AMERICAN Volume 92 FERN Number J O U R N A q January—March 2002 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Obituary: Rolla Milton Tryon, Jr. (1916-2001) Gerald J. Gastony, David S. Barrington, and David S. Conant 1 a Associated with the Intertracheid Pit seas) of the Woody Fern Botrychium multifidum ela C. Morrow and Roland R. Dute 10 A New Filmy Fern from the Dominican Republic Carlos Sanchez 20 Adiantum argutum, an Unrecognized Species of the A. latifolium Group Jefferson Prado and David B. Lellinger 23 Polypodium cseapage Plants Sporulate viacpecsndaeed in a Non-seasonal Glasshouse Envi- ronmen a E. Simdn and Elizabeth Sheffield 30 The American Fern Society Council for 2002 CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS 66045-2016. President TOM RANKER, University Museum, Campus Box 265, University of Colorado, Boulder, art eu ‘65. President W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI eg 1478. Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916-1110. Treasurer GEORGE YATSKIEVYCH, Missouri Botanical Garden, P. O. Box 200, St. oe MO 63166-0299. rship Secretary JAMES D. MONTGOMERY, Ecology HI, R.D. 1, Box 1795, Berwick, PA 18603-9801. Back — Curator R. JAMES HICKEY, Botany Dept., Miami University, Oxford, OH 45056. rnal Editor DAVID B. LELLINGER, U.S. National Herbarium MRC-166 Smithsonian Institution, Washington, DC 20560-0166. Memoir Editor CINDY JOHNSON-GROH, Dept. ‘of pom — Adolphus College, 800 W. College Ave., St. Peter, MN 56082 Bulletin Editor American Fern Journal EDITOR R. JAMES HICKEY Botany De ami University, Oxford, OH 45056 ph. (513) £5 Sone e-mail: hickeyrj @ muohio.edu ASSOCIATE EDITORS GERALD J. GASTONY .............. Dept. of Biology, Indiana University, ages aig IN 47405-6801 CHRISTOPHER H. HAUFLER ...... Dept. of Botan get University of Kansa 66045-2106 ROBBIN C. MORAN ew York Botanical Gecten: Bike NY 10458-5126 JAMES H. PECK ser of — University of Arkansas—Little Rock, 2801 S. University Ave., Litthe Rock, AR 72204 The “American Fern Journal” (ISSN 0002- inapls is an illustrated quarter! iene a the sf acting study of ferns. It is owned by the American Fern Society, and ame at ate st St VA 22180-4151. Periodicals postage paid at Vienna, VA, and addition — for bare: issues, made 6 months (domestic) to 12 sani sei ‘cael after the date of issue, and r back issues should be addressed to Dr. James D. Montgomery, Ecology Il, R.D. 1, Berwick ies 18603. 9801. Chan of address, dues, and applications for membership should be sent to the Membership Secretary. General inquiries posianeag ferns should be addressed to the Secretary. Subscriptions $20.00 gross, $19.50 net if paid through an agency (agency fee $0.50); sent free to members of the prorestians Fern Soaew. (annual dues, $15.00 + wecbe mailing surcharge beyond U.S.A.; life oni ne rship, $300.00 + $140.00 mailing surcharge beyond U.S.A. Back volumes are available for most years as printed issues or on microfiche. Please contact the Back Issues Curator for prices and availability. POsTMASTER: Send address changes to AMERICAN FERN JOURNAL, 326 West St. NW., Vienna, VA 22180-4151. FIDDLEHEAD FORUM The editor - the Bulletin of the American Fern Society welcomes contributions from members and non-members, including miscellaneous notes, offers to ee or purchase materials, personalia, horticultural sae and reviews of non-technical books on SPORE EXCHANGE Mr. Stephen McDaniel, 1716 Piermont Dr., Hacienda Hts., CA 91745-3678, is Director. pote exchanged and lists of available spores sent on request. http://www.amerfernsoc.org/sporexy GIFTS AND BEQUESTS Gifts and dais ssn ps to ia Society enable it to expend it its services to members and to others interested in ferns. Ba other gifts are always welcomed and are tax-deductible. Tnonin lA hy - > thn < me 4) te OCCretary American Fern Journal 92(1):1—9 (2002) MISSoy Ay Obituary: Rolla Milton Tryon, Jr. g4p 1 4 2099 (1916-2001) DEN LiBRap GERALD J. GASTONY Department of Biology, Indiana University, Bloomington, IN 47405-3700 Davip S. BARRINGTON Department of Botany, University of Vermont, Burlington, VT 05404-0086 Davip S. CONANT Department of Natural Sciences, Lyndon State College, Lyndonville, VT 05851 Rolla Tryon, a member of the American Fern Society since 1932 and one of the twentieth century’s most eminent students of pteridophytes, was born on August 26, 1916 in Chicago, Illinois. His father, a professor of American history and education at the University of Chicago, maintained a summer cottage in Chesterton, Indiana in addition to his home in Chicago. Rolla’s fascination with ferns and fern allies developed during boyhood forays from that Ches- The photograph was taken by Dr. Walter H. Hodge in Mexico City in December, 1972 and was made available by the Hunt Institute for Botanical Documentation. 2 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) terton cottage into sand dune habitats along Lake Michigan in the northwest of Indiana. At the age of 18, he published his first paper, relating his obser- vations on Osmunda plants in the Indiana Dunes (see complete bibliography below). As a boy, Rolla was greatly influenced by, and in turn influenced, Charles Deam (author of the 1940 Flora of Indiana), advising Deam about fern species he had found in the dunes area. When a doubting Deam appeared at the cottage door one day asking to meet Rolla and to be shown these ferns in situ, he was surprised to learn that Rolla was not the adult of the family but a mere boy of 14. Thus began a productive friendship documented in Rolla’s correspondence with Deam from May, 1935 to January, 1953. All the penny postcards and letters he received from Deam have been carefully maintained in one of Rolla’s files, now archived at Indiana University—fascinating reading. Rolla’s insatiable boyhood appetite for ferns got him into a bit of trouble at home, however, when his father learned that he had charged Bower's three volumes on The Ferns to his account at Brentano’s bookstore in Chicago. Rolla built a solid academic superstructure on the foundation of these boy- hood experiences. Among his academic accomplishments were an A.A. degree in 1935 and a B.S. degree in 1937, both from the University of Chicago, and a Ph.M. in 1938 from the University of Wisconsin. In 1940 he earned an MS. and in 1941 a Ph.D., both from Harvard University. During his days as a Har- vard student, he contracted malaria in South Carolina while collecting plants for M. L. Fernald, and during the war-torn year following completion of his Ph.D. he served as a lab technician in the U. S. Chemical Warfare Service at Massachusetts Institute of Technology. His father thought he should follow his Ph.D. in botany with another, this time in chemistry, so that he could earn a living, but instead Rolla became an Instructor in Botany first at Dartmouth College, then at the University of Wisconsin before becoming an Assistant Professor in Botany at the University of Minnesota in 1945. While an Assistant Professor at the University of Wisconsin, Rolla met Alice Faber. Their marriage in 1945 initiated not only a happy and enduring domestic partnership but also a research synergism whose productivity has nourished pteridologists through- out the world. In 1947 Rolla became Associate Professor in Botany at Washington Univer- sity, St. Louis and Assistant Curator of the herbarium of the Missouri Botanical Garden, positions he held to 1957. During this appointment, he and Alice were the original organizers of the Missouri Botanical Garden’s annual Systematics Symposium, whose 48th meeting was held 12-13 October, 2001. This highly successful annual meeting has received continuous support from the National Science Foundation from its second year (1954) to the present (with the lapse of a single year). From 1946 to 1957 Rolla served as curator and librarian of the American Fern Society’s library and herbarium, responding to members’ requests for loans of materials. That herbarium and library was subsequently entrusted to Warren H. Wagner at the University of Michigan. Following a year as Research Associate at the University of California, Berkeley, Rolla went to the Gray Herbarium of Harvard University as Associate Curator and Curator of Ferns in 1958 and became Curator of the Gray Herbarium in 1967. Rolla ROLLA MILTON TRYON, JR. 3 and Alice traveled the world extensively, attending international meetings, conducting field work, studying specimens at major herbaria in the Americas, Europe, and Africa, and conducting field courses on the ferns. In addition to other services to professional societies, Rolla served for many years as Asso- ciate Editor of Rhodora and the American Fern Journal, as Associate Editor of Brittonia (1961-1964), as Editor-in-Chief of Rhodora (1977-1982), and as Pres- ident of the New England Botanical Club and the American Fern Society. A framed photograph of his revered mentor, Charles A. Weatherby (see American Fern Journal 40[1] for a remarkable series of papers honoring this unusually respected and beloved botanist), was always prominently displayed on Rolla’s desk at Harvard, undoubtedly inspiring his own welcoming, patient, and supportive response to all who entered his office seeking counsel. In 1970 Rolla initiated an annual New England Fern Conference at Harvard Forest. For 20 years this provided a stimulating intellectual setting in which students of fern biology discussed and developed their ideas. In 1972 he became Professor of Biology at Harvard University, holding both the Curatorship and Professor- ship until his retirement in 1987. He remained at Harvard as Professor Emer- itus from 1987 to 1989 when he moved to the University of South Florida in Tampa as Adjunct Professor, bringing with him his extensive library of fern and biogeographic literature. To mark the retirements of Alice and Rolla Tryon from Harvard, a festschrift of 13 papers plus introduction was published in their honor in the Annals of the Missouri Botanical Garden (vol. 77: 225-339. 1990). At the University of South Florida, Rolla and Alice helped found the Insti- tute for Systematic Botany and endowed the Tryon Lecture Series that brings several internationally known botanists to the university each year. In their research-active office on the Tampa campus, he and Alice continued their pter- idological work, as his following bibliography indicates. Rolla Tryon’s publication list exceeds 100 titles and includes a great breadth of topics. Papers ranged from articles on pteridophytes for the 1943 Encyclo- pedia Britannica to a glossary of terms relating to the fern leaf, discussions of the history of pteridology and fern classification, a remembrance of his grad- uate mentor and counselor Charles A. Weatherby, discussions of the formali- ties of fern nomenclature, and many book reviews. His monographs and re- visions focused mostly on ferns but also included angiosperms Convolvulus and Elymus. Signal among these were his revisions of Pteridium, Doryopteris, the Selaginella rupestris group, American Notholaena, and the Cyatheaceae. His papers on fern biogeography began with Doryopteris in 1944, matured in his exposition of geographic speciation in Selaginella in 1971, and continued to his and Alice’s 1999 discussion of the phytogeography of eastern North American ferns (honoring Ching Ren-Chang). Floristic and taxonomic notes on ferns ranged from simple observations of growth forms and hybrids to eluci- dations of complex taxonomic and nomenclatural issues. For his 1955 publi- cation on the taxonomy of cycads (coauthored with students in his Washington University class) he was awarded the 1956 Robert Montgomery award of the Fairchild Tropical Garden for distinguished achievement in the world of palms 4 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) and cycads. At the time of his death, he had a book review in press in Rhodora and a paper in press in Bradea (coauthored with his former student Paulo Windisch). In addition to his numerous papers on ferns and other topics, Rolla is no- table for his books. Among these are two editions of his Ferns and Fern Allies of Wisconsin (1940, 1953) and Ferns and Fern Allies of Minnesota (1954, 1980). He is renowned for his knowledge of the ferns of Peru, first expressed in his 1964 Ferns of Peru (250 pages in the Contributions from the Gray Her- barium of Harvard University). This treatment was updated and completed in six subsequent parts entitled Pteridophyta of Peru between 1989 and 1994, mostly coauthored with Robert Stolze but with several portions contributed by other pteridological specialists. His monumental 1982 book Ferns and AIl- lied Plants with Special Reference to Tropical America, coauthored with Alice Tryon, is an encyclopedic treatment of this subject that continues to stimulate new research, as does his treatment of Pteridaceae, with Alice Tryon and Karl Kramer, in volume 1 of Families and Genera of Vascular Plants edited by K. Kramer and P. S. Green. Rolla’s kindly and perceptive mentoring and his outstanding contributions to our knowledge of ferns is signaled by having the following four fern taxa named in his honor. 1) Asplenium tryonii Correll. In describing this species, Donovan Correll (1961) said “It is a pleasure to name this species for Dr. Tryon, who has always been most gracious in helping his fellow-workers with their never- ending problems in the study of ferns.” Known only from Chihuahua, Mex- ico, this species was further discussed and illustrated in Ferns and Fern Allies of Chihuahua by Knobloch and Correll (1962). Alsophila tryonorum Riba. The eminent Mexican pteridologist Ram6n Riba (see American Fern Journal 90:112—118, 2000) stated that “this species is named after Dr. Rolla M. Tryon and Dr. Alice F. Tryon for their contributions to the taxonomy of the ferns” (Riba, 1967). The plural specific epithet rec- ognizes the close professional relationship between this highly productive research team. This tree fern species is now known as Trichipteris tryono- rum (Riba) R. Tryon following its transfer by Rolla in his 1970 paper on the classification of the Cyatheaceae. Nephelea tryoniana Gastony. ‘“‘I am pleased to name this species for my mentor, Dr. Rolla M. Tryon, in recognition of his outstanding contribution to the understanding of the systematics and evolution of the family Cy- atheaceae” (Gastony, 1973). Subsequent research by Conant (Conant and Cooper-Driver, 1980; Conant, 1983) revealed that this tree fern species is a reproductively stabilized diploid hybrid species that is now regarded as Alsophila tryoniana (Gastony) Conant. Tryonella Pichi Sermolli. This new generic name was established by Pichi Sermolli (1974) “in honour of the eminent pteridologist R. M. Tryon, Jr., author of many important papers on ferns, who, inter alia, supported the distinction of the present genus from Doryopteris, though without giving it i) — ioe) — ns — ROLLA MILTON TRYON, JR. 5 anew name.” This name is currently regarded as a synonym of Doryopteris by Tryon and Tryon (1982) and Tryon, Tryon, and Kramer (1990). Among the doctoral graduate students he trained, Rolla counted the follow- ing (those with asterisks received their degrees from other institutions): Alice F. Tryon, *Karl Kramer, *Ramon Riba, Gerald Gastony, Lawrence Palkovic, David Barrington, David Conant, Paulo Windisch, *R. James Hickey, *Robbin Moran, Sonia Sultan, and Calvin Sperling. He also mentored Robert Stolze in his taxonomic revision of Cnemidaria at the Field Museum. Always available to his students, he modeled his supportive and insightful mentoring on his experiences with his own graduate mentor, Charles Weatherby. For this he has earned our love as well as our respect. His impact on his students, and their students, and their students is incalculable. In 1978 Rolla M. Tryon, Jr. was elected to honorary membership in the Amer- ican Fern Society, a special category of membership for persons who have made outstanding contributions to the study of ferns. In 1984 he received a Merit Award from the Botanical Society of America “In recognition of distin- guished achievement in and contributions to the advancement of botanical science. Pre-eminently knowledgeable in matters of taxonomy and nomencla- ture, this foremost pteridologist is a perceptive student of phytogeography and of the evolutionary impact of the selective process during plant migration.” In addition to Rolla’s botanical activities he was also highly skilled in run- ning a family farm in Knox County, Indiana for many years. He visited the farm, overseeing its management, a few times each year. The extensive records he kept in managing crops and livestock illustrate his practical ability in man- aging business as well as scientific data. On August 20, 2001, six days before his eighty-fifth birthday, Rolla Milton Tryon, Jr., left us to continue our work with the pteridophytes of the world, and to delight in them, without him. We do this fortified by his writings, the echoes of his encouraging words, and his everlasting example. He was the beloved husband of Alice Faber Tryon, the benefactor of countless students of pteridophytes, including many who never knew him personally, an inspiration and counselor to many collaborators and coauthors, the advisor of doctoral students, the teacher of innumerable undergraduates, and our dear friend and mentor. He will be deeply missed. He already is. LITERATURE CITED CONANT, D. S. 1983. sis revision of the genus Alsophila (Cyatheaceae) in the Americas. J. Arnold Arbor. 64: 333-— Conant, D. S. and : ones DRIVER. 1980. Autogamous allohomoploidy in Alsophila and Ne- phelea (Cyatheaceae): a new hypothesis for speciation in homoploid homosporous ferns. mer. J. Bot. 67: 1269-1288. CorreLL, D. S. 1961. Two Texas-Chihuahuan ferns. Wrightia 2: 200-203. — G. J. 1973. A revision of the fern genus Nephelea. Contr. Gray Herb. 203: 81— eure _ I. W. and D. S. CorreLL. 1962. Ferns and fern allies of Chihuahua, pace — a Foundation, Renner, Tex PICHI pameene R. E. G. 1974. seit a Webbia 29: 1-16. 6 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) Risa, R. 1967. New taxa in the genus Alsophila. Rhodora 69: 65-68. BIBLIOGRAPHY OF ROLLA M. TRYON, JR. (1916-2001) TRYON, R. M., JR. 1934. Some observations on Osmundas. Amer. Fern J. 24: TRYON, R. M., JR. 1936. Ferns of the dune region of Indiana. Amer. Mid. en 17: ite TRYON, R. M., JR. 1936. Botrychium dissecturn and forma obliquum. Amer. sg n J. TRYON, R. M., JR. 1938. The phenomenon of forking in ferns. Amer. Fern J. 2 TRYON, R. M., JR. 1938. Recent additions to the flora of Indiana. Proc. “la poe Sci. 47: 76— 17 TRYON, R. M., JR. 1939. Notes on the ferns of Wisconsin. Amer. Fern J. 2 9, TRYON, R. M., JR. 1939. The varieties of Convolvulus spithamaeus and a C. sepium. Rhodora 41: 5—423. TRYON, R. M., JR. 1940. Notes on some Indiana plants. Proc. Indiana Acad. Sci. 49: 89-90. TRYON, R. M., JR., N. C. Fassett, D. W. DUNLOP, and M. E. 6 EMER. 1940. The ferns and fern allies of Sip — Dept., Univ. of Wisconsin, Madiso TRYON, R. M., JR. 0. An Osmunda hybrid. Amer. Fern J. 30: 65-66. TRYON, R. M., JR. 1 mi a revision of the genus Pteridium. Bhai 43: 1-31, 37-67. Reprinted as Contr. Gray Herb. 1 -7 bia R. M., JR. 1942. ae Roland, A. E., The ferns of Nova Scotia. Amer. Fern J. 32: 73— Perey R. M., JR. 1942. A new Dryopteris hybrid. Amer. Fern J. 32: 81— TRYON, R. M., JR. 1942. A revision of the genus Doryopteris. Contr. G ae 80. TRYON, R. M., JR. 1942. Review of Brown, C. A. and D. S. Correll, The ferns ak 7e allies of Louisiana. Rhodora 44: 484—485. sia TRYON, R. M., JR. 1943. Several articles on Pteridophyta for Encyclopedia Brittanic TRYON, R. M., JR. ma Abstracts of the American Fern Jour nal for Biological Abstracts. TRYON, R. M., JR. and J. W. Moore. . Notes on aquatic and prairie acca in ane i 6. TRYON, R. M., JR. and J. W. Moore. 1946. A preliminary checklist of the flowering plants, ferns and fern allies of Minnesota. Univ. of Minnesota, Minneapolis. TRYON, R. M., JR. 1946. A new Doryopteris hybrid. Amer. Fern J. 36: 4 OorE, J. W. and R. M. TRYON, JR. 1946. A new record of Isoétes inalsinpod. Amer. Fern J. 36: 89-91 vista F, K and R. M. TRYON, JR. 1948. A fertile mutant of a Woodsia hybrid. Amer. Fern J. 35: 13 TRYON, M., JR. 1948. Some Woodsias from the north shore of Lake Superior. Amer. Fern J. 38: 159-170. Boouer, L. E. and R. M. TRYON, JR. 1948. A study of Elymus in Minnesota. Rhodora 50: 80-91. TRYON, R. M., JR. 1950. Charles Alfred Weatherby—teacher and counselor. Amer. Fern J. 40: 9-10. TRYON, R. M., JR. 1950. A new erect species of the Selaginella rupestris group. Amer. Fern J. 40: 69-74. TRYON, R. M., JR. — Ferns of the Missouri Ozark region. Missouri Bot. Gard. Bull. 39: 136-138. TRYON, R. M., JR. 1951. Review of Manton, I., Problems of cytology and evolution in the Pterido- phyta. Ecology 32: 769-770. TRYON, R. M., JR. 1952. A sketch of the history of fern classification. Ann. Missouri. Bot. Gard. 39: 6 TRYON, R. M., JR., N. C. Fassett, D. W. DUNLOP and M. E. DiEMER. 1953. The ferns and fern allies of Wisconsin, Ed. 2. Univ. Wisconsin Press, Madison. TRYON, R. M., JR. 1954. The ferns and fern allies of Minnesota. Univ. Minnesota Press, Minneap- olis TRYON, R. M., JR, E. BARBOUR, E. Davis, H. Kipp, R. Lonc, C. Marvin, B. MIKULA, and R. MOHLEN- BROCK. 1955. Ancient seed plants: the cycads. Missouri. Bot. Gard. Bull. 43: 65-80. ROLLA MILTON TRYON, JR. TRYON, iy M., re ney ep ae ager vee its allies. Ann. Missouri. Bot. Gard. 42: 1-99. KRAME N, JR. 1955. eeu. new species of Doryopteris from Surinam. Ann. Missouri Bot a < _ ner oe Sti . — JR. 1956. A revision of the American species of Notholaena. Contr. Gray Herb. 179. aa. . a JR. 1957. Adianturn in Peru: new species and combinations. Amer. Fern J. 47: 139— ah KosuskI, C. E., C. V. MorTON, M. Ownsey, and R. M. TRYON. 1958. Report of the committee for recommendations on desirable procedures in herbarium practice and ethics. Brittonia 10: 93-95. TRYON, R. M., JR. and A. F. TRYON. 1959. Observations on cultivated ferns: the hardy species of tree ferns (cha and Cyatheaceae). Amer. Fern. J. 49: 129-142. Reprinted in Lasca Leaves 10: 26-33. 1960. TRYON, R. M., Ir. 1960. The ecology of Peruvian ferns. Amer. Fern J. 50: 4 TRYON, R. 1960. A glossary of some terms relating to the en — Taxon c 108-109, Reprinted in Russian iio in Bot. Jour. Akad. Nauk, U.S.S 736-739. TRYON, M., JR. 1960. Review of Pteridophyta in Munz, P. a o balou rene Amer. Fern J. 50: seit TRYON, R M. JR. 1960. A review of the genus Dennstaedtia in America. Contr. Gray Herb. 187: 23— TRYON, * 4 JR. 1960. New species of ferns from serait mei South America. Rhodora 62: 1-10. TRYON, R. M., JR. 1961. Taxonomic fern notes. I. Rhodora 63: 7-88 TRYON, R. M., JR. 1962. The fern genus Doryopteris in aa ee and Rio Grande do Sul, Brazil. Sellowia 14: 51-59 TRYON, R. M., JR. 1962, Review af Wherry, E. T., The Fern Guide. Amer. Fern J. 52: 89-91. TRYON, R. M., JR. 1962. A note on Nephrolepis pa ie cv. Duffii. Amer. Fern J. = 153-155. TRYON, R. M., JR. 1962. Taxonomic Fern Notes. II. Pityrogramma (including Trismeria) and Ano- gramma. Contr. Gray Herb. 189. 52-76. TRYON, R. 1962. Taxonomic fern notes. III. Contr. Gray Herb. 191: 91— TRYON, R. M., JR. 1962. Review of Knobloch, I. W. and D. S. Sessil ow and Fern Allies of Chihuahua. Rhodora 64: 347-348 TRYON, R. 1962. A commentary o axpeetiudus names. Taxon 11: 116—120. TRYON, R. 1963. Nomenclatural er a Taxon 12: 28-285. TRYON, R. M. 1964. Evolution in the leaf of living ferns. Mem. Torrey Bot. Club 21: 73-85. TRYON, R. 1964. erie ferns of Peru. Polypodiaceae (Dennstaedtieae to Oleandreae). Contr. Gray Herb. 1 194: 1-— TRYON, R. 1964. rite ie Fern Notes IV. Rhodora 66: 110—117. PICHI-SERMOLLI, R. E. G., F. BALLARD, R. E. HOLTTuM, H. IT, F. M. JARRETT, A. C. JERMY, E. A. C. L. E. SCHELPE, M.-L. TARDIEU-BLOT, and R. M. TRYON. 1965. Index Filicum Supplementum Quartum pro Annis 1934-1960. Internat. Bureau for Plant Taxonomy and Nomenclature, Utrecht, Netherlands. TRYON, R. 1965. rig. te in the ferns. Taxon 14: 213-218. TRYON, R. and D. M. BRITTON. 1966. A study of variation in the cytotypes of Dryopteris spinul Rhodora 68: 59-92. TRYON, R. 1967. Taxonomic fern notes V. New combinations in Peruvian species of Thelypteris. odora 69: 5- TRYON, R. 1967. paviine of Hara, H., The flora of eastern Himalaya. Rhodora 69: 456-457. Boum, B. A. and R. M. TRYON. 1967. Phenolic compounds in ferns. I. A survey of some ferns for c TRYON, R. M. and A. F. TRYON. 1968. Edith Scamman kage 1967). Amer. Feri 5 58: 1-4. TRYON, R. 1968. Nomenclatural proposals. Taxon 17: 588-590. TRYON, R. 1969. Pteridology, Pp. 97-102 in J. Ewan (ed.) A short history of botany in the United States. sone Publishing Co., New York. TRYON, R. 1969. Taxonomic problems in the geography of North American ferns. BioScience 19: iat 8 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) paige R. 1970. aes of Correll, D. S. and M. C. Johnston., Manual of the vascular plants of xas. Rhodora 72: 533-535. Pines R. 1970. The le of the Cyatheaceae. Contr. Gray Herb. 200: 3-53. TrYON, R. 1970. Development and evolution of fern floras of oceanic islands. Biotropica 2: 76-84. Reprinted 1971 as Pp. 54-62 in W. L. Stern (ed.) Adaptive aspects of insular evolution. Wash TRYON, R. 1971. The American at ferns allied to i trata seeps ang 73: 1-19. TRYON, R. 1971. Ferns of the Andes and Amazon. Morris Arbor. Bull. 2 TrYON, R. 1971. The process of evolutionary migration in species oe sae Brittonia 23: 89— 100. TRYON, : 1972. Endemic areas and geographic speciation in tropical American ferns. Biotropica 4: 121-131. TRYON, s 1972. Taxonomic fern notes, VI—New species of American Cyatheaceae. Rhodora 74: TRYON, R. M., Jr. and A. F. TRYON. 1973. spac 7 - evolutionary relations in the cheilanthoid ferns. Bot. J. Linn. Soc. 67: Suppl. 1: TRYON, A. and R. TRYON. 1973. Thelypteris in se ae ee America. Amer. Fern J. 63: 65— 76 TRYON, R., B. VOELLER, A. TRYON, and R. RIBA. 1973. Fern biology in Mexico (a cl field program). BioScience 23: 28-33. TRYON, R., C.-J. WIDEN, A. HUHTIKANGAS, and M. LOUNASMAA. 1973. pony nana derivatives in Dryopteris parallelogram and D. patula. Phytochemistry 12: TRYON, A. and R. TRY 974. Ge ogaph — in temperate rence fant and some rela- tionships in Galas Amer. Fern. J. 64: 99-104. TRYON, R. 1975. The Benjamin D. G Gilbert fern nea Amer. Fern J. 65: 60. TRYON, R. 7 and G. J. GAsToNny. 1975. The biogeography of endemism in the Cyatheaceae. Fern Gaz 573-79. TRYON, R. 4 and D. S. CONANT. 1975. The ferns of Brazilian Amazonia. re paiitias 5: 23-34. TRYON, R. 1976. A oa of the genus Cyathea. Contr. Gray Herb. 206: GASTONY, G. J. a . M. TRYON. 1976. Spore morphology in the Cyatheacoe Il. nae genera Lophosoria, soe Sphaeropteris, Alsophila, and Nephelea. Amer. J. Bot. 63: 738-758. aug R. M. 1977. Studies on the American Cyatheaceae and on the pie ees of on family. Males. Bull. 30: 2839— ete R M. and G. VITALE. 1977. dua for antheridogen production and its mediation of a mating system in natural populations of fern gametophytes. Bot. J. Linn. Soc. 74: 243-249. TRYON, R. 1979. Biogeography of the Antillean fern flora. Pp. 55-68 in D. Bramwell (ed.) Plants and islands. Academic Press, New York. TRYON, A. F., R. TRYON, and F. BADRE. 1980. Classification, spores, and nomenclature of the marsh fern. Rhodora 82: 461-474 TRYON, R. 1980. Review of Sepensnck, J. A. and O. Huber, Flora of Avila. Amer. Fern J. 70: 79. Tryon, R. M., JR. 1980. The ferns and fern allies of Minnesota, Ed. 2. Univ. Minnesota Press, Minneapolis. wer - ft A. TRYON. 1981. Taxonomic and nomenclatural notes on ferns. Rhodora 83: 133-— wee - 1982. — “e a A. R., Pteridophytes in D. E. Breedlove (ed.) Flora of Chiapas Part 2. Madrofio TRYON, R. pane ALF. aly a Additional taxonomic and nomenclatural notes on ferns. Rho- TRYON, R. M. and A. F. TRYON. 1982. Ferns and allied plants with special reference to tropical America. Springer-Verlag, New TRYON, R. 1984. An unusual new Bapheeyplesenctin from Peru. Amer. Fern J. 74: 108-110. TRYON, R. 1985. Fern speciation and biogeography. Proc. Roy. Soc. Edinburgh, B. 86: 353-360. TRYON, R. 1986. Dicksoniaceae. Lophosoriaceae. Metaxyaceae. Cyatheaceae. Pp. 1-59 in G. Harling and L. Andersson (eds.) Flora of Ecuador, No. 27. Swedish Research Councils, Stockholm. TRYON, R. 1986. Some new names and combinations in Pteridaceae. Amer. Fern J. 76: 184—186. ROLLA MILTON TRYON, JR. 9 TRYON, R. nip — biogeography of species, with special reference to ferns. Bot. Rev. (Lancaster) S2otks— TRYON, R. i pee Pp. 327-338 in H. Lieth and M. J. A. Werger (eds.) Tropical rain forest ecosystems. let Amster TRYON, R. M. and R. G. STOLZE. 1989. Pied eae of Peru. Part I. 1. Ophioglossaceae—12. Cy- eee. Fieldiana, Bot ‘New eries 20: 1-145. Tryon, R. M. and R. G. STozE. 1989. Pteridophyta of Peru. Part II. 13. Pteridaceae—15. Dennstaed- tiaceae. Fieldiana, Bot. New Series 22: 1-128. TRYON, R. M., A. F. TRYON, and K. U. KRAMER. 1990. nbpigh Sues Pp: 230-256, in K. Kubitzki (ed.) ee eee and genera a a plants. Vol. 1. oo and gymnosperms. Vol. K. U. Kramer and P. S. Green. Springer-Verlag, a6 as RM and R. G. SToLzE. 1991. Pteridophyta of Peru. a IV. 17. Dryopteridaceae (with collaboration of] T. Mickel and R. CG. Moran). Fieldiana, Bot. New Series 27: 1-176. TrYON, R. M. and R. G. STOLZE. 1992. Pteridophyta of oF has Ill. 16. Thelypteridaceae (con- tributed ‘ne A. R. Sensi Fieldiana, Bot. New Series 2 TRYON, R. M. and R. G. STOLZE. 1993. Pteridophyta of a < V1 ie ariaernepRi Poly- podiaceae (with gepacieiers of B. Leén). Fieldiana, Bot. New Series —190. TRYON, R. M. and R. G. STOLZ 1994. Pteridophyta of Peru. Part VI. 22. spe akan sir (with collaboration of = \.H ckey and B. Ollgaard). ne apr Bot. New Series 34: 1— TRYON, R. 1997. Systematic ee on Oleandra. Rhodora 9 -3 TRYON, R. 1997. Proposal to reject the name Acrostichurn a a (Pardaceas), Taxon 46: 339— 340 CHURCHILL, H., R. TrYON, and D. S. BARRINGTON. 1998. Development of the sorus in tree ferns: Dicksoniaceae. Canad. J. Bot. ne 1245-1252. TRYON, R. M. and A. F. TRYON. 1999. Observations on the phytogeography of eastern North Amer- ican ferns. Pp 250-273 in X-C Zang and K-H Shing (eds.). Ching Memorial Volume. Institute of Botany, Chinese Acad. Sci., Beijin TRYON, R. 2000. Systematic notes on the Old World fern genus Oleandra. Rhodora 102: 428-438. TRYON, R. M. 2001. Review of Hoshizaki, B. J. and R. C. Moran, Fern grower's manual, revised and expanded edition. ego 103: 34 Winoiscu, P. G. and R. TRYON. 2001. The pane Ricardo Franco (State i “is Grosso, Brazil) as probable migration ean and its present fern flora. Bradea 8: 267— American Fern Journal 92(1):10—-19 (2002) Crystals Associated with the Intertracheid Pit Membrane of the Woody Fern Botrychium multifidum ANGELA C. MORROW AND ROLAND R. DUTE Department of Biological Sciences and Alabama Agricultural Experiment Station, Auburn University, Auburn, Al 36849, USA ABSTRACT.—CALCIUM-containing crystals have been found in the lumens of secondary tracheids in the rhizome of the woody fern Botrychium multifidum. These crystals are styloids with rough, pyramid- shaped ends. The crystals are usually single; however, conjoined or grouped crystals were also found. Crystal formation apparently has no constant relation to the pit membrane, but crystals of mature tracheids are often associated with the pit membrane or are located in the pit areas. Crystals were also located between the helical thickenings of the lumen walls. No crystal chamber or crystal sheath was found in association with the crystal body. Crystals are a common feature in many plant tissues (Scurfield and Mitchell, 1973), and more than 1000 crystal producing woody plants, spanning 160 fam- ilies, were described at the light microscopic level by Chattaway (1955, 1956). Scanning electron microscopy has allowed for more rapid identification of crystals in plant tissues and a clearer picture of their morphology (Scurfield and Mitchell, 1973). Although crystals in xylem tissue have been reported in the vessels of Intsia Thouars (Fabaceae; Hillis, 1996), Torreya yunnanesis C.Y, Cheng & L.K. Fu (Taxaceae; Kondo et al., 1996), and Polyalthia Blume (An- nonaceae; Scurfield and Mitchell, 1973), they are most commonly found in the xylem parenchyma, septate fibers, or vessel tyloses (Scurfield and Mitchell, 1973). The formation of crystals by Botrychium, the only extant fern that produces wood (Gifford and Foster, 1989), has not been previously reported. During our studies of the torus-bearing pit membrane in the tracheid of Botrychium mul- tifidum (S.G. Gmelin) Rupr., we discovered occasional instances of crystals associated with the pit membrane. This paper describes the morphology of these crystals as observed with SEM. The discoverey of these very small crys- tals was unexpected, and our exploration of them to this point has been strictly descriptive. However, in our discussion we explore several possible reasons for crystal formation in this wood. MATERIALS AND METHODS Rhizome samples of upright or orthotropous rhizomes of Botrychium mul- tifidum. were collected by Dr. D. W. Stevenson (New York Botanical Garden, New York City, U.S.A.) from Plumas County, California and fixed in FPA. The collection site (elevation 2000 m) is rocky mountain soil at the edge of a mead- ow and the ground is frozen for much of the year. The samples were typical rhizomes selected as random samples and representative of the population. In MORROW AND DUTE: CRYSTALS IN BOTRYCHIUM MULTIFIDUM 11 our lab, samples were cut transversely into 1-2 mm pieces that were placed into 50% ethanol and then dehydrated through a graded alcohol series. Sam- ples then were cut into small wedges, placed into hexamethyldisilazane (HMDS) for 2 hours (Nation, 1983), and subsequently placed under a chemical hood overnight to dry. Dry samples were attached to aluminum stubs with double-sided sticky tape and coated with gold-palladium. For comparison pur- poses, samples of Botrychium dissectum Sprengel and B. virginianum (L.) Swartz from Lee County, Alabama were prepared in the same manner as B.multifidum. Specimens were viewed with a Zeiss DSM 940 at 5,10, or 15 kV. Qualitative element identification was performed using energy dispersive spectroscopy (Tracor Northern Micro Z II) coupled to the SEM. RESULTS Secondary xylem tracheids of Botrychium multifidum contain helical wall thickenings and intertracheid circular bordered pits (Fig. 1). Thickenings, as seen in longitudinal section, are uniform neither in height nor in distance between gyres, and thickenings are sometimes branched (Fig. 1). The pit mem- brane is almost always differentiated into a torus and margo (Fig. 2). Microfi- brils of the pit membrane are loosely woven in the margo region, but tightly woven in the torus. Tearing of the pit membrane was sometimes evident in the margo (Fig. 2). Crystals were found in association with torus-bearing pit membranes of tracheids (Fig. 11), as well as in tracheid lumen (Figs. 1, 3). These crystals were not apparent at the light level. Crystals associated with these tracheids are styloids (Frey-Wyssling, 1981; Carlquist, 1988); they are rectangular columnar with pyramidal ends. Intact crystals have columns that are four-sided and are smooth-surfaced. The pyramidal crystal ends consist of four equilateral triangles, although wedge-shaped ends also were observed (Fig. 4). Crystal ends, when visible, typically appeared to be rough (Fig. 3), although some crystals with smooth ends were observed (Fig. 4). Crystals ranged in size from 4.3 to 12 wm in length, and 1.14 to 2.4 pm in width (N = 12). The mean crystal length is 7.27 »m, mean width is 1.55 pm, and mean ration of width-to-length is 1: 4.7. The crystals were not always regular in shape and sometimes appeared to have their growth modified by the presence of a helical wall thickening (Fig. 3). By energy dispersive spectroscopy (EDS), these crystals were found to be composed of a calcium compound, most likely calcium oxalate (Fig. 13). Although single isolated crystals were most commonly found, joined double crystals with U-shaped conjoined end were also observed (Fig. 4). Crystals in groups of two or more were also encountered and had either parallel or per- pendicular orientation to each other (Figs. 1,5). Crystals were found in various positions within the tracheary lumen. They were located between either the helical thickenings of the wall material, laying flat on the inner cell wall, or projecting out from the pit membrane (Fig. 1). It was clear from some specimens that the crystals were composite structures (Figs. 4, 6-8). In some instances the subunits resembled raphides (Figs. 6,7), 12 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) Fics. 14. SEM micrographs of intertracheary pit membranes and crystals. 1) Tracheary lumen with helical wall thickenings (W), crystals (arrow), pit aperture (A), and circular bordered pit membrane (P); scale bar = 5 um. 2) Intertracheary pit membrane with torus (T) a margo (M). e pit border was removed when the wood was split during preparation; scale = 2 pm. 3) Crystal entering a pit aperture (A). Note how = crystal appears to have grown tsa the helical thickening to the right (double arrow). R = rough end of crystal; scale bar = 2 um. 4) Double crystal joined at one end (arrow); scale bar = m. MORROW AND DUTE: CRYSTALS IN BOTRYCHIUM MULTIFIDUM FIGs SEM micrographs of crystals in which crystals revea Multiple crystals with parallel orientation and perpendicular orientation. Note subunits crystal (arrow); scale bar = 5 pm. 6) E | their subunit composition. 5) s in broken view of composite crystal formed by smaller raphide shaped crystals (arrow); scale bar = 500 nm. 7) Composite crystal w ith styloid (arrow) and raphide crystal (double arrow) shaped subunite: scale bar = 2 pm. 8) Composite crystal with styloid crystal subunits (arrow); scale bar = 2 pm. 14 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) whereas in others they resembled small styloid crystals that were fused to form one large crystal (Figs. 7, 8). Both types of subunits appear to integrate into one another (Fig. 7). One shattered example had a hollow center (Fig. 9). Crystals were observed to traverse the pit aperture (Figs. 3, 10) and contact the pit membrane (Figs. 11, 12). These did not appear to penetrate the pit membrane, but we are uncertain of this point due to the poor preservation of the pit membranes in our samples. Fig. 12 demonstrates a unique occurrence in which a pit membrane is approached by a crystal from either side. Due to the position of the crystal relative to the pit membrane, we were unable to confirm the presence of a torus in each crystal-associated pit membrane; how- ever a torus was present in the samples that exhibited a crystal behind the pit membrane (fig. 11). No noticeable chamber or crystal sheath was ever observed in association with a crystal. No evidence of a surrounding membrane was discovered, al- though the ends of the crystals often were rough (Fig. 3). Efforts to examine crystals with TEM to determine the presence or absence of a chamber or crystal sheath were not successful. Crystals were not observed in either Botrychium dissectum or B. virginianum. DISCUSSION Three major systems of mineralization occur in plants. These include silic- ification, calcium carbonate crystallization, and calcium oxalate crystallization (Grimson et al., 1982). Calcium oxalate crystals, either in the monohydrate or polyhydrate state, are the most common mineral deposits (Webb and Arnott, 1982). EDS evidence indicates that our crystals contain calcium. The bipyra- midal shape of the crystals’ ends, and the rectangular columns, suggest that they are crystals of calcium oxalate in the polyhydrate form (Frey-Wyssling, 1981). Usually, acid solubility tests are used to confirm crystal composition in plants (Webb and Arnott, 1982). In addition, the oxalate nature of a calcium crystal can be tested with cupric acetate and ferric sulphate (Deshpande and Vishwakarma, 1992). However, due to the small size and sparse number of crystals found in Botrychium multifidum, these tests were not performed. The location of crystals in tracheid lumens in unusual. Crystals in wood are most frequently found in ray or axial parenchyma cells (Chattaway, 1955, 1956), although they may also be found in septate fibers, vessel tyloses, and even in vascular cambia (Deshpande and Vishwakarma, 1992). In Polyalthia, vessels contained a crystalline mass (Scurfield and Mitchell, 1973). In the cur- rent study, crystals were isolated within the tracheary lumen, and there was no evidence suggesting attachment to cell walls. Some crystals appeared to be touching, but were not attached to, the pit membrane. Crystals also were found that had no apparent association with a pit membrane. Therefore, it appeared that crystal formation was not directly related to pit membranes. Crystals in plants often are formed in membrane-bound compartments with- in the vacuole (Arnott and Pautard, 1970; Franceschi, 1984: Webb et al., 1995). As proposed by Arnott and Pautard (1970), the cell membrane may control MORROW AND DUTE: CRYSTALS IN BOTRYCHIUM MULTIFIDUM cv Fics. 9-12. SEM micrographs of crystals in association with pit area and pit membranes. 9 Fractured crystal with hollow center (arrow). Note the aperture behind the crystal (double arrow); scale bar = 5 jm. 10) Crystal entering a pit aperture. Note shaping of the crystal around helical thickening (arrow); scale bar = 2 1) Crystal behind pit membrane; scale bar = 2 pm. 12) Pit membrane associated with crystals from contiguous tracheids; A = aperture; C = crystal; P = pit membrane; scale bar = 2 pr 16 A. Experimental AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) PEAK LISTING RGY AREA cp wh EL. AND LINE - KA B. Control I PEAK LISTING RGY AREA OU bh oN EL. AND LINE a I KA 7 Pic. 13. EDS of tracheid. A. Spectral tracing of crystal within a tracheid. The calcium compone nt of the spe ectrum is conspicuous and is indicated by the peak labeled CA. B. EDS of tracheid without represents calcium in the middle lamella. AU = palladium; SI = MORROW AND DUTE: CRYSTALS IN BOTRYCHIUM MULTIFIDUM 17 both shape and growth of crystals. There was no direct evidence that crystals of B. multifidum were once enclosed in a membrane, but Scurfield and Mitch- ell (1973) suggest that a rough area on a crystal is indicative of the adhering remnants of membrane. In crystals of B. multifidum only the membrane’s im- pression on a crystal would be evident because living portions of the tracheid has undergone autolysis and no membrane remains. If the vacuole with its membrane-covered crystal pressed against either the cell wall or a cell wall thickening as a crystal formed, this contact could explain the shape of these crystals. Water flow though the xylem could also deposit crystals (no longer enclosed by cytoplasm) randomly throughout a tracheid, including on top of a pit mem- brane or between wall thickenings. Due to erosion, water flow might also change crystal shape. Another aspect of crystal development in plant cells in isolation of a crystal by wall material or a suberized sheath after the crystal has formed within a vacuole. This process would, in essence, externalize the crystal (Frank and Jensen, 1970). In Agave (Agavaceae), crystals are produced in such extraplasm- ic compartments (Wattendorff, 1976a, b). Wattendorff (1976b) found that all styloid idioblast of Agave, where they did not touch the wall, were surrounded by a suberized sheath. Although crystals of B. multifidum are styloids, they appear neither to be associated with a sheath of any sort nor to be isolated by cell wall material. The reason a cell forms a crystal is not well understood. Crystal formation may represent a crystallization of waste material or storage of minerals (Desh- pande and Vishwakarma, 1992). Crystal formation also may be associated with ionic balance, and therefore, the formation of a crystal could be a form of osmoregulation (Franceschi and Horner, 1980). Franceschi and Horner (1979) correlated the amount of calcium in the growth medium and the number of crystals formed in Psychotria L. (Rubiaceae) callus. Lane (1994) has suggested that calcium oxalate crystals may promote the polymerization of lignin which of course, would be occurring in the developing tracheids of B. multifidum. It is evident at times that crystal formation in plants is under genetic control (Frey-Wyssling, 1981; Webb, 1999); however, genetic control of the formation of all crystals has not been proven. The cell in which a crystal is produced undergoes many changes at macro, micro, and ultrastructural levels, as well as, changes in cell chemistry. These changes, documented in other plant taxa during crystal formation, make it unlikely that crystal formation could be sim- ply the result of precipitation or crystallization (Franceschi and Horner, 1980), although crystal formation may represent crystallization of waste material or storage of minerals (Deshpande and Vishwakarma, 1992). Deshpande and Vish- wakarma (1992) also identified seasonal fluctuation in crystal formation after the cessation of cambial activity. Gourley and Grime (1994) described crystals that were more commonly found in the late wood of Acacia Mill.(Fabaceae). The availability of water was also determined to be a factor in crystal formation (Gourley and Grime, 1994). It was impossible to determine for certain whether crystals in the tracheids 18 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) of B. multifidum formed before or after cell death. Perhaps due to greater water flow resistance occurring at the pit membrane, there would have been a greater chance for calcium precipitation in the pit area rather than in the tracheary lumen. If this were the case, crystals could at the pit membrane form after the death of a tracheid. However, the rough ends observed on some crystals sug- gest they may have been enclosed at one time by a membrane. Additionally, the crystals appear to conform to the shape of the pit aperture or cell wall thickenings and do not appear to have been randomly distributed by water flow. Perhaps the best explanation of where these crystals develop is in mem- brane compartment within vacuoles of living tracheids. The enlargement of a crystal in a plane perpendicular to the cell’s axis would result in its abutting a wall or pit membrane, thus influencing crystal shape. Crystals that elongated parallel to a cell’s axis would not encounter these boundaries and would not be shaped by them. Crystals that were not pressed into a cell wall or pit mem- brane also would not develop this shaping and might settle between wall thickenings, or after cell death, move with the xylem water stream. However, the positions of crystals in Figures 3, 10, and 12 with respect to the pit mem- brane suggest that crystal position is not the result of water flow. Crystal formation has also been associated with the products of fungal me- tabolism within the plant cells (Scurfield and Mitchell, 1973). In our study, no fungal hyphae were found near any of the crystals. Therefore, this possibility in B. multifidum seems unlikely. If crystal manufacture is under genetic control, what advantage does the cell gain from its production? This is an especially intriguing question with regards to Botrychium as crystal production would be occurring in cells about to die. Lane (1994) has suggested that calcium oxalate crystals play a role in lignin polymerization and perhaps this may be true in these lignified tracheids. How- ever, the lack of crystals in other Botrychium species indicates that this would be true only under certain environmental conditions. As previously men- tioned, some authors believe that crystals represent the storage of calcium that could be either reserve calcium or waste calcium (Deshpande and Vishwak- arma, 1992; Webb, 1999). Storage of needed calcium in a near-death cell would be unlikely. However, these crystals may have been produced by a cell for ionic balance or osmoregulation. Ionic balance and osmoregulation are critical for immature cells (Franceschi and Horner, 1980). If a plant were growing on soil with high nitrate levels, assimilation of this compound would increase cell pH, and oxalic acid might be produced to counter this effect. The oxalate anion could then react with calcium to form a crystal that would remove the excess anion from cell sap (Franceschi and Horner, 1980). Another explanation could be protection from herbivores, although crystal production in a leaf cell would be more plausible for defense purposes. The crystals in B. multifidum are too small and too few in number for this type of protection. Fire protection was listed as an explanation by Gourley and Grime (1994) for crystals in Acacia, but again this is unlikely for a rhizome. Based on our data the best explanation for crystal formation in the xylem of B. multifidum is that crystals are the result of excess calcium precipitation, MORROW AND DUTE: CRYSTALS IN BOTRYCHIUM MULTIFIDUM 19 which could represent either waste, storage, or osmoregulation in the plant. Because the crystals are located in dead cells, active resolubilization by a cell would be unlikely; however, if crystals were dissolved by water flow in the xylem, their minerals could be carried in the transpiration stream. Deshpande and Vishwakarma (1992) have suggested that formation of calcium crystals may be a reversible process in some tissues. Therefore, these crystals do not necessarily represent a calcium loss for the plant. LITERATURE CITED ARNOTT, H. J., and F. G. E. PAUTARD. 1970. Calcification in plants. Pp 375-446, in maida saledicatian: Cellular and molecular aspects. H. Schraer, ed. North-Holland, Amsterdam CHaTTaway, M. M. 1955. Crystals in woody tissues. Part I. Trop. Woods 102:55—74, ———. 1956. Crystals in woody tissues. Part II. Trop. Woods 104:100—24. DESHPANDE, B. P., and A. K. VISHWAKARMA. 1992. Calcium beam a in the fusiform cells of the pli es of Gmelina arborea. [AWA Bull. N.S.. isa V. R. 1984. Developmental — : calcium ee an sand depositions in vulgaris L. leaves. Protoplasma 12 H. T. HORNER. 1979. Use of aera eis callus in study of calcium oxalate crystal idioblast formation. Z. Pflanzenphysiol. 92:1-75. 1980. Calcium oxalate crystals in plants. Bot. Rev. (Lancaster) 46:361—42 FRANK, E. and W. A. ae 1970. On the formation of the pattern of crystal idioblast in ade ensiformis D.C. IV. The fine structure of the crystal cells. Planta 95:202—217. FREY-WYSSLING, A. 1981. Crystallography of the two hydrates of crystalline calcium oxalate in plants. Amer. J. Bot. 68:130-141 GIFFORD, E. M., and A. S. FOSTER. 1989. oe and evolution of vascular plants. 3rd ed., eeu and Co., Salt Lake C Gour.ay, L. D., and G. W. GRIME. an. Calc cium oxalate . in african Acacia isa and their says by scanning proton a (SPM IAWA Bull. N.S. 15:137-1 ee M. J., H. J. ARNOTT, and M. A. WEBB. 1982. A scanning electron eo ins of 71133-1140. winged point — in the bean legume. Scanning Electron Microscopy III Hits, W. E. . Formation of robinetin crystals in vessels of Intsia species. ore ng vie N.S. x reaione Konpo, Y., T. Fuyl, ¥. HAYASHI, and A. Kato. 1996. Organic crystals in the tracheids of Torreya yunnanenss IAWA Bull. N.S. 17:393—403. Lang, B. G. 1994. Oxalate, germin, and the extracellular matrix of higher plants. F.A.S.E.B.J. 8: 294-301. NATION, J. ca 1983. A new method using sre ea ca for preparation of soft insect tissues for scanning electron microscopy. Stain Technology ScurFIELD, G., and A. J. MITCHELL. 1973. Crystals in woody ely Bot. J. Linn. Soc. 66:277—289. WarTTENDoREF, J. 1976a. A third type of raphide crystal in the plant kingdom: six-sided raphides i 4a ioe animate of the suberized svioid crystal cells in pam leaves. Planta (Berl.) ee pier Wess, M. A. see Cell-mediated crystalliztion of calcium oxalate in plants. Pl.Cell 11:751—761. , and H. J. ARNOTT. 1982. A survey of calcium oxalate crystals and other mineral inclusions in seeds. eres Electron Microscopy III:1109-1131 ALETTO, N. C. Carpira, L. E. Lopez, and H. J. ARNOTT. 1995. The intravaculolar oat fae associated with calcium oxalate crystals in leaves of Vitis. Plant J. 7:633-648 American Fern Journal 92(1):20—22 (2002) A New Filmy Fern from the Dominican Republic CARLOS SANCHEZ Jardin Botaénico Nacional, Carretera del Rocio, km 3.5, Calabazar, Boyeros, C. P. 19230 Ciudad de la Habana, Cuba ABSTRACT.—A new species of Hymenophyllum subg. Hymenophyllum with entire involucral valves is described from the Dominican Republic on the island of Hispaniola. During the prepartion of a revision of the filmy ferns (Hymenophyllaceae) for the Flora of the Greater Antilles project, a peculiar species was discovered among the undetermined specimens in the Gray Herbarium. The taxon belongs to Hymenophyllum subg. Hymenophyllum, following Morton (1968, pp. 162-164), a subgenus represented by only two species in the Antilles. Hymenophyllum tunbrigense (L.) Smith is known only from Ja- maica and Hispaniola (Proctor 1985, p. 90), whereas H. fucoides (Sw.) Sw., is more widely distributed (Proctor, 1985, p. 92; 1989, p. 58). The members of this subgenus are characterized by having toothed segment margins, and most have sinuous to toothed involucral valves as well. The new species is de- scribed as Hymenophyllum integrivalvatum C. Sanchez sp. nov. Fig. 1 Ab speciebus aliis antillanis subgeneris Hymenophylli valvis integris, stip- itibus brevissimis, segmentis pinnarum paucis (1 vel 2), necnon laminis gla- berrimis diversa. TypE—Dominican Republic: Pcia. La Vega: Near the pyramid ca. 13 km from Valle Nuevo on the road to San José de Ocoa, ca. 2500 m elev., 22 August 1957, Gastony, Jones & Norris 740 (GH; isotype US). Rhizomes creeping, filiform, 0.1-0.3 mm in diam., clothed with deciduous, brownish, pluricellular trichomes, with a few conspicuous, straight roots ca. 5 mm distant. Fronds small, erect, determinate, approximate, 1.15—2.1 cm long.; stipes 0.1-0.3 mm long., 0.2 mm in diam., very narrowly alate thorough- out, dark brown, glabrous or with a few brownish, often 2-celled trichomes; laminae narrowly ovate, lanceolate or oblong 1.4-1.8 cm long, X 0.8-1 cm wide, pinnate-pinnatifid; rachises notably flexuous, narrowly and evenly alate, the alae less than 0.1 mm wide, dark brown, glabrous; pinnae 5-8 pairs, spreading to ascending, mostly with 2 acroscopic segments; segments narrowly elliptic, oblong, or linear-oblong, 1.2-1.8 mm wide, glabrous, the margins dis- tantly toothed, the teeth usually more distant than their length and ascending, the midvein dark brown, the lamina tissue olivaceous-green when dry; sori conspicuous in size in comparison with the length of the lamina, borne at the lamina apex or in the distal half, subaxillary on the acroscopic side of the SANCHEZ: A NEW FILMY FERN FROM THE DOMINICAN REPUBLIC 21 B22, 4. Cale ry Se Fic. 1. Holotype of Hymenophyllum integrivalvatum (GH). A. Habit of plant. B. Detail of an ultimate segment. C. Detail of a sorus. pinnae; involucres 1.6—2.2 mm long, 1.4 mm wide, broadly elliptic or broadly ovate, bivalvate, the valves wider than the sterile segments, the margin entire, the filiform receptacle included. DISTRIBUTION.—Endemic to Hispaniola (Dominican Republic), known only from the type collection. Hasirat.—Epipetric in very moist burned and timbered pinelands, forming thick mats on rocks along streams in very moist ravines, according to the in- formation on the label. The new species is most closely related to H. tunbrigense (L.) J. E. Smith, which is also known from a few collections from Hispaniola. Hymenophyllum tunbrigense differs in having larger fronds, straight rachises with wider alae, more divided pinnae, narrower segments, and sinuous involucres. The entire involucral valves, very small fronds, and the absence of trichomes on the re- mainder of the lamina separate the new species from all other Antillean spe- cies of subg. Hymenophyllum. ACKNOWLEDGMENTS I am indebted to the following individuals for their help and hospitality during my visit to different North American herbaria to study Greater Antillean pteridophytes: John T. Mickel and Thomas Zanoni (New York Botanical Garden), David B. Lellinger (United States National Herbar- ium), and Emily Wood and David E. Boufford (Gray Herbarium, Harvard University). I am grateful to the New York Botanical Garden and The John D. & Catherine T. MacArthur Foundation for financial support. Thanks also to David B. Lellinger for revising the English version of the man- uscript. I also wish to thank Manuel G. Caluff for the illustrations. 22 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) LITERATURE CITED Morton, C. V. 1968. The Genera, Subgenera, and Sections of the Hymenophyllaceae. Contr. U. S. Natl. Herb. 38:153-214. Proctor, G. R. 1985. Ferns of Jamaica: A guide to the pteridophytes. British Museum (Natural History), London, . 1989. Ferns of Puerto Rico and the Virgin Islands. Mem. New. York Bot. Gard. 53: 1-389. American Fern Journal 92(1):23—29 (2002) Adiantum argutum, an Unrecognized Species of the A. latifolium Group JEFFERSON PRADO Secdo de Briologia e Pteridologia, Instituto de Botanica, Caixa Postal 4005, 01061-970 Sao Paulo, SP, Brasil Davip B. LELLINGER Department of Botany, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0166 ABSTRACT.—The present paper distinguishes A. argutum, an unrecognized but widespread species from South America, from the related A. Jatifolium, and designates a lectotype for A. argutum. Several pinnate or bipinnate Adiantum species have an indument like that of the A. serratodentatum group, but differ in having fewer, larger, less di- midiate pinnules and thin, very long-creeping rhizomes. Among the species of this group are A. argutum Splitg., A. incertum Lindm., which is based on Lindman Regnell Exped. I. A2083 (S not seen; isotypes B, GH) from Paraguay, the widespread A. Jatifolium Lam., and A. viviesii Proctor, which is based on Proctor 41389 (US; isotypes IJ, SJ) from Puerto Rico. Adiantum glaziovii Baker, which is based on Glaziou 13345 (K, isotype US) from Rio de Janeiro, Brazil, is a synonym of A. Jatifolium. The early, unplaced name A. elatum Desv., which is based on a Brazilian specimen from the Herb. Dombey (P-Herb. Juss. Cat. 1421 not seen Morton photo 3153) will likely displace one of the later named species found in Brazil. The specimen was said by Morton to have almost glabrous segments; it needs to be examined critically. Adiantum argutum Splitgerber, Tijdschr. voor Natuurl. Gesch. en Physiol. [Leiden] 7: 427. 1840. Figs. 1, 2. Lectotype (chosen here): “in sylvis montosis Surin. prope Bleauwe Berg,” Surinam, May 1838, Splitgerber 891 (L not seen, photo US). Other syntype: idem, id., Splitgerber 290 (L not seen, photo US). Adiantum fovearum Raddi var. reductum Jenm. Ferns Br. W. Ind. & Guiana 87. 1899. TYPE: Not clearly stated, but presumably Guyana, Jenman (NY? not seen). Splitgerber (1840) described A. argutum based on material he collected in Surinam. The author did not mention any specimens in the protologue, only the locality. Original materials were found by Morton at Leiden (Morton photos 193 and 194, both US), and they may be considered the syntypes of Splitger- ber’s taxon. The lectotype selected here (Splitberger 891) has an original label, handwritten by the author with the same information he published with the 24 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) Ss ° J gv - is ° \ 10 o 2 ¢ 3 o “Sa ‘4 e e o —— a : ° e t eveinaph Ottis Pl ° en Me ind <— aa yy. ‘, ane = Be = é bad e \ a \ ° r bane 0 , ° a ° f eS @ Fic. 1. Range of Adiantum argutum Splitg. original description. The photograph of the other syntype (Splitgerber 290) resembles A. Jatifolium Lam., although we can not place it there with certainty. According to Splitgerber (1840), A. argutum has a long-creeping rhizome; the laminae are lustrous adaxially and have 3 or 4 pinna pairs, acuminate pinnules, a subrhombic terminal pinnule, reduced and flabellate basal pin- nules, sparse, minute setiform scales abaxially, and oblong sori. In fact, these characters distinguish this species from all others closely related to it. Other important features to recognize A. argutum are the distant fronds and the id- ioblasts on the abaxial surface of the pinnules. Unfortunately, over the years, the Splitgerber species was included within the concept of A. Jatifolium by Vareschi (1969, p. 734), Kramer (1978, p. 91), Tryon and Stolze (1989, p. 66), and Smith (1995, p. 259). In other cases, the PRADO AND LELLINGER: ADIANTUM ARGUTUM 25 name was synonymized under this species, by Posthumus (1928, p. 105), Lel- linger (1989, p. 148), and Cremers and Hoff (1990, p. 19). Adiantum argutum demonstrates its close affinity to A. Jatifolium mainly in its slender, long-creeping rhizome, 2-pinnate fronds, and stipe and rachis cov- ered by deltate to lanceolate scales with a pectinate base. However, A. latifol- ium differs by its smaller, obtuse to subacute pinnules with a roundish apex that are abaxially glaucous, glabrous, and without idioblasts. Adiantum incertum Lindm. differs from A. argutum and A. Jatifolium in having the scales on the abaxial surface of the pinnules hairlike and with a few basal processes, rather than having such scales with a pectinate base or totally lacking scales. In addition, it is a species restricted to Paraguay and extra-Amazonian Brazil: Goids (Maurilandia, Rio dos Bois, Hatschbach 34271, MBM, MO, NY, UC), Mato Grosso (Santa Terezinha, 21 km SW of Portal da Amazonia, Thomas et al. 4334, NY, US), SAo Paulo (Sao Carlos, 9 km NNE of the BR Station at Santa Eud6xia, Eiten & Eiten 3488, US), and Parana (Foz do Iguacu, Parque Nacional das Cataratas, Hatschbach 23171, HB, MBM, MO, UC, UPCB). Adiantum obliquum Willd., although its fronds look much like those of A. argutum, can be distinguished by its short-creeping rhizomes, approximate and usually 1-pinnate fronds, and pinnules with conspicuous idioblasts on both surfaces. It may be more related to A. Jucidum Cav. and perhaps to A. petiolatum Desv., with which it hybridizes. These three species may form a separate group. Adiantum argutum has a more restricted area of distribution than its closest relative A. Jatifolium. It occurs in northern South America (Colombia to French Guiana) and in the Amazonian regions of Peru, Bolivia, and Brazil. It grows in primary and secondary forests, on dark red lateritic clay soils, from 50 to 1000 m elevation. Representative specimens of A. argutum studied: COLOMBIA: Meta: Sierra de La Macarena, Cajfio Entrada, Philipson & Idrobo 1748 (US); Villavicencio, Pennell 1607 (GH). Boyaca: Los Llanos, Haught 2833 and 2844 (both GH). Vichada: San José de Ocumé: near Rio Vichada at Botomi, ca. 14 km NW, Hermann 11107 (US); NE de Pto. Infrida, 3°58’N, 67°50’W, Churchill et al. 17748 (NY). VENEZUELA: Bolivar: La Tomasa, Williams 1295 (US); Rio Paragua, Isla El Casabe, Killip 37301 (US); Salto Alta, Alto Orinoco, Croizat 486 (NY); Dtto. Sifontes, Concesién Minera Oro Uno, 7 km NW of la Clarita, 6°13'N, 61°27'W, Aymard et al. 3976 (NY); Sierra Imataca betw Rio La Reforma and Puerto Rico, N of El Palmar, Steyermark 88012 (US). Amazonas: Around the margin of the Rio Orinoco above Tamatama, Williams 15199 (GH); Cuenca del Rio Manapi- are, 5°5’N, 66°03’W, Huber 435 (NY). Delta Amacuro: Rio Cuyubini, Cerro de la Paloma, Steyermark 87649 (NY). Mérida: Near border Rio Grande de Toro, 61°44’W, 80°4'N, Breteler 3781 (US). TRINIDAD: Fendler 2 (NY). GUYANA: Cuyuni-Mazaruni: 8 km N of Bartica on W bank of Essequibo River, 06°29’N, 58°38’W, Henkel & Chin 297 (US). U. Takutu-U. Essequibo: i io” ” a INS PRADO AND LELLINGER: ADIANTUM ARGUTUM 27 Rupununi area, Surama Village, 04°08’N, 59°04’W, Acevedo et al. P3297 (US); Marudi River, 02°11’N, 59°11'W, Henkel et al. 2902 (NY), 3032 (US); Kuyuwini River, 02°11'N, 59°11’W, Henkel et al. 3022 (US); NW Kanuku Mts., 3°21'N, 59°30'W, Hoffman & Foster 3510 (US); Rupununi River, Jansen-Jacobs et al. 4207 (US). Potaro-Siparuni: On 0.5 km island in Essequibo River, 1 km S of Fairview, 4°40'N, 58°40'W, McDowell 3371 (US); Iwokrama Mts., Annai-Ka- rupukari Rd., 04°19’N, 58°51’W, Hoffman et al. 1409 (US); River Isherton, 2°20'N, Smith 2432 (GH, NY). Barima-Waini: Head of Barima River, Ayamba Falls, 4.5 mi W of Eclipse Falls, ca. 10 km W of Arakaka, 7°39’N, 60°09’, Pipoly & Lall 8200 (NY); Head of Barima River, NW of Kariako River, 7°30'N, 60°35’W, McDowell 4393 (NY); Labbakaka Creek, Tiger Creek, Sandwith 1209 (K, NY). SURINAM: Haut Litany: Basin du Litany, 2°31'N, 54°45’W, Granville et al. 12040 (US). Nickerie: Area of Kabalebo Dam project, Lindeman & Roon 884 (US); Area of Kabalebo Dam project, 4°—5°N, 57°30’—-58°W, Lindeman et al. 165 and 343 (both NY); Sectie O, along railroad, vic. Km. 70, Maguire & Stahel 23605 (GH, NY). Brokopondo: 2.4 km S of village Gansee, Donselaar 1189 and 1276 (both GH); Zuid River, 3°10’—3°20’N, 56°29’-56°49’W, Kayser Airstrip, 45 km above the confluence with Lucie River, Irwin et al. 57697 (NY). FRENCH GUIANA: Cayenne, Inini River, 3°28’N, 52°36'30’"W, Cremers et al. 8781 (US); Camp Eugene, Basin du Sinnamary, 4°51’S, 53°4'W, Cremers & Granville 13727 (NY); Gobaya Soula, Basin du Maroni, 53°58’W, 3°37’S, Cre- mers et al. 10125 (US); Saoul, 3°37’N, 53°12’W and vicinity, Route de Bélizon, N of Eaux Claires, Heald & Yahr 56 and 65 (both NY); Comté., degrad auprés de Crique Martineau, Oldeman 1426 (NY); Mt. Balbao, Secteur Sud, 3°35’N, 53°20'W, Granville et al. 8958 (NY). PERU: Madre de Dios: Near the confluence of Rio Tambopata and Rio La Torre, 39 km SW of Puerto Maldonado, 12°50’S, 69°20’W, Smith & Condor 1114 (US) and 1363 (NY); Tambopata, Vargas 18577 (GH); Tambopata, vic of Moho towards Piedra Redonda, at the Bolivian frontier, 12°30'S, 69°40’W, Nu- fiez et al. 9695 (GH, NY); Tambopata, SSW of Pto. Maldonado at the confluence of the R. La Torre and the R. Tambopata (SE bank), Tambopata Nature Reserve, 12°49’S, 69°17'W, Barbour 4763 (NY), Loépez 4585 (GH). BOLIVIA: Beni: Pcia. Ballivian: Rio Colorado, Collegio Técnico Agropecu- ario de Rfo Colorado, 15°00’S, 67°10'W, Fay & Fay 2105, 2640, 2652, 2654, 2681 (all US); 18.4 Km E of Riberalta, then 1 km NE on old road to Cachuela Esperanza, 11°05’S, 65°50’W, Solomon 7804 (NY); Isla Capanario, 50 m from the San Borja—San Ignacio de Moxos road, 212 km from Campoamento Totai- zal, Rolleri 140 (NY); Pcia. Moxos: Chimanes Forest, 15°10’S, 66°37’W, Fay & Fay 2794 (US). Sta. Cruz: Bella Vista, Rio Blanco, Scolnik & Luti 681 (US); Pcia. Ichilo: Old meander loop of the Rio Ichilo, 1-1.5 km SW of the Buena Vista—-Villa Tunari Hwy., 17°18'S, 64°12’W, Nee & Moran 45225 (NY). Pando: Nicolas Suarez, SW of Cobija on the Rio Naraueda, 11°08’S, 69°08’W, Sperling oo Fic. 2. Adiantum argutum Splitg. Fig. 2A. Habit. Fig. 2B. Abaxial surface of some pinnules. Fig. 2C. Abaxial surface of a pinnule showing the scales. 28 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) & King 6475 (GH, NY, UEC); Ca. 20 km from Cobija towards Castro Erifia, Casas & Sussana 8123 (NY); W bank of the R. Madeira betw Cachoeiras Madeira and Misericordia, Prance et al. 6612 (NY). La Paz: Pcia. Iturralde, Siete Cielos, R. Manupare, 12°27'S, 67°37’W, Solomon 16947 (NY). BRAZIL: Amapa: Serra do Navio, bank of the Rio Amapari, Emmerich & Andrade 745 (HB, R). Roraima: Posto Mucajai, Rio Mucajaf, vic of Mucajai airstrip, Prance et al. 10991 (GH, R, UC). Para: Lageira, airstrip on Rio Mai- curt, 0°55'S, 54°26’W, Strudwick et al. 3088 (NY), 3129 (MG, NY, US), 3580 (US); Curud S.A., near Alenquer, Santarém, Piggott 2547 (K, NY); Breu Branco, ca. 40 km S of Tucurui, 4°03’S, 49°40’W, Daly et al. 1376 (GH, MO, NY, US); Serra dos Carajas, Serra Norte, ca. 15 km W of AMZA Exploration Camp., 6°S, 50°15’W, Berg & Henderson BG472 (GH, NY, UC, UEC, US); Serra dos Carajas, 6°4'S, 50°8’W, Secco 286 (GH, K, MG, NY, SPF); BR-163, Cuiabé—Santarém Highway, km 885.5, Prance et al. P25171 (MG, NY, UC, US); Parque Indigena do Tumucumaque, Rio Parti de Oeste, Miss&o Tiriyo, 2°20’N, 55°45’W, Caval- cante 2401 (K, MG, NY, US); Conceigaéo do Araguaia range of low hills ca. 20 km W of Redengao, near Sado Jodo and Troncamento Santa Teresa, 8°03’S, 50°10’W, Plowman et al. 8635 (GH, NY), 8757 (GH, NY, US); Rio Xingu, Balée 2398 (NY); Confluéncia com Rio Pardo, Vasconcelos et al. 260a (NY); Rio Ita- caiunas, affluent of the Rio Tocantins, Serra Buritirama, 50°15’W, 5°30’S, Pires 12427 (NY); Rio Cumina, Ducke (Hb. Mus. Goeldi 8885, 15163) (both HB, MG). Amazonas: 1—5 km road Boca do Acre to Rio Branco, Prance et al. 2533 (GH, MG, NY, R, US); Vic. of Tototobi, Basin of the Rio Demeni, Prance et al. 10208 (NY, UC, US); Rio Curuquete, Providencia, Prance et al. 14632 (NY, UC); Sao Paulo de Olivenga, 30 km above the mouth of the Rio Coti, Prance et al. 14444 (B, NY); Vic. of Macujai airstrip, Prance et al. 10991 (MG, NY, UC); Borba, 4°02'S, 59°06’W, W side of the Rio Cunama, Hill et al. 12868 (MO, NY). Ron- dénia: Mineragao Campo Novo, BR-421, a 2 km a Oeste da Mineragéo Campo Novo, 10°35'5"S, 63°37’W, Vieira et al. 517 (NY, US); Basin of Rio Madeira, Trail north of Rio Madeira from 2 km, below confluence of Rio Abuna, Prance et al. 8345 (K, NY, UC, US). Acre: Palacio de Castro, fazenda Mococa, ramal no km 120 da rod. Rio Branco—Pérto Velho, Santos et al. 122 (MG, NY, US); 9 km from Rio Branco on Rio Branco—Pérto Acre road at cut-off for Colénia Cinco Mil, Lowrie et al. 650 (MG, NY, R, US); Sena Madureira, Bacia do Rio Purus, varagao para o Seringal Fonte Boa, 10°07’'S, 69°13’W, Silveira et al. 668 (MO, NY); Xapuri, Seringal Cachoeira, 35 km SE of Xapuri, Pinard 809 (NY), Kainer 126 (NY). Mato Grosso: Colider, Comunidade Sao Francisco, Salino 284 (GH); Serra Ricardo Franco, 15°S, 60°W, Windisch 1503 (HRCB); Rio Pei- xoto de Azevedo, Faz. Sao José (Cachimbo), Bokermann 6747 (UEC); Santa Terezinha, BR-158, Vila Confresa, pr. ao aeroporto da Fazenda Confesa, 10°35'S, 5°35’W, Windisch 5987 (UC). ACKNOWLEDGMENTS The first author appreciates the financial support of the Brazilian Research Council CNPq (Proc. n. 300843/93-3 and 450658/99-6) and of the Smithsonian Institution, Washington DC (Short-Term Visitor Grant). We thank also Sra. Emiko Naruto for preparing the illustration. PRADO AND LELLINGER: ADIANTUM ARGUTUM 29 LITERATURE CITED Cremers, G. and M. Horr. 1990. Inventaire taxonomique des plantes de La Guyane Francaise. I- Les ace ta Museum National d’Histoire Naturelle, Inventaires de Faune et Flore. Fasc MER, K. U. 1978. ors pteridophytes of Suriname. Uitgaven Natuurw. Stud. Suriname Nederl. Antillen 93:1-19 LELLINGER, D. B. 1989 ais ferns and fern-allies of ssi a Panama and Chocé (Part 1: Psilo- taceae through Dicksoniaceae) haope 2A PosTHUMUS, O. 1928. The ferns of Surinam. pre ie Java. 1 SMITH, A. R. 1995. poe L. Pp. eae in P. E. Berry, B. K. Holst, ae Vatachievsdhi (eds.), Flora of the Venezuelan Guayana, vol. 2: Pteridophytes, Srempatophytan: Acanthaceae—Ar- aceae. Timber Press, Portland. SPLITGERBER, F, L. 1840, Enumeratio Filicum et Lycopodiacearum quas in Surinamo legit F. L. iteadies Tijdschr. Natuurl. Gesch. Physiol. 7:391-444. TrYON, R. M. and R. G. STOLZE. ee ah eae of Peru, Part II. 13.Pteridaceae—15.Dennstaed- tiaceae. Fieldiana Bot., n.s. 2 VARESCHI, V. 1969. Helechos. = oat sod in T. Lasser (ed.), Flora de Venezuela. vol. 1, Tomo 2. Instituto Boténico, Caracas American Fern Journal 92(1):30—38 (2002) Polypodium vulgare Plants Sporulate Continuously in a Non-Seasonal Glasshouse Environment SANNA E. SIMAN AND ELIZABETH SHEFFIELD School of Biological Sciences, 3.614 Stopford Building, The ee of Manchester, Oxford Road, Manchester M13 9PT, U ABSTRACT.—In their natural environments pteridophytes usually have regular sporing periods, the onset of which is triggered by the interaction of climatic and nutritional factors. Little, however, is known about what changes there may be in the sporing behaviour of a fern when it is transferred from its natural habitat to an artificial environment, such as a glasshouse. We recorded sporing behaviour in relation to vegetative growth in two genetically matched populations of Polypodium vulgare. One population was placed in a controlled-climate glasshouse, the other was left outside. The recruitment of new fronds was significantly higher in the indoor population than in the outdoor population. The indoor population also maintained a high proportion of actively sporing fronds throughout the winter. There was no net recruitment of new fronds in the outdoor ce lation during the winter and early spring. Some elements of the glasshouse envi t, probably the enhanced light and temperature, induced continuous sporing in this fern. Considering the ever-increasing interest in ferns as ornamental plants, and the growing body of evidence of toxic and allergenic effects caused by fern spores, this kind of sporing behaviour may have implications for human health. Ferns in their natural environments usually have regular and predictable periods of spore production and release. In the temperate zones and the sea- sonal tropics they tend to release their spores towards the end of the growing season. In the wet tropics, where the growing season is much longer or even continuous, initiation of new fronds and maturation of older fronds take place throughout the year (Page, 1979). Most ferns have been said to show very little fluctuation in annual spore output with variations in climate, in contrast with good and bad seed years in angiosperms and conifers (Page, 1979). This does not apply to all fern species. Page (1976) pointed out that for Pteridium aquil- inum (bracken) the spore yield can vary widely between different years. Sim- ilarly, Steeves (1959) noted that, for Osmunda cinnamomea, a hot dry summer is usually followed by a high degree of fertility in the following spring, where- as a cooler moister summer leads to reduced fertility. Furthermore, the onset of the reproductive phase in fern sporophytes can be demonstrated to be reg- ulated by the interaction of several factors, such as light exposure, temperature and the nutritional status of the plant. Field observations have suggested that the onset of the reproductive phase in ferns (as in many flowering plants) is induced by particular photoperiods, but the evidence so far published is scanty (Wardlaw and Sharma, 1963). periments carried out by Wardlaw and Sharma (1963) indicated that there is a more or less direct relationship between active photosynthesis and/or pho- toperiodic perception in the expanded leaves and the induction and devel- opment of sori in the next inner leaves of Dryopteris austriaca. Harvey and SIMAN & SHEFFIELD: POLYPODIUM VULGARE SPORULATION 31 Caponetti (1972), however, demonstrated that increasing light intensities in- hibited sporophyll differentiation in Osmunda cinnamomea. Maximal initia- tion of sporangia occurred in total darkness in this species, so it may be that green and non-green spored ferns respond to light in different ways. There may be doubts about the importance of photoperiodic induction, but in deter- mining the extent of the fertility in ferns, photosynthetically available radia- tion and the duration of exposure to light are probably of major importance. Steeves (1959) compared the incidence of fertility between O. cinnamomea plants in heavy wood and open areas and found a greater incidence of fertility in the latter plants. Conway (1957), Dring (1965) and Page (1976) suggested the same property in Pteridium (bracken), in which they found a gradual de- crease in fertility with increasing degree of shade, although vegetative growth in the latter may be little impaired. The enhancement of sporogenesis by high light has recently been confirmed in experiments conducted with clones of bracken grown in high and low levels of photosynthetically available radiation (Wynn et al., 2000). Temperature also plays a role in the onset of the reproductive phase as shown by Labouriau (1958). He found that initiation of sporangia was stimu- lated by exposure of the developing outermost set of Osmunda claytoniana fronds to a temperature of 26°C; plants kept at a lower temperature remained sterile. Similar trends were reported in Pteridium by Sheffield (1996) and Wynn et al. (2000). Allsopp (1964, 1965) suggested that nutritional conditions, particularly car- bohydrate supply, appear to be of greater importance for the induction of spo- rangia in pteridophytes than photoperiodic or similar stimuli. Several studies indicate that the nutritional status of the plant is indeed of great importance for the initiation of the spore-productive stage (Wardlaw and Sharma, 1963). Goebel (1887, 1905, 1908) and Atkinson (1896) concluded, from experiments with Onocleoid ferns, that if carbohydrate supplies are inadequate, developing leaves tend to remain in the vegetative state. Goebel (1928) found that imma- ture sporophylls of certain ferns developed as vegetative fronds when the ster- ile fronds of the plant were removed. This has been repeated in numerous fern species (e.g. by Labouriau (1958), Wardlaw and Sharma (1963), and Steeves and Wetmore (1953)), who thus linked induction of sporogenous tissue to car- bohydrate supply. Sussex and Steeves (1958) cultured excised leaf primordia of Leptopteris hymenophylloides, Todea barbara and Osmunda cinnamomea, and found that high sucrose concentrations in the medium was essential for the inception and early development of sori and sporangia in those species. They also showed that an increased supply of inorganic nitrogen promotes the onset and extent of fertility in T’ barbara. According to Wardlaw and Sharma (1963), there is a positive relationship between the amount of photosynthesis- ing leaf surface and the induction of fertility. It is apparent from the above that we have some limited knowledge of what triggers sporing in ferns in natural conditions. The sporing behaviour of ferns transferred from their natural habitat to an artificial environment, e.g. a glass- house, however, remains largely unexplored. Considering the ever-increasing 32 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) interest in ferns as household and garden ornamentals (Gress, 1996) and the fact that fern spores may cause adverse health effects in humans (Siman et al., 1999), more knowledge in this field is urgently required. The aim of the current experiment was to compare the reproductive perfor- mance of Polypodium vulgare plants placed in either a seasonal outdoor en- vironment or a constant high-temperature and high-light glasshouse environ- ment. MATERIAL AND METHODS Polypodium vulgare plants were collected in March 1998, from stone walls along the south-east side of road B4403, running along the south-east shore of Llyn Tegid (Bala Lake) (N 52°53’, W 3°38’), Wales, U.K. Polypodium was cho- sen as it represents a widespread genus, including a broad range of horticul- tural favorites, such as P. amorphum, P. cambricum and P. interjectum, of sim- ilar morphology and life history (Mickel, 1994). he plants were potted in commercial potting compost within a day of col- lection. Lengths of rhizome were split into two equal parts (i.e. bearing the same number of fronds in each of the two pots in a pair), and each was placed in one pot. In this way 64 pots were prepared, i.e. 32 pairs of pots containing clones. The length of the potted rhizomes varied from 1 to 5 cm, but was much the same within each pair. In order to minimise the impact on the results of the growth of any apical meristems, only median pieces of rhizome were used. All plants were allowed a six-month settling-in period (mid-March to mid- September) outdoors, after which one pot of each pair was left outdoors, to the north-east of a glasshouse in the Manchester University Experimental Grounds, Manchester, U.K. These were the ‘‘outdoor population”. The other group of the plants was put inside a glasshouse (mean day temperature: 28°C, range 20-38; mean night temperature: 15°C, range 11—27; photosynthetically available radiation: c. 110 pmol m~? s~'), at the aforementioned Experimental Grounds. These were the “indoor population”. At the start, the total number of fronds in each population was very similar. No plants were given any ad- ditional nutrients during the course of the experiment. Those outside were subject to ambient rainfall, those inside were watered regularly. During the settling-in period all fronds in 14 pots, seven in each population, died. Eight of the 14 pots belonged to matched pairs, so the aim of ensuring a genetic similarity between the two populations was still met to a high degree. Weekly records were taken of the number of fronds in each pot with a) no sori, b) immature (green) sori, c) sporing (yellow-orange) sori and d) empty (brown) sori, from the beginning of October 1998 until June 1999 for the indoor population. The outdoor population was recorded until the end of its growing season in mid-September 1999. The proportions of recruited fronds during the experimental period by the two populations were compared with a x? test. The differences in numbers and proportions of fronds of each developmental stage in the two populations were compared numerically. SIMAN & SHEFFIELD: POLYPODIUM VULGARE SPORULATION 33 |= outdoors mean number of fronds per pot 28 8 oe Cc = ee = = o 18/11/98 = 8 N = ~ Nn i=) & —~ oO = 30/12/98 8 3 10/02/99 24/02/99 10/03/99 Fic. 1. Changes in the mean number of fronds per pot in two genetically matched populations of Polypodium vulgare. The indoor aaa was placed in a controlled-climate greenhouse (mean day ek anarae 28°C, mean night temperature: 15°C, photosynthetically available radiation: ca 110 pmol m~? s~'); the outdoor seatiitten was left outside (U.K. natural weather conditions). The re- cruitment of new fronds in the indoor population occurred in three waves, one from October 1998 to mid-January 1999, the second from mid-January 1999 to late March 1999, and the third from early April 1999 to the end of the experiment. Error bars show standard error of the mean RESULTS The recruitment of new fronds from October 1998 to early June 1999 was significantly higher in the indoor population than in the outdoor population (x2-test, x = 127, df = 1, p<0.01) (Fig. 1). During this time the indoor popu- lation increased its number of fronds more than fourfold. The recruitment of new fronds took place in three waves (Figs. 1 and 2a), each of which increased the number of fronds by a factor between 1.5 and 1.7. The outdoor population increased its number of fronds by a factor 1.2 from October 1998 to June 1999 (Fig. 1). All recruitment of new fronds in the outdoor population took place from mid-April 1999 onwards. The increase in the number of fronds in the outdoor population continued during the summer months until mid Septem- ber 1999 (Fig. 3a). The majority of the new fronds recruited in the indoor population during the course of the experiment was fertile (Fig. 2a), so the indoor population maintained a high proportion of actively sporing fronds throughout the winter. ach wave of recruitment in the indoor population began with a sudden in- crease in the number of initially sterile fronds; a number which decreased as sori began to appear. The proportion of fronds that remained sterile throughout the wave decreased with each wave. Thus, at the end of January 1999 the proportion of sterile fronds was 37%, in early April 1999 28% of the fronds were sterile and in early June 1999 the proportion of sterile fronds was 17% 34 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) number of fronds 100 S86&8S88a88 percentages of fronds co eieresetege 33a 3 g BSses ede Sbaz2 23 5 date non at&éF4C aw FSS Ee Fic. 2. Sporing behaviour in a population of Polypodium vulgare kept in a controlled-climate bars represents i) sterile fronds and fronds with ii) green sori, iii) sporing sori and ix) empty sori, as indicated by the key in the figure. (a) Total number of fronds in the population (full bars) and number of fronds in each of the four groups (i.e. sterile, green, sporing, empty) for each week of the experiment. (b) Proportions of sterile, green, sporing and empty fronds, respectively, for each week of the experiment. (Fig. 2b). The frond mortality represented 7.8% of the total number of fronds gained over the experimental period and was entirely due to old fronds wither- ing and falling off. In the outdoor population there was no net recruitment of fronds during the winter and early spring. When the new fronds started to emerge, in late April, SIMAN & SHEFFIELD: POLYPODIUM VULGARE SPORULATION 35 Number of fronds 3 g co. raed 2S Percentages of fronds os 8 88 8 $ 3 co Fic. 3. Sporing behaviour in a population sa painir to vulgare kept outside in the Manchester University Experimental Grounds, U.K. (natural weather conditions) from October 1998 to Sep- tember 1999. The subdivision of the bars nanan i) sterile fronds and fronds with ii) green sori, iii) sporing sori and iv) empty sori, as indicated by the colour code key in the figure. (a) Total number of fronds in the population (full bars) and number of fronds in each of the four groups (i.e. sterile, green, sporing, empty) for each week of the experiment. (b) Proportions of sterile, green, sporing and empty fronds, respectively, for each week of the experiment. most of them soon turned into fertile fronds, so there was a steady increase in the number and proportion of fertile fronds over the summer (Figs. 3a and 3b). The proportion of sterile fronds decreased simultaneously so towards the end of the growing season in mid-September 1999, the proportion of sterile fronds was 17% (Fig. 3b), i.e. the same as for the indoor population at the end of its 36 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) third wave in early June 1999. The increase in number of fronds in the outdoor population was slightly impaired at two points (24/June/99 and 22/July/99) by herbivory from snails, but the fronds thus lost represented no more than 5% of the population. DISCUSSION It is clear that the conditions of a warm and illuminated glasshouse stimu- lated the vegetative growth and spore output of the P. vulgare plants. In the indoor population, the initial response to the glasshouse conditions was increased vegetative growth. At a time when the outdoor population stopped producing new fronds, the indoor population continued recruiting. Similar continuous growth has been observed in Pteridium grown in glass- houses (Wynn et al, 2000), but it is interesting that Thomson (2000) reports that Pteridium plants from four places (Honshu, Japan; Kiev, Ukraine; Bridg- ton, Maine, U.S.A. and Waterloo, Michigan, U.S.A.) require cold treatment (4°C) for four to eight weeks to ensure successful spring emergence of croziers when cultivated in the relative warmth of Sydney Royal Botanic Gardens (summer temp: max. 25.5°C, min. 18.2°C; winter temp: max. 16.8°C, min. 8.7°C). It seems that fluctuating temperatures warmer than those of the natural environment of a fern can have adverse effects on frond recruitment. The majority of the new fronds emerging in the indoor population of the present study became fertile, so a high proportion of fertile fronds was main- tained throughout the winter. Steeves and Wetmore (1953) concluded, after experiments with Osmunda cinnamomea, that the factors which determine fertility exercise their effects during the year before the leaves expanded. As- suming, in the present experiment, that each wave of recruitment created in the indoor population mimicked one growing season, we could suggest that the warm and bright indoor climate had an effect on the fertility of the second and third waves of new fronds. The proportion of fertile fronds in each of those two waves was higher than in its preceding wave. This may well be an effect of the enhanced nutritional status of the population, caused by a high production of photosynthate, which, transported as sugars to the bud primor- dia, might induce fertility, as suggested by Harvey and Caponetti (1972). In its natural environment P. vulgare produces ripe spores from July/August. The ripening of the spores is a gradual process, occupying a period of several months. Within a single sorus some sporangia shed early and others will take longer to ripen and shed later (Wright and Wright, 1999). The persistent pro- portion of 10-20% sporing fronds in the outdoor population from October 1998 to March 1999 is evidence of this behaviour. Each wave of recruitment indoors, as well as the single period of recruitment outdoors, (i.e. the growing season) increased the number of fronds by a factor of c. 1.5. This suggests that there were, at the beginning of the experiment, an equal number of dormant buds in the rhizomes of the populations and that the number of dormant buds an existing number of fronds can initiate for the next generation is restricted by something other than purely environmental SIMAN & SHEFFIELD: POLYPODIUM VULGARE SPORULATION 37 factors. This could explain the occurrence of the proportionally similar waves of recruitment. The present study shows that by enhancing the light and temperature it is possible to interrupt the strict reproductive cycle and induce continuous spor- ing ina pteridophyte population. Evidence of similar behaviour has recently been obtained in another study, in which dormant Pteridium rhizomes pro- duced fertile fronds within 13 weeks of being put into warm, well lit condi- tions (Wynn et al., 2000). Air samples taken in glasshouse and fernery envi- ronments in the UK do include ferns spores at all times of the year (e.g. Win- ston, 1998; Siman, 2000). This study suggests that transfer of plants to glasshouses could benefit fern spore collectors by inducing continuous sporing in plants. There are less welcome implications of fern spore production in indoor environments, how- ever, especially for species that do generate vastly more spores in glasshouse settings than those in natural environments. A glasshouse is an enclosed en- vironment with little chance of biological particles being blown away by winds. This means that there is a higher risk of inhalation of fern spores in a fern-rich glasshouse than in most places outdoors. Based on the growing body of evidence of toxic and allergenic effects caused by fern spores (as reviewed by Simén et al., 1999, see also Siman et al., 2000), we suggest that some pro- tective measures (e.g. face masks) should be taken by people who regularly work in or visit indoor fern-rich environments. ACKNOWLEDGMENTS authors wish to thank David Newton for looking after the plants in the Experimental Grounds, Gareth Ballance for standing in to take records in those weeks when we were away an the University of Manchester Research Support Fund and the F. C. Moore Studentship Fund for financing the study. LITERATURE CITED ALLsopp, A. 1964. The metabolic status and morphogenesis. saab aplaiatinigs 14:1-27. ALLSOPP, A. 1965. The significance for development of water supply, osmotic relations and nutri- Handbuch der Pflanzenphysiologie 15:504—552. Srewene G. F. 1896. The probable influence of disturbed nutrition on the evolution of the veg- etative phase of the sporophyte. Amer. Naturalist 30:353-357. Conway, E. 1957. Spore production in bracken (Pteridium aquilinum (L.) Kuhn). J. Ecol. 45:273- 284 DRING, M. Si ba The influence of shaded conditions on the fertility of bracken. Brit. Fern Gaz. 9:2 EVANS, a ie The nature of flower induction. Pp. 457-480, in L. T. Evans, ed. The induction of flov gear Macmillan of Australia, Melbourne. GOEBEL, K. 18 Uber kunstliche Vergrunung der Sporophylle von Onoclea struthiopteris. Ber. Deutsch. ae Ges. 5:69. GOEBEL, K. 1905. Onecnouraglhiy of Plants. Part II. (English ed.). Oxford University Press. GoEBEL, K. 1908. Einleitung in die experimentelle Morphologie der Pflanzen. Druck und Verlag von B. Me ype Leipzig "sheng K. 1928. Organographie der Pflanzen. Teil I. png Jena. S, A. ae oo, Hort. Week 28-Mar 1996:2 38 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 1 (2002) Harvey, W. H., AND J. D. CAPONETTI. 1972. In vitro studies on the induction of sporogenous tissue on leaves of cinnamon fern. I. Environmental factors. Can. J. Bot. 50:2673—2682. LABOURIAU, L. G. 1958. Studies on the initiation of sporangia in ferns. Arq. Mus. Nac. 46:119-202. PAGE, Cc. = 1976. The taxonomy and phytogeography of bracken—a review. Bot. J. Linn. ae 75s) Boy Ay e 1979. Experimental aspects of fern ecology. Pp. 105-140, in A. F. Dyer, ed. The ex- perimental papas of ferns. Academic Press, London bance E. 1996. From pteridophyte spore to lena it in the natural environment. Pp. 541— ar aa and R,J. Johns, eds. Pteridology in perspective. Royal Botanic Gardens, SIMAN, . E. 2000. Fern spores and human health. PhD dissertation. University of conan eee SIMAN, S. E., A. C. POVEY, AND E. SHEFFIELD. 1999. Human health risks from fern spores?—a revie 87. SIMAN, S. E, A.C. POVEY, ra Warp, G.P. MARGISON AND E. SHEFFIELD. 2000. Fern spore extracts can damage DNA. Brit. J. Cinice 83:69-73. STEEVES, T. A. 1959. An eriieeanias al two forms of Osmunda cinnamomea. Rhodora 61:223— 230. STEEVES, T. A., AND R. H. WETMORE. 1953. Morphogenetic studies on eae cinnamomea L.: some aspects . “ general morphology. Phytomorphology 3:339 Sussex, I. M., AND T. A. STEEVES. 19 ie + aust on the control of ya of fern leaves in sterile culture. ot Gaz. 119:203-— spas J. A. Morphometric and genomic diversity in the genus Pteridium (Dennstaedtiaceae). n. Bot. 85:77-99. Waou E W., AND D. N. SHARMA. 1963. Experimental and analytical studies of pteridophytes. Bot. ne 101-121, eho D. 1998. Risk assessment of environmental exposure to fern spores. M.Sc. dissertation, hota! of Mancheste WRIGHT, B., AND A. WRIGHT. om The “Wright” way to collect and clean fern spores. Pteridologist 3:62 ote Wynn, J. M., J. L. SMALL, R. J. PAKEMAN, AND E. SHEFFIELD, 2000, An assessment of genetic and envixonmental effects on sporangial development in img (Pteridium aquilinum (L.) Kuhn) using a novel quantitative method. Ann. Bot. 85:113-115. 3 Oe on re » aang plot yh? > Elliowsen: 7 pk . oar - 7 : iJ ae des 7 f : : - - i 7 = a, ar cam’ - ; Vai _ - ie 7 aa 4 : iat tae. yy, = - a a : = : _ ae ek ae wD SS ae Fe es Se a hcinn ag te > a a _ =, ee Sol 3 Cae 3 : ; x ae Reto al ie ara: a! : x wo ee ee hier - f 7 7 / ; ; 7 - ir =f — ‘ Pp = | - 7 a 7 ele ene ih le A ite a i 2 a fe. oe on a tanical Garden Libra I IN Authors are encouraged to submit manuscripts pertinent to pteridology for pub- lication in the American Fern Journal. Manuscripts should be sent to the Editor. Acceptance of papers for publication depends on merit as judged by two or more referees. Authors are encouraged to contribute toward publishing costs; however, the payment or non-payment of page charges will affect neither the acceptability of manuscripts nor the date of publication. Authors should adhere to the following guidelines; manuscripts not so prepared may be returned for revision prior to review. Submit manuscripts in triplicate (xerocopies acceptable), including review copies of illustrations and originals of illustrations. After review, submission of final versions of manuscripts on diskette (in PC- or Mac-compatible formats) is strongly encouraged. Use standard 8% by 11 inch paper of good quality, not “erasable” paper. Double space manuscripts throughout, including title, authors’ names and addresses, short, informative ab- stract, text (including heads and keys), literature cited, tables (separate from text), and figure captions (grouped as consecutive paragraphs separate from figures). Arrange parts of manuscript in order just given. Include author’s name and page number in upper right corner of every sheet. Provide margins of at least 25 mm all around on typed pages. Do not submit right-justified copy, avoid footnotes, and do not break words at ends of lines. Make table headings and figure captions self-explanatory. Use S.I. (metric) units for all measures (e.g., distance, elevation, weight) unless quoted or cited from another source (e.g., specimen citations). For nomenclatural matter (i.e., synonymy and typification), use one paragraph per basionym (see Regnum Veg. 58:39—40. 1968). Abbreviate titles of serial publi- cations according to Botanico-Periodicum-Huntianum (Lawrence et al., 1968, Hunt Botanical Library, Pittsburgh) and its supplement (1991). References cited only as part of nomenclatural matter are not included in literature cited. For shorter notes and reviews, omit the abstract and put all references parenthetically in text. Use ae Herbariorum (Regnum Veg. 120:1—693. 1990) for designations of her- bari icatetik should be proportioned to fit page width with caption on the same page. Provide margins of at least 25 mm on all illustrations. For continuous-tone illustrations, design originals for reproduction without reduction or by uniform amount. In composite blocks, abut edges of adjacent photographs. Avoid com- bining continuous-tone and line-copy in single illustrations or blocks. Coordinate sequence and numbering of figures (and of tables) with order of citation in text. Explain scales and symbols in figures themselves, not in captions. Include a scale and reference to latitude and longitude in each map. Proofs and reprint order forms are sent to authors by the printer. Authors should send corrected proofs to the editor and reprint orders to the printer. Authors will be assessed charges for extensive alterations made after type has been set. For other matter of form or style, consult recent issues of American Fern Jour- nal and The Chicago Manual of Style, 14th ed. (1993, Univ. Chicago Press, Chicago). Occasionally, departure from these guidelines may be justified. Authors are encouraged to consult the editor for assistance with any aspect of manuscript preparation. Papers longer than 32 printed pages may be sent to the Editor of Preridologia (Memoir Editor, see cover 2). PTERIDOLOGIA ISSUES IN PRINT The following issues of Pteridologia, the memoir series of the American Fern Society, are available for purchase: 1. Wagner, David H. 1979. Systematics of Polystichum in Western North America North of Mexico. 64 pp. $10.00 postpaid. 2A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Choc6 (Part 1: Psilotaceae through Dicksoniaceae). 364 pp. $32.00 postpaid. Send your order with a check or money order to: American Fern Society, Inc., c/o U.S. National Herbarium MRC-166, Smithsonian Institution, Washing- ton, DC 20560. AMERICAN FERN JOURNAL ON MICROFICHE Volumes 1-61 of the American Fern Journal are available as archival quality, silver positive microfiches. Single volumes or the entire run may be purchased. The fiches are easily read with 10X or greater magnification (using a dissecting microscope and transmitted illumination or a fiche reader). Silver negative micro- fiches of vols. 1-50 are also available. The price is $4.00 per volume or $244.00 per set of 61 volumes, postpaid. Send your inquiry or order with a check or money order to: American Fern Society, Inc., c/o Dr. James D. Montgomery, Ecology III, Inc., R.D. 1, Box 1795, Berwick, PA 18603. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://www.amerfernsoc.org/ ee AMERICAN eee FERN aici J O UJ R N A 1 April-June 2002 IN MEMORIAM: WARREN HERBERT WAGNER, JR. Obituary: Warren H. Wagner, Jr. (1920-2000) Donald R. Farrar 39 Bibliography of Warren Herbert Wagner, Jr. David B. Lellinger 50 The Mating Systems of Some stele Polypodiaceae n-Liang Chiou, Donald R. Farrar and Tom A. Ranker 65 Belowground enseicras and Abundance of Botrychium Gametophytes and Juvenile Sporoph ay Cindy Johnson-Groh, Chandra Riedel, Laura Schoessler and Krissa Skogen 80 Additions to the Fern Flora of Saba, Netherlands Antilles David B. Lellinger 93 Taxonomic Notes on Hawaiian Pteridophytes Daniel D. Palmer 97 Novelties in Pteridaceae from South America Jefferson Prado and Alan R. Smith 105 Is Gametophyte Sexuality in the Laboratory a — Predictor of Sexuality in N. m A. Ranker and Heather A. pnd 112 Additional Support for Two Subgenera of Anemia (Schizaeaceae) from Data for the Chloroplast Intergenic Spacer Region eo and Morphology Skog, E. A. Zimmer and J. T. Mickel 119 Intrafamilial Relationships of the Thelypteroid cy (Thelypteridaceae) n R. Smith and Raymond B. Cranfill 131 Two New Species of Moonworts (Botrychium subg. Botrychium) from Alaska Mary Clay Stensvold, Donald R. Farrar and Cindy Johnson-Groh 150 Isoétes X herb-wagne ri, an [ pecific Hybrid of I. bolanderi x I. echinospora o (Isoétaceae) W. Carl Taylor 161 Botrychium alaskense, a New Moonwort ea the Interior of Alask Warren Herb Wagner, Ir and Jason R. Grant 164 A New Name for an Old Fern from North Alabama James E. Watkins, Jr. and Donald R. Farrar 171 Continued Pteridophyte Invasion of Hawaii Kenneth A. Wilson 179 The American Fern Society Council for 2002 CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS erase 2016. resident TOM RANKER, University Museum, Campus Box 265, University of Colorado, Boulder, CO pa 0265. Vice President W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI 53233-1478. Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916-1110. Treasurer GEORGE YATSKIEVYCH, Missouri Botanical Garden, P. O. Box 299, St. Louis, MO 63166-0299. Membership Secretary JAMES D. MONTGOMERY, Ecology II, R.D. 1, Box 1795, Berwick, PA 18603-9801. Back Issues Curator R. JAMES HICKEY, Botany Dept.,; Miami rcebie Oxford, OH 45056. Journal Editor DAVID B. LELLINGER, U.S. National Herbarium MRC-166, Smithsonian Institution, Washington, DC 2 0560-0166. Memoir Editor CINDY JOHNSON-GROH, Dept. of ae oS Adolphus College, 800 W. College Ave., St. Peter, MN 56082- Bulletin Editor American Fern Journal EDITOR R. JAMES HICKEY Botany Department, Miami University, Oxford, OH 45056 ph. (513) 529-6000, e-mail: hickeyrj @ muohio.edu ASSOCIATE EDITORS GERALD J. GASTONY ~..00.....6.. uae of ae Indiana Sa ee IN 47405-6801 RISTOPHER H. HAUFLER .... Dept. of Botany, University of Kansas, Lawrence, KS 66045-2106 ROBBIN C. MORAN New York Botanical Gaaa hie NY 10458-5126 JAMES H. PECK Dept. of Biology, University of Arkansas—Little Rock, 2801 S. University Ave., Little Rock, AR 72204 ian “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly vero to the Asti study of ferns. It is owned by the American Fern Society, and published at The Am n Fern Soc % Missouri Botanical ana P.O. Box 299, St. Louis, MO 63166-0299. Peciodicgle postage oar. é St. Louis, MO, and addition — aims for missing oe potste 6 months semica pry to ks fr en (foreign) after the date of i ee rders for pg oom should be addressed to Dr. J . Montgomery, Ecology II, R.D. cece PA 1860: i Changes of sees dues, and applications for membership should be sent to the Membership tary. General inquiries concerning ferns should be addressed to the Secre Subscriptions $20.00 gross, $19.50 net if paid through an agency (agency fee $0.50); sent free to members of the American i ociety (annual dues, $15.00 + $5.00 mailing surcharge beyond U.S.A. Back volumes are available for most years as printed issues or on microfiche. Please contact the Back ae Curator for prices and availability. TMASTER: Send address changes to AMERICAN FERN JOURNAL, 326 West St. NW., Vienna, VA SieeateL FIDDLEHEAD FORUM The editor of the Bulletin of the American Fern Society welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on SPORE Ex: n McDaniel, 1716 Piermont Dr., Hacienda Hts., CA 91745-3678, is Director. Spores cinhanea set lists of available spores sent on request fe / html GIFTS AND ccs enable bed to — its eervwes to members. and to = eerened MISSOURI BOTANICAL & :39-49 (2( Acs American Fern Journal 92(2 AUG 2. 9 2002 GARDEN LIBRARY Obituary: Warren H. Wagner, Jr. (1920-2000) DONALD R. FARRAR Department of Botany, Iowa State University, Ames, IA 50011 Warren Herbert Wagner, Jr. was born on 29 August 1920 and raised in Washington D.C., the son of Warren Herbert Wagner and Harriet Claflin Wagner. His early interests in natural history took him. frequently to the Smith- sonian Institution, where he became acquainted with the experts, including 40 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) the eminent pteridologists William R. Maxon and Conrad V. Morton. In college at the University of Pennsylvania, he became the enthusiastic field compan- ion of Edgar T. Wherry, author of the “The Fern Guide.”” Wherry was a mineralogist who became an expert on fern habitats and the first to point out the important associations of epipetric ferns with particular rock types. Graduating from the University of Pennsylvania in 1942, Herb entered the U.S. Naval Air Corps, serving first in the Atlantic, then in the Pacific Fleet, where he was a Naval Air Navigator. In the Pacific islands he spent his off- duty hours collecting ferns and butterflies, later publishing (with David Grether) ‘“Pteridophytes of Guam”’ (1948) as well as articles on the pterido- phytes and butterflies of the Admiralty Islands. During this time he also flew into California, taking his specimens to E. B. Copeland at Berkeley, re- nowned expert on Philippine ferns. This was the beginning of an association that would bring him back to Berkeley for graduate study. While in the Navy, he also began what was to become a life-long study of the ferns of the Hawaiian Islands. At the University of California—Berkeley, in 1945, Herb joined student colleagues Charles Heiser, Ernest Gifford, Frank Ranzoni, Vern Grant, Art Krukeberg, and others in courses instructed by G. Ledyard Stebbins, Adriance Foster, and Herb’s major professor, Lincoln Constance. Copeland, although retired, was still active and advised informally on Herb’s research. Also among Herb’s student colleagues in systematic botany was Florence Signaigo. Herb and Florence married in 1948 and began a lifelong productive partnership in research and publication. After receiving his Ph.D. in 1950, Herb spent a year as a Gray Herbarium Fellow at Harvard, then moved to the University of Michigan in 1951, where he remained throughout his illustrious career in teaching and research. He chaired or co-chaired the graduate programs of more than 45 Ph.D. students. He taught a variety of graduate and undergraduate courses, including his highly popular “Systematic Botany” and “Biology of Woody Plants,’ both of which he continued to co-teach after ‘‘retirement” (1991) through the fall of 1999. Herb’s public lectures and seminars were equally popular. Few biolo- gists have been in such demand as a visiting speaker. His curriculum vitae list of “invited lectures” totaled 169—since 1992! From 1966 to 1971 Herb served as Director of the University of Michigan’s Matthaei Botanical Garden. His popularity with garden clubs, amateur bota- nists, and conservation groups made the Gardens a center of outreach activities in this arena, as well as a center for research and display of con- servatory plants. He chaired the Department of Botany in the Division of Bio- logical Sciences from 1974 to 1977 and chaired many additional department and college committees, including the University of Michigan’s Tropical Studies Committee from 1983 through 1997. He was chairman or president of nine professional societies, including the American Fern Society, American Society of Plant Taxonomists, and Botanical Society of America, and council member/trustee/advisor to dozens of organizations. He served as an editor for the Flora of North America, co-editing the Pteridophytes in volume two. FARRAR: WARREN H. WAGNER, JR. 41 As reviewer of countless journal manuscripts, grant proposals, and botany/ biology programs, he gave freely of his time while continuing to teach and maintain a research program generating over 240 research publications. “Probably the best-known botanist ever to work at the University (of Michigan),’”’ Herb Wagner’s influence on his science was immense. He has been referred to as “the founder of modern day systematics” in reference to his seminal contributions to cladistic analysis of phylogenetic relationships. His studies in the recognition of species hybrids, their value in understand- ing species relationships and their role in speciation and evolution became classics. His knowledge of ferns worldwide was phenomenal and allowed unique insights into fern ecology and life history, as well as systematics. Along with awards recognizing his contributions to systematic biology (including Willi Hennig Fellow, National Academy of Science, and the Asa Gray Award of the American Society of Plant Taxonomists), his name is indelibly inscribed in cladistic literature via the universally recognized ‘‘Wagner Tree’’ representation of phylogenetic relationships. In writing about the career of Edgar Wherry, Herb stated that Wherry “was one of those rare individuals—a real naturalist.’’ Extended as the highest of compliments, the same could be said of Herb Wagner. His had an exception- ally keen ability to observe the small and intricate details of plants interact- ing with their environment. Study the knowable. Accumulate information on the parts. With time, dedication, and an open mind, the big picture will emerge. These gems of Herb’s philosophy attributed value to all research, no matter how small the project or whom the researcher. All good data were worth getting excited about—and he did. Herb’s ability to inspire others through his interest in their studies and their knowledge not only fostered independent research, but also created a legion of professionals and amateurs eager to contribute data to Herb’s projects. The total productivity of this syn- ergism, though unquantifiable, remains hugely visible. Herb maintained a rigorous schedule of teaching, research, invited lectures and symposia presentations until just weeks before his death on January 8, 2000 at the age of 79. He was very much looking forward to participating in the 2000 Botanical Society of America symposium on the ‘Biology and Conservation of the Ophioglossaceae’’ that he helped to organize. In the summer of 1999, Herb and Florence conducted field work in Alaska and in southwestern Canada. From both places they returned with, of course, new species of Botrychium. The foregoing highlights of the career and achievements of Warren ‘‘Herb”’ Wagner fall short in communicating the extraordinary nature of his per- sonality, his gift for teaching, and his full influence on pteridology and pterido- logists of the last half-century. The more personal reminiscences that follow portray a rare individual whose life we are all so fortunate to have shared. ACKNOWLEDGEMENTS Information on the early years of Herb’s career were graciously provided by Florence Wagner. Factual information is taken from Herb Wagner’s 1999 curriculum vitae. Other observations 42 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) extend from my long association with the Wagners, as a graduate student in Ann Arbor and in many subsequent field trips and discussions of plants, people, and philosophy. Portions of this article are excerpted with permission from Taxon 49:585-592. August 2000, “Warren H Wagner, Jr. (1920-2000),” by Donald R. Farrar. Quotations are by Julian P. Adams and William R. Anderson of the University of Michigan in The University Record 55:2-4. January 17, 2000, ‘Warren H. (‘Herb’) Wagner,” by William R. Anderson. REMINISCENCES Meeting Herb Wagner.—I first met Warren Herb Wagner, Jr., on the eve- ning of 13 June 1980, at Flathead Lake Biological Field Station in far north- western Montana. This meeting has stuck in my mind, perhaps because in many respects it typified my relationship with him. I was a new student in pteridology and had come to Flathead Lake to take his (and Florence’s) four- week-long fern course. Earlier that year I had phoned Herb to tell him that I had enrolled in his course and to let him know about a new, unnamed fern hybrid I had found in the Shawnee Hills of southern Illinois, a cross be- tween the walking fern (Asplenium rhizophyllum) and the maidenhair spleenwort (A. trichomanes). As it turned out, Herb was a few days late for the course because of a conflict with the final days of the International Bota- nical Congress in Vancouver. He finally arrived at the biological station about 10:30 p.m., when I was alone in the lab and identifying plants. The lab door suddenly opened and Herb burst into the room. He walked about half its length before acknowledging my presence. We introduced ourselves. He explained that he had been collecting moonworts all day en route to the field station, but it didn’t seem like it to me. Instead of being exhausted after a long day of fieldwork, he was animated and lecturing me about the biology of moonworts and ferns in general. I listened spellbound and thrilled that he would take the time to explain it all to me. “Let’s see your hybrid fern,” he demanded. “It’s in my cabin,” I replied, ‘and my cabin is a long way away, and it’s pitch-black outside, and I’ve lost my flashlight. Can I show it to you tomorrow?” Then—and I'll never forget this—he gave me the most odd sup- plicating look, and with hands clasped together as if praying, he said, ‘““Won’t you please go get it—now!”’ How could I refuse? I stumbled back to my cabin in darkness, found the specimens, and retraced my steps to the lab. Herb examined the specimens. “Yes, that’s it! Asplenium rhizophyllum xX A. trichomanes. Congratulations!’” He shook my hand; I had received his imprimatur. Then, without further ado, he described how he wanted the lab rearranged—by me, that is. Tables, benches, plant presses, cardboard divid- ers, microscopes—all these were to be moved according to a plan that he had devised while I was retrieving the specimens. After explaining where every- thing should be relocated, he said, ‘‘Ok, gotta blow!” and he left the lab as abruptly as he had entered. I was once again alone in the lab, excited by what I had just learned about moonworts and spleenworts, and too motivated to mind having been pressed into service.—Robbin C. Moran, The New York Botanical Garden, Bronx, NY 10458-5126. FARRAR: WARREN H. WAGNER, JR. 43 Herb at ‘“‘Bug Camp’’.—My first memory of Herb Wagner is from Spring, 1952. During my senior year at the University of Michigan, about to graduate with a major in Biological Science, I worked as an undergraduate lab assis- tant in botany. That job involved lugging five-gallon bottles of distilled water, spraying the greenhouse for insects, and other tasks. One day I was called in to meet a new faculty member, Dr. Warren H. Wagner, Jr. He asked me to be his graduate assistant that summer at the bug camp, the University of Michigan Biological Station at Douglas Lake. With no other plans, I agreed That marked my beginning as a botanist. I assisted Wagner in courses in Phycology and Pteridology. To be a graduate student, I had to take a course, so I signed up for an independent study with him. Since the genus Equisetum was well represented around the bug camp, Wagner assigned me to work with its taxonomy. Herb showed himself to be an enthusiastic, knowledgeable field botanist. I well remember, after a hard day bent in a half-crouch searching out Botrychium, stopping with him for ice cream on the way back to Douglas Lake. And, I remember evenings all of us gathered around a piano to hear him play. Most evenings were spent in the lab until 11 p.m. or so, working with Herb, on the materials gathered during the long days in the field. That summer, Herb advised me to go to some other university for my mas- ter’s degree, to widen my exposure to the whole field of botany, with the understanding that I would return to Michigan to work for my doctoral de- gree under him. I went off to Florida State University as a graduate research assistant on a tidal marsh study, then to the University of California, to study the anatomy of Equisetum stomata with Dr. Adriance Foster and get an M. A. in Botany. After a stint in the U. S. Army, I returned to Ann Arbor, to my mentor, Dr. Warren H. Wagner, Jr., and to the taxonomy of Equisetum.—Richard L. Hauke, 900 N. Stafford St., Apt. 1103, Arlington, VA 22203-1844. Unflappable Herb.—If you know something of Herb’s activities during World War II, it is no surprise that he was unflappable. He had been an offi- cer in the U. S. Navy, a navigator on air flights, and for a time was based on Guam. Like the pilots, he was not called upon to fly every day; on his days off he collected butterflies and ferns. After the end of the war, he and David F. Grether published an account of the pteridophytes of Guam based princi- pally on their collections and those of other U. S. military personnel. (Occ. Pap. Bishop Mus. 19(2):25-99. 1948). He was a native of Washington, DC, and his butterfly collection was given to the Smithsonian Institution. The best collecting for butterflies, Herb found, was beyond the American defense lines that were maintained by Marine guards and patrols. Some well armed Japanese troops were still at large beyond those lines, along with numbers of their fallen colleagues who were very attractive to butterflies. Because of the danger, the Marine guards didn’t like Herb’s excursions, but as an officer, he was able to go where he wanted. A butterfly net was of the utmost importance to those excursions. It was the equivalent of a white 44 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) flag that signaled the bearer’s non-combatant status. Even so, there must have been a considerable element of danger, although Herb said he was never shot at by the Japanese while carrying a net. (A plant press, even loaded with ferns, did not confer the same protection, presumably because its function was unknown to the Japanese soldiers. In Ann Arbor, Herb’s spring wildflower course was always popular. His humor, knowledge, and enthusiasm were delightful. Who could forget his earnest description of a hand lens and how it was to be used and worn around the neck? He then pulled from within the lectern a large hand lens attached to a red ribbon so wide that it was practically a sash! Although Herb was generally unflappable, I do remember one exception that was, de- pending upon your point of view, hilarious. The year I assisted him, his lectures were presented in three classrooms that had been converted into a long, narrow lecture hall. The slide projector, which I manned, was placed well back in the hall, perhaps 40 feet from the desk, lectern and screen at the front. In order to signal for a change of slides, Herb would hold a wooden pointer vertically and drop it dramatically on the concrete floor next to his feet. Although none of us knew it at the time, the repeated thunk was not popular with the Dept. of Geology professors who had offices on the floor below. One afternoon, Herb called for darkness. The shades were drawn while Herb turned out the room lights. I turned on the projector and put the first slide on the screen. Naturally, I was looking at the projector when Herb drop- ped the pointer to request the second slide. There was a tremendous det- onation in the front of that quiet room. My head snapped up in time to see Herb still coming down behind the desk after what must have been a terrific leap, nearly a pole vault. He strode over to the door, clicked on the lights, took out and lit with trembling hands one of the small cigars he smoked at that time, inhaled deeply, and said ‘‘Take a five-minute break.” Of course, the students could not contain themselves. By evening, most of the graduate students knew what had happened: one of them had taped a few caps from a cap pistol to the blunt end of the pointer. After the recess and cigar, Herb’s lecture and slides continued as before, punctuated only by an ordina thunk. The Geology professors were not at all amused.—David B. Lellinger, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0166. Graciousness and Support.—Like most graduate students back in the 1970's, it wasn’t long before beginning my study of ferns that I was introduced to Herb Wagner. My first impressions were that he was brilliant, dynamic, mesmerizing, and exactly the type of scientist that I wanted to be. Of course, that wasn’t possible because he was certainly one of a kind. He reviewed one of the first papers that I wrote concerning meiotic studies in Ceratopteris. I remember that the review was less than favorable about some aspects of the study and of the way that it was presented. He was right, of course, but for some reason, perhaps extensive re-writing, it eventually was published. FARRAR: WARREN H. WAGNER, JR. 45 A couple of years later, | wrote a paper that made an attempt to propose an alternative explanation for the pattern of reticulate evolution that Herb had described for Appalachian Dryopteris. This was the last piece of my doctoral dissertation. I soon graduated and while I was at work as a young assistant professor, I actually received a reprint request for this paper from Hugh Iltis with a big “‘congratulations”’ scrawled on it! But any glory was to be short-lived. A couple of years later, during a meeting of the Southeastern fern group at the Duke Marine lab, I remember sitting quietly in the front row as Herb methodically destroyed my brash interpretation of Appalachian Dryopteris. | was humbled, but not humiliated. He was far too gracious for at. Although I was not one of his students in any type of academic lineage, he was a constant and helpful influence in my career. Letters of recommen- dation for jobs and promotions, reviews of my academic department, and oc- casional suggestions spanned the course of over 20 years. My last personal encounter with him was during a visit to Knoxville in the late 1990’s where he gave an inspiring seminar to a group of faculty and students. I was teach- ing a sophomore genetics class that semester and made a point of inviting my students to hear his seminar. Although they weren’t botany majors and generally had very little interest in plants, it was obvious that he had not lost his charm, dynamism, and ability to mesmerize even these supposedly disinterested young people in his story about moonworts! I keep a small pic- ture of Herb behind me in my office and often turn and sneak a peek at him for inspiration when I am trying to figure out how to do a better job of pre- paring a lesson or writing up a laboratory exercise.—Les Hickok, Department of Botany, The University of Tennessee, Knoxville, TN 37996. Humanity and Professionalism.—He combined intellect with a humanity and friendliness. Herb Wagner, as we called him when he befriended us, traveled widely and infected nearly everyone he met with an enthusiasm for life and for inquiry into the natural world. His influence upon me came early in my professional career. While a stu- dent at the University of Kentucky between 1961 and 1963, Herb had already visited the campus, and as always, took a field trip into the Kentucky hills. I was told that an old man walking along a road of a cove yelled, ‘“‘Hey, what are you fellers doing down there?’ Herb’s immediate response, ‘‘We’re just botriculating.”” And though suspicious of revenuers, the old man responded, “Oh well, go right ahead.” It was shortly after I arrived in Cullowhee in 1969 that Herb made a visit to Highlands Biological Station in preparation for the fern conference in 1970. He had come across a paper that Jim Horton and I published listing new county records that included Asplenium heteroresiliens, a hybrid on marl known only from the coast. He visited our herbarium and in his in- structive but gentle way said, ‘I see why you identified Asplenium tricho- manes for heteroresiliens, but it isn’t.” Our somewhat depauperate and poorly dried specimen didn’t exactly match either species. Herb also got me 46 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) to be more careful in my identification of other species, pointing out that what appeared to be Populus grandidentata growing beside the science build- ing was something else, which we later identified as Populus X canescens. In the early 1970’s Herb was President of the American Fern Society, pop- ularizing the profession. About this time I had a free evening and took in the movie, “A New Leaf,” a spoof of a millionaire and his attraction to a wall- flower type of lady who happened to be a student of ferns. To my amaze- ment, she came to the point of verifying her identity that the tropical fern she discovered in Alaska was confirmed by ‘‘Dr. Wagner at the University of Michigan.” Such was his fame as it expanded beyond the ivory tower into society in general. Herb visited Cullowhee many times. On his last trip here in April, 1997, not only did he give the visiting scholar lectures but also wanted to return to some nice place before he returned home. His 4 p.m. seminar that Friday afternoon was so well received that a couple of visitors, Bob Dellinger, local naturalist, and Gary Kauffman, USDA Forest Service botanist, asked to tag along with us the next day to the Joyce Kilmer area. Although Herb had in- dulged in country ham the previous days (he was supposed to be on a low salt diet) and was a bit tired from the activities the previous two days, he de- lighted in finding Carex austro-caroliniana with peduncles several centi- meters long. Such was his exuberance over those things he found fascinating.—Dan Pittillo, Department of Biology, Western Carolina Univer- sity, Cullowhee, NC 28723. In His Element.—When | think of Herb, I tend to think of him in a field trip setting. For instance, when he was invited to the annual meeting of the Field Botanists of Ontario in Midland, Ontario a number of years ago, I re- member the lead car leaving the sandy gravelly country road and entering an open sandy field with a few copses of junipers. In what resembled a “cops and robbers” film on T.V., the lead car had not come to a full stop when Herb rolled out the front door of the car. He took a giant step towards the copse and was almost immediately flat on the ground under the prickly Juni- per. Suddenly there was a victory call for all and sundry within 200 yards— “Botrychiums.” All Herb’s followers were on the spot in a few seconds and each looked in marvel at some very small specimens between the juniper and the sand—a gap of less than a foot. Herb was in his element in the field, and his disciples enjoyed every minute of it!—D. M. Britton, Professor Emer- itus, Department of Molecular Science & Genetics, University of Guelph, Guelph, Ontario, N1G 2W1 Canada. One of the All-time Giants.—My first contact with Herb Wagner came about in early 1978 when I was at the University of Cambridge, England. I had been intrigued for some time by a note in Herb’s paper on spores in rela- tion to fern phylogeny (Ann. Missouri Bot. Gard. 61: 346. 1974) on the pres- ence of large spores in ferns in the higher altitude rainforests of Hawaii not being linked to polyploidy. I had noted that in the New Zealand species of Grammitis (Grammitidaceae) there is a trend towards the production of FARRAR: WARREN H. WAGNER, JR. 47 larger spores, and frequently fewer spores per plant, in the species occurring at higher altitudes and latitudes. Although few chromosome counts are avail- able this does not seem to be linked to polyploidy. I had assumed that this wasn’t a parallel of the strategy of higher plant taxa that at higher altitudes produce fewer and larger seeds to contain resources needed during longer periods of dormancy, because dormancy is not an issue with Grammitid spores, which are photosynthetic. Perhaps Herb’s ‘‘selection for precinctive- ness’’ was working here? So I wrote to him concerning my observations; he promptly replied that he thought it was, and that in Hawaii, for example, any spore or any propagule that is especially likely to be carried by wind is going to lose out. His observations on butterflies had some part in his logic, with the most conspicuous Hawaiian butterfly being a very sluggish inhabi- tant in mountain valleys, rather than ridges, where it was likely to be blown away. I’m still very much undecided about selection for precintiveness in Grammitidaceae after further work, as the high altitude New Guinea species do not show the same increased spore size as the New Zealand ones, but Herb’s prompt, encouraging and discursive response to a junior unknown pteridological correspondent made a deep and lasting impression on me that here was one of the all-time giants of pteridology. Nine years later, when I had moved to the Royal Botanic Gardens, Kew, I finally met both Herb and Florence at the International Botanical Congress in Berlin in 1987, and the impression I had of Herb from his correspondence was strongly reinforced. He enjoyed ferns and people and talk, and I really envied all of his students for having such a stimulating mentor! A talk with Florence at Berlin turned up a surprising coincidence—one of my M. Sc. supervisors at the University of Auckland, New Zealand, Prof. Jack Ratten- bury, had been best man at Florence and Herb’s wedding—it’s a very small botanical world.—Barbara Parris, Fern Research Foundation, 21 James Kemp Place, Kerikeri, Bay of Islands, New Zealand. Tutor and Friend.—Herb Wagner was my tutor and friend during the last dozen years of his life and he assisted me in gaining the skills needed to study the Hawaiian ferns. Herb worked with the Hawaiian fern flora for more than half a century, during which time he greatly increased the know]l- edge of its diversity and biology. He described more than three dozen new species, varieties, and hybrids. With his contributions, our organized under- standing of the Hawaiian ferns was much advanced. Herb made several long visits to Hawaii during which he went on exten- sive field trips and spent much time in local herbaria. During these visits he taught courses reviewing the Hawaiian ferns and took his students on field excursions. A remarkable and enthusiastic teacher, the passionate sharing of his deep knowledge of Hawaiian ferns inspired many in Hawaii to study this neglected group of plants. His lectures and seminars were remarkable for his enthusiastic, thorough, and effective presentation of material to students with only a minimal knowledge of pteridophytes. I will remember the ballet like pirouettes he would do at the blackboard when emphasizing a point, 48 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) and I remember the vast collection of anecdotes and jokes he used to enliven a lecture and to help students remember the subject under discussion. He would not let the decreasing physical vigor he experienced in later years interfere with active field work. On his return to Ann Arbor from Hawaii he always left behind a whirlwind of enthusiasm for the study of Hawaiian ferns. Florence Wagner’s collaboration and organizational abilities allowed the Wagner family to effectively pursue the twin goals of teaching and research. Many of us in Hawaii will fondly remember a day’s end with Herb and Florence, where a good dinner, cocktails, stimulating conversation, and en- joyable company combined to make for a very pleasant evening. He is missed by all who knew him in Hawaii. His stimulating visits will be missed and the time he spent here will be remembered here with nostalgia.—Daniel D. Palmer, 1975 Judd Hillside, Honolulu, HI 96822. Missing Herb.—I miss Herb too much to find words to describe how misera- ble I am at knowing he is not at the end of the phone needing a this or a that, or welcoming my weird plant ID, or to confirm one of those peripheral vegeta- tive idioblastic forms that my heart hoped might be something new. I remem- ber the summers he visited Iowa Lakeside Lab, the AFS field trips, particularly the one in northern Michigan. I did not know Herb from Michigan as many did, as he was my academic grandfather—and father in Iowa by remote-control, a fitting way to be for us hybrid studying types. He thought that my following Darwinian fates of individuals and natural history approach to fern reproduction and population structure was significant when most around me spoke only of things homoeologous, bar codes of molecules on gels, boot- strapping clades, and such. In comfort, Herb offered that he often no longer knew what cladistic people meant with the language they used to describe his ground-plan divergence method he created so long ago. I only wanted to be a teacher, and Herb’s way with a gathering of students still is my exemplar. We will all miss him; he will always be with us.—James H. Peck, Department of Biology, University of Arkansas—Little Rock, Little Rock, AR 72204. A Personal Gift.—I didn’t go to the University of Michigan to study with Herb Wagner. My interests were in plant physiology, but in my first semester a fellow graduate student, Bob Faden, who was then a student of Herb’s, secured an invitation for me to accompany the annual Wagner entourage to the hills and canyons of southern Ohio and Kentucky. Herb had just pub- lished papers on the independent occurrence of gametophytes of Vittaria and Trichomanes in the eastern United States. We had no trouble finding great mats of Vittaria gametophytes and, with some squirming on our backs under overhanging sandstone cliffs, we also found the green, threadlike gametophytic filaments of Trichomanes. Returning with collections of these intriguing plants, I was thrilled and convinced that with my growing physiological expertise I would coax them in culture into producing sporo- phytes and would resolve the mystery of their independent (without sporo- phyte production) occurrence in the wild. Nearly 40 years later I still work on that. FARRAR: WARREN H. WAGNER, JR. 49 After two years of graduate school I was having doubts about my future in academia. With considerable trepidation I informed Herb of my intention to take some time off and join the Peace Corps. To my surprise, he supported my decision, but then gently reminded me that I was now the ‘world’s expert” on independent gametophytes and I ‘‘owed it to science’ to stay one more year to publish what I had learned. By the time I did that, of course, my thinking had changed and I was forever hooked on fern gametophyte biology. Herb probably expected that outcome, but it wouldn’t have hap- pened without his encouragement and confidence in me. The excitement of field trips with Herb and Florence remain highlights of my graduate years at Michigan, as do warm memories of holidays at the Wagner home, cutting out snowflakes or whatever project Florence had de- signed for their “extended family” of graduate students. This nurturing of an Ozark farm boy a long way from home made a difference. It was a personal gift. Yet I know it was only one of many such personal gifts, bestowed on many others as well, by Herb and Florence. For those gifts we all say thanks!—Don Farrar, Department of Botany, Iowa State University, Ames, Iowa 50011. American Fern Journal 92(2):50-64 (2002) Bibliography of Warren Herbert Wagner, Jr. Davin B. LELLINGER Department of Botany, Smithsonian Institution, Washington, DC 20560-0166 Asstract.—This bibliography contains 395 entries, all the scientific publications of Warren H. Exact publication dates were found for most journal articles. However, month or estimated dates were used for other lire eciseem books, which usually are - datable within a = All the entries are contained in a searchable database now maintained by Florence S. Wagne The eral categorizes the ee according to their subject matter and includes a more or less exact publication dates that are the basis for the onan list. Wacner, W. H., Jk. 1941. New localities for Botrychium matricariaefolium in Maryland. Amer. Fern J. 31:21—22. ———.. 1941. Butterfly hunting: Why and How? The Naturalist 1:6 1941. [Mimeographed publi- cation of the Naturalists Field Club of the University of Pennsylvania] . 1941. District of Columbia Butterfly Notes (Lepidoptera: weit tale Entomol. News 52: 196-200, 245-249. The parts were published 18 Jul and 8 Oct . 1942. Bipinnate Christmas ferns. Amer. Fern J. 32:27—29. 9—70. ia) 1943. New locality for a rare Hairstreak (Lepidoptera: ane Entomol. News 54:11. 1943. scat: sae by remote control. Amer. Fern J. 3 7 944, rms new to Trinidad. Amer. Fern J. 34 oe 1944, pare occurrence of the apparent hybrid Cystopteris. Amer. Fern J. 34:125—127. 1945. Fern hunt in Puerto Rico. Amer. Fern J. 3 1945. Ferns on pacific island coconut trees. Amer. Forn J. 35:74—76. 1946. Fern field notes in the Washington—Baltimore area. Castanea 11:59-60. . 1946. Notes on the protection of rare ferns in the Wochisister Malti asces area. Bull. on dis Washington-Baltimore Area Flora 10:5-6. [Mimeographed] - 1946. Botrychium multifidum in Virginia. Amer. Fern J. 36:117-1 : 1946. Pteridophyta. Pp. 3-8 in Hermann, F. J. A checklist of olnnts of the Washington— Baltimore area, ed. 2. [Mimeographed . 1947. Tree-climbing Gleichenias. Amer. Fern J. 37:90-95 & D. F. GRETHER. 1948. Pteridophytes of Guam. Occas. Pap. Bernice Pauahi Bishop Mus. 19(2):25-99, ———.. 1948. A new fern from Rota, Mariana Islands. Pacific Sci. 2:214—-21 RETHER. 1948. The Pteridophytes of the Admiralty Islands. Unity, Calif. Publ. Bot. 23(2): 17-110, t. 5-25. & D. F. GreTHer. 1948 [1949]. The butterflies of the Admiralty Islands. Proc. U. S. Natl. Mus. 98:163-186. ——.. 1949. A reinterpretation of Schizostege lidgatei (Baker) Hillebrand. Bull. Torrey Bot. Club 76:444—461. . 1950. The Hart’s-tongue Fern. 18th Annual Calif, Spring Garden Show, April 21-28, pp. i, = | . | | | | Re | 1950. The habitat of Diellia. Amer. Fern J. 40:21— - 1950. Ferns naturalized in Hawaii. Occas. Pap. ae Pauahi Bishop Mus. 20(8):95— 121. - 1951. A new species of Diellia from Oahu. Amer. Fern J. 41:9-13. LELLINGER: BIBLIOGRAPHY OF WARREN HERBERT WAGNER, JR. 51 . 1951; Sie pe a analysis of evolution in Pteridophyta. [Review of: Manton, I. 1950. Problems of Cytology and sig in the Pteridophyta. Cambridge Univ. Press, London sir sire York.] Ss. oe 5:177— 951. Review of: Manton, I. pi Problems of Cytology and gigerton ie in the Pterido- ha Cambridge Univ. Press, London and New York. Amer. Fern J. 4 93. . 1952. The flowerless plants. [Review of: Billington, C. 1952. Fems of Michigan. Cran- — Institate oe Science, Bloomfield Hills, MI.] Sci. Monthly 75(5):3 rm Genus Diellia: Its Structure, Affinities and Ao Univ. Calif. Publ. ree a) leah t. 1-21. . 1952. Review of: McVaugh, R. & J. H. degen 1951. Ferns of Georgia. Univ. of Georgia mae pe Michigan pee 58(21):366— ———. 1952. Review of: Billington, C. 1952. caine of Michigan. Cranbrook Inst. of Science, Bloomfield Hills, MI. Cranbrook Inst. Sci. News Letter 22(1): ee . 1952. Juvenile leaves of two Polypodies. Amer. Fern J. 42:8 . 1952. Types of foliar dichotomy in living ferns. Amer. J. i pe 578-59 92. . 1953. The genus Diellia and the value of characters in determining fern affinities. Amer. J. Bot. 40:34—40. 2 1953; An Asplenium prototype ne be genus Diellia. one — Bot. Club 80:76-94. . 1953. Report of the Secretary for 2. A . 1953. A cytological study of the pam Spleonw i re mer. Fern J. 43:109-114 1953. Review of: Hubbard, D. H. 1952. Ferns of Hawaii National Park. Hawaii Nat. Hist. ——— & D. J. HaGenan. 1954. A natural hybrid sy et apaerias lonchitis and P. acrostichoides en the Bruce Peninsula. Rhodora 56:1-6, t. 4. A Bolivian Elaphoglossum of unique “leaf structure. Bull. Torrey Bot. Club 81: 1954. Report of the Secretary for 1953. Amer. Fern J. 44:21-24 1954. Reticulate evolution in the Appalachian Asphiiiana, Evolution 8:103-118 [Reprinted as pp. 136-151 in Orndoff, R. 1967. Papers on Plant Systematics. Little, — Boston. ] oes The evidence used recent classifications of the ferns. Rapports Comm. aux Sect. : , 6, Ville Congr. Int. B WILSON, - ci & W. H. WAGNER he pris a of: Lawalrée, A. 1950. Flore cnn de Belgi- que: Pt spy Jardin Botanique de |’Etat, Bruxelles. Amer. Fern J. 4 Wacner, W. H. Jr. 1955. Cytotaxonomic observations on North American “on phen 57:219- 240. . 1955. Should - American Hart’s-tongue be interpreted as a distinct species? Amer. Fern J. 45:127-1 . 1955. Austin ie Clark. re News 9(4,5):151-157. . 1955. Review of: Manton, I. . A. Sledge. 1954. Observations on the cytology and acne of the sean sh For Ceylon. Phil. Trans. Roy. Soc. London, Biol. Sci. 238:127-185. Amer. Fern J. 4 . 1956. A reoviany rege mii en a C. C. Hall. Madrofio 13:195—205 D. J. Hac . 1956. A diploid variety in the Cystopteris fragilis sienpies Rhodora 58:79-87, t. ee. . 1956. Asplenium ee x ot gla a new triploid hybrid produced under arti- fical conditions. Amer. Fern J. 4 Voss, E. G. & W. H. WAGNER . 1956. in on Pieris virginiensis and Erora laeta—two butterflies hitherto enikciovted from Michigan. Lepidopterists’ News 10:1 & L. P. Lorp. 1956. The morphological and cytological distinctness of Botrychium minga- swe and B. lun in Michigan. Bull. Torrey Bot. Club —280. D. J. emia oe [1957]. a a on some tralia: -producing populations of the Cystopteris fragilis complex. Amer. Fern J. 46:137-146. 1957. Heteroblastic leaf moepholgy in juvenile plants of Dicranopteris linearis (Gleiche- tincead). Phytomorphology 7 52 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) . 1957. Botanical Phase. Pp. 1-14 in J. M. Sheldon & E. W. Hewson. Atmospheric Pollu- tion by Aeroallergens, Progress Report No. 1. Eng. Res. Inst. 2421-1-P, Ann Arbor, MI J. DaruING, JR. 1957. Synthetic and wild Asplenium gravesii. Brittonia 9:57-63. STEINER, E., A. S. SUssMAN & W. H. WacnerR Jr. 1957. Botany Laboratory Manual. Dryden, New York Wacner, W. H. Jr. & R. S. WHiTMIRE. 1957. Spontaneous production of a morphologically distinct, fertile oo by a sterile diploid of Asplenium ebenoides. Bull. Torrey Bot. Club seg i is GILBERT, 1957. An unusual new Cheilanthoid fern from California. Amer. J. Bot. : 743. Constance, L., H. Lewis, R. C. Rouuins, R. F. THORNE & W. H. WAGNER JR. 1958. Suggested outline for at aprons botany. Pl. Sci. Bull. 4(1):1-3 esis ni H. Jr. 1958. Notes on the distribution of a kentuckiense. Amer. Fern J. 9-43. eee Botanical Phase. Pp. 1-13 in J. M. ci & E. W. Hewson. Messiah sone Pollu- tion rc neering a, Progress Report No. 2. Eng. Res. Inst. 2421—-2-P, Ann Arbor, & EA 8. Perennial ragweeds fAnabocetia) in Michigan, with the lem of a new, vito Si ie Rhodora 60:177—204 1958. Review of: Bold, H. C. 1957. Marntolosy of Plants. Harper, New York. Science 128:1079. Dincte, A. N., J. C. Gin, W. H. Wacner Jr. & E. W. Hewson. 1959. The emission, dispersion and deposition of ragweed pollen. Adv. Geophys. 6:367—387 Wacner, W. H. Jr. 1959. Botanical research on atmospheric pollution. Proc. 13th Ann. Meeting Northeastern Weed Control Conf., Rutgers Univ.:262-268. [Mimeographed] K. wood and its relationship to a peculiar plant from West Virginia. Amer. Fern J. 48:146— 159. - 1958 [1959]. The hybrid ragweed, Ambrosia artemisiifolia x trifida. Rhodora 60:309- - 1959. Problems in the classification of ferns. Proc. IXth Intl. Bot. Congr. 2(Abstrs.): 418-419. CanTLON, J. E., W. H. WAGNER JR. & W. is Riripon 1959. A range extension in Arctic Alaska for cares lunaria, Amer. Fern J. 49:25-29. Wacner, W. H. Jr. 1959. Hybrid swarms of mittee [Review of: Walker, T. ss 1958. Hybridization in some — of Pteris L. Evolution 12:82—92.] Amer. Fern J. 49:4 , ed. 9. An annotated bibliography of ragweed (Ambrosia). a Ailacay Appl. Immu- nol. . ae - 1959. Botanical phase. Pp. 1-22 in J. M. Sheldon & E. W. Hewson. Atmospheric Pollu- tion ky Aeroallergens, Progress Report No. 3. Univ. Mich. Res. Inst. 2421-3-P, Ann MI. . SCHWEMMIN & W. H. Wacner JR. 1959. Pollen release in the common ragweed todanaie cape tagioersd Bot. Gaz. 120:235—243. W. . 1959 can Grapeferns resembling Botrychium ternatum: A preliminary . 1960. Evergreen Grapeferns and the pres of infraspecific categories as used in North American Pteridophytes. Amer. Fern J. 50:32—45. . 1960. The proliferations of Asplenium pipet Castanea 25:74-79. . 1960. Ragweeds. Cranbrook Inst. Sci. po Letter 30(1):2-8. - 1960. Botanical phase. Pp. 1-25 in J. M. Sheldon & E. W. Hewson. a ao tion by sy gi Progress Report No. es Univ. Mich. Off. Res. Admin. 034 Ann . 196 ae and pigmentation in aa subg. Sceptridium in the northeastern United ‘Staten, Bull. Torrey Bot. Club 87:3 961. Filicineae, Frond. Pp. 396-398, pooh in P. Gray, ed. The Encyclopedia of the nies Sciences. Reinhold, New York. ee 9 LELLINGER: BIBLIOGRAPHY OF WARREN HERBERT WAGNER, JR. 53 . 1961. Some new data on the vernation differences of Botrychium dissetum and B. terna- tum. Amer. Fern J. 51:31-— . 1961. Problems in the elaneification of ferns. Pp. 841-844 in Recent Advances in Botany, ———. 1961. On the relative development of the fertile segments in Botrychium dissectum and B. ear Amer. Fern J. 51:75-81. 1961. The Goldback Ferns of California. Review of: Alt, K. S. & V. Grant. 1960. Cytotaxo- nomic observations on the Goldback Fern. Brittonia 12:153-170. Amer. Fern J. 51:105-106. . 1961. Spore studies and speciation in the ferns on Hawaii. 10th Pac. Sci. Congr. Abstrs. Symp. Pap. 135-136 . 1961. Roots and the naa differences between Botrychium oneidense and B. dissec- tum. Rhodora 63:164-1 & K. E. Boypston. io, en new hybrid showing homology between Asplenium ebenoides and A. pinnatifidum. Brittonia 13:286—289. 961. Nomenclature and typification of two Botrychiums of the southeastern United States, Taxon 10:165-169. 1 Review of: Wherry, E. T. 1961. The Fern Guide. Northeastern and Midland Eiesiliie States and Adjacent Canada. Doubleday, Garden City. Amer. Mid]. Naturalist 67: & D. J. HaGENAH. 1962. Dryopteris in the Huron Mountain Club area of Michigan. Brit- a 14:90—100. 962. Cytological observations on Adiantum xX tracyi C. C. Hall. Madrofio 16:158—161. ae, grein The endemic Botrychiums of the southeastern United States. ASB Bull. 9:40. Morzentl, V. M. & W. H. WacNER Jr. 1962. Southeastern American ‘‘Blackstem spleenworts” of 0-41 Wacner, W. H. Jr. 1962. Plant co —— and leaf production in iene multifidum “‘ssp. icum’ and ‘forma dentatum.’’ Amer. Fern J. 52:1-18. . 1962. The synthesis and expression of nou pennies data. Pp. 273, 276, 277 in L. Benson. 1962. Plant Taxonomy: Methods and Principles. Ronald Press, New York. . 1962. A graphic method for orn inka: based upon group correlations of indexes of divergence. Pp. 415 n L. Benson. Plant Taxonomy, Methods and Princi- ples. Ronald Press, New York. . 1962. Irregular morphological development in hybrid ferns. Phytomorphology 12:87—100. . 1962. Ferns and their allies. [30-minute film] AIBS Secondary School Biological Sciences Film Series, McGraw-Hill 962. oe crosses and spore characteristics in the genus Psilotum (Psilotaceae). ages J.B 962. soiree boschianum in Arkansas. Amer. Fern J. 52:85—86. Geanen, ri I., W. W. Payne & W. H. WacNner JR. 1962. Botanical phase. Pp. 1-16 in J. M. Sheldon & E. W. Hewson. Atmospheric Pollution by Aeroallergens, Progress Report No. 5. Univ sch: Off. Res. Admin. 04202, 04771, 04772, 04773-1-P, Ann Arbor, MI. Wacner, W. H. Jr. & D. E. RAWLINGS. 1962. A sampling of pins agen Subg. Sceptridium in the vicinity of Leonardtown, St. Mary’s Co., Md. Castanea 27:132-1 toa ips and taxonomic categories in lower mated plants. Pp. 63-71 in V. H. Harwue . Léve, eds. Symposium on Biosystematics. Intl. Assn. for Plant Taxono- my, Utrecht, ie ds. . 1963. A biosystematic survey of United States ferns—Preliminary abstract. Amer. Fern J. 53:1—16. . 1963. Gros and herbarium studies, University of Washington, June 23-28, 1963. [Mimeo- — 12p 5 ee raat 1963. gee prothallial clones (Trichomanes?) locally abundant on Illinois Reon Amer. J. B 0:623. 1963. Pteridology in oie Newslett. Hawaiian Bot. Soc. 2(8):117-123. [Mimeo- graphed] 54 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) J. SHARP. 1963. A cnet reduced vascular plant in the United States. Science 142:1483-1484, cover picture . 1963 [1964]. Pteridaphytes of the concn mye area, Giles County, Virginia, including notes from Whitetop Mountain. Castanea 28:1 964. Chemistry and taxonomy. [Review = ie ee ed. 1963. Chemical Plant Taxon- ony Academic Press, New York.] Science 143:1428. . 1964. vig vce Filicinae. Taxon 13:56-64. & K. L. Cuen. 1964. [23 new chromosome records] in Love, A. & O. T. Solbrig. IOPB oe ae eports. I. Taxon 13:99-102 . Fern Geecihen ten in terms of divergence patterns. Abstrs. Xth Intl. Bot. Congr. a sie Assessment of species relationships in the Filicineae. Abstrs. Xth Intl. Bot. Congr. 147-14 Evans, A. “ & W. H. Wacner Jr. 1964. me cae i x intermedia—a natural Woodfern cross of noteworthy morphology. Rhodor. 266. Wacner, W. H. Jr. 1964. vig ps Pteridophyta an cai 1-15.] Pp. 1, 2, 39-48 i . Rad- ford, H. E. Ahles & C. R. Bell. Guide to the Vascular Flora of the Carolinas. he, a North Carolina, Aran Hill. Scora, R. W. & W. H. Wacner Jr. 1964. A oo chromatographic study of eastern Ameri- can pce he Hn Fern J. 54:10 Wacner, W. oa JR. 1964. The neha “een of living ferns. Mem. Torrey Bot. Club 21(5):86—95. . 1964 [1965]. ees creer Copeland (1873-1964) and his contributions to pteridology. Amer. senha 177- . 1965. Edwin Bin any Copeland 1873-1964. Taxon 14:33—41. . 1965. ietitadies maps of Familes 1-15, Pteridophyta] Pp. 1-6 in A. E. Radford, H. E. Abiles & C. R. Bell. Atlas of the Vascular Flora of the Carolinas. Univ. No. Car. Tech. Bull. 165. & K. L. CHEN. 1965. eee o ie as and sporangia as a tool in the detection of Dryo- eee hybrids. Amer. Fern J. 5 ——— & F. S. Wacner. 1965. Rochester area ee Ferns (Dryopteris celsa) and their hybrids. Proc. yal Sa . Sei. 11(2):59— STEINER, E., A. S. SUSSMAN & W. H. ia JR. 1965. Botany Laboratory Manual, rev. ed. Holt, Rinehart ; Winston, New York. Wacner, W. H. Jr. 1965. Pellaea oo in North Carolina and the question of its origin. J. Elisha Mitchell Sci. Soc. 81:9 . 1965. Bibliography. pete BioScience 15:809-810. . 1965. Fern paraphyses: acipusaaig on recent papers. Taxon 1 » D. R. Farrar & K. L. CHEN. 1 ye A new sexual form of Pellaea sails var. glabella from Missouri. Amer. Fern J. 55:171-178. & J.B. Sm IN. 1966. ae ie Sections and Societies. Botanical Sciences (G). Science 151:865—866. MickeL, J. T., W. H. Wacner JR. & K. L. CHEN. 1966. Chromosome observations on the ferns of Mexico. Caryologia 19:95-102. Wacner, W. H. Jr. 1966. Two new species of ferns from the United States. Amer. Fern 1: 56: sa 1966. Report of the Librarian and Curator for 1965. Amer. Fern J. 56:45-46, - 1966. New data on North American Oak Ferns, oo eg 68:121-138. - 1966. Modern Research on Evolution in the Ferns. Pp. 64-184 i . A. Jensen & L. G. Kavaljian, eds. Plant Biology Today. Advances and os ee ed. : Sn Belmont, . 1966. gman in the Ferns. Bull. Fairchild Trop. Gard. 21(3):4—8,17-22, cover picture. & F Wacner. 1966. Pteridophytes of sek Mountain Lake Area, Giles Co., Virginia: Bio- repent beste. 3 1964-65. Castanea 31:1 966. The e new University of Michigan psoas Gardens. Taxon 15:285—286. LELLINGER: BIBLIOGRAPHY OF WARREN HERBERT WAGNER, JR. 55 . 1966. Illustrations of transient fern forms. Amer. Fern J. 56:101—107. . 1967. Reports of Sections and Societies. a Sciences (G). Science 155:880. Eva , J. T. MickeL, D. B. LELLINGER & A. M. 967. Pteridophytes of Costa Rica: Prelimi- nary chockliet ad species known to be or ie occurring in Costa Rica. [Mimeographed, 54 as ] R. C. Woopsipe. 1967. Ferns for an indoor limestone boulder habitat. Amer. Hort. Mag. oa 219-223 . E967. Rexiow of: R. L. Taylor & R. A. Ludwig, eds. 1966. The evolution of Canada’s flora. Univ. of Toronto, Toronto. Canad. Field-Naturalist 81:206—207. . 1968. Hybridization, taxonomy, and evolution. Pp. 113-138 in V. H. Heywood, ed. Mod- ern Methods in Plant Taxonomy. Academic Press, London . L. McWituiaMs. 1968. The University of Michigan Botanical Gardens special collec- tions. Arb. & Bot. Gard. Bull. 2:43-46. . 1968. recreate stamens and carpels of a willow from northern Michigan. Michigan Bot. 7:113—120. 196 ge taxonomy and modern systematics. BioScience 18:96—100. Farrar, D. : & W. H. WAGNER Jr. 1968. The gametophyte of voter holopterum Kunze. Bot. Gaz. 129:210-219. Wacner, W. H. Jr. 1968. Review of: R. H. Mohlenbrock. 1967. The Illustrated Flora of Illinois: Ferns. So. Ill. Univ. Press, Carbondale. Quart. Rev. Biol. 43(4):461 ———. 1968. The role of a botanical garden in a modern marae Scietiee 162:1165- . 1969. The construction of a classification. Pp. 67-90 in Systematic Biology. or 1692 cit the National Academy of Science, A erscnts DC. (Translated by E. R. de la Sota and published as La Construccién de una Clasificacién. Quid de la Ciencia, la Tecnologia y la Educacion 2(18):440-448. 1983, haa 532-538. 1984.] oe, es Richard Eric Ho sean sagen kgeray pteridologist. Amer. Fern J. 59:1-3. ——.,F.S. “Wa GNER & D. J. Ha . 1969. The Log Fern (Dryoptreris celsa) and its hybrids in sen ig preliminary a Michigan Bot. 8:137—145. oe . Bio —— important is it to basic evolutionary systematics? Abstrs. xh a Bot. Congr. 230. F. S. WacneR. 1969. A new natural hybrid in the Appalachian Asplenium complex and its taxonomic significance. Brittonia 21:178-1 . 1969. The role and taxonomic treatment of ntti. BioScience 19:785-—789. . 1970. . The Barnes pyres — Populus X barnesii, hybr. nov—A nomenclatural case in eal Michigan Bot. 9:53— ——. 19 Sg-te sane wad evo ia ionary noise. Taxon 19:146—151. . 1970. Natural oe of floating stems of Scouring-rush, Equise- tum hyemale. ‘Miche Bot. 9:166-17 , D. R. Farrar & B. W. McALpPIN. oa epee ae _ eee Biological Station area, ees ee J. Elisha Mitchell Sci. & L. D. Gomez P. . Field research on ferns in co ae Ophioglossum, Dictyoxi- phium, Geetha rae 10-12, 1971. iaiwnsaeiepties 4pp.] 1971. Evolution of naires in ia to the a Virginia Polytech. Inst. State Univ. Res. Div. Mongr. 2:147—19 1971. The Sonthetaions Adder’s- —— Ca epee vulgatum var. pycnostichum, ig for the first time in Michigan. Michigan Bot. 971. Ferns in biology: Some final comments. nose 21:311, 323-324. nt Report of the President for 1970. Amer. Fern Soc. News & Views 2: — . 1971. Lindsaea (Schizoloma) ae Swartz in Hawaii. Amer. Fern J. 61 58. Wuirtier, D. P. & W. H. WaGNER JR. 1971. The variation in spore size and pre he in Dryo- pteris taxa. Amer. Fern J. 61: ofeagee 1972. Botanical research at botanical gardens. Pp. 28-33 in P. F. Rice, ed. agen, of the Symposium—A National Psion Garden System phi Canada. Techn. Bull. N Royal Botanical Gardens, Hamilton, Ontario, Canada. 56 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) snes 2 iy by & W. H. Wacner Jr. 1972. A native Woodfern garden. Amer. Horticulturist 51(1 WAGNER, ae 7 Jr. 1972. Estimation of the evolutionary ane Bibliographic references —— to the en gga Method. [Mimeographed, Pp. Cain, S. A. & W. H. WaGNER ash Dale J. Hagenah (1908-1971): An outstanding Michigan botanist. Michigan Bot. 1 ae Wacner, W. H. Jr. 1972. io pe the President for 1971. Amer. Fern Soc. News & Views 6:1—4. . 1972. Dale J. Hagenah. Amer. Fern J. 62:2 , F. S. Wacner & L. D. cia P; 1972. The Cantal American fern genus Pleuroderris: its — yr significance r. J. Bot. 59:677. eo a nonst a little-known fern epiphyte with dimorphic stems. Amer. ae 62:3 oar 1° . 1972 eal Review of: R. M. Lloyd. 1971. perenne of the Onocleoid ferns. Univ. Calif. Publ. Bot. 61:1-93. Bull. Torrey Bot. Club 99: . 1972 [1973]. Disjunctions in homosporous a plants. Ann. Missouri Bot. Gard. 59:203—217. & D. R. Farrar. 1973. The oo fern genus Hyalotricha and its family relationships. lage J. Bot. 60(4, suppl.):3 973. An orchid asc for monarch butterflies (Danaidae). J. Lepidopterists’ Soc. 27: apg . 1973. Reticulation of Holly shea ee in the western United States and ad- janes er ia mer. Fern J. 63 973. Some future Pee of pes systematics and phylogeny. Pp. 245-256 in A. C. Jermy, J. * Crabbe & B. A. Thomas, eds. 1973. The Phylogeny and Classification of the Ferns. Bot. J. Linn. Soc. 7 Suppl. 1 Tuomas, R. D., W. H. Wacner Jr. & M. R. Masta 1973. Log Fern (Dryopteris celsa) and related species in Louisiana. Castanea 38:269—27 Wacner, W. H. Jr., F. S. WAGNrR, J. A. ata & J. F, MaTTHEws. 1973 [1974]. Asplenium monta- num X platyneuron, a new primary member of the A i te < ltiaiail complex from useing Mountain, N.C. J. Elisha Mitchell Sci. Soc. 89(3):218-223. ——. 1974. es hackberry (Ulmaceae: Celtis tenuifolia) in the a Lakes region. Michigan tg 13:73-9 MEsLeR, M. R. & 4 H. Wacner Jr. 1974. The venation of Ophioglossum dendroneuron and its systematic significance. Amer. J. Bot. ey Suppl.):37 Wacner, W. H. Jr., J. T. Micke. & L. D. z P: 1974. a dubium and its bearing on the penis of generic relationships ey Cyrtomium and Phanerophlebia. Amer. J. Bot. 61(5, Suppl.):39 . 1974. The classification of leaf types of land plants. Amer. J. Bot. 61(5, Suppl.):67. - 1974. Fern. Pp. 237-248 in Encyclopaedia Brittanica, ed. 15, vol. 7. H. H. Benton, Chicas, HL. Hit, R. H. & W. H. hiiggen ag 1974. Seasonality and spore type of the pteridophytes of Michi- gan. Michigan Bot. 13 Wacner, ae H. Jr. 1974. aa years of botany 1947-1972: Introduction. Ann. Missouri Bot. Gard. 61:1-— pe 25 years of botany 1947-1972: Pteridology. Ann. Missouri Bot. Gard. 61:86—111. - 1974. Structure of spores in relation to fern phylogeny. Ann. Missouri Bot. Gard. 61:332-—353. Dancik, B. P., B. V. BARNES & W. H. Wacner Jr. 1974. Aberrant pistillate catkins of Betula alleg- haniensis. Michigan Bot. 13:177-179 Wacner, W. H. Jr. 1975. Review of: Hirabayashi, = sexi Cytogeographic Studies on Dryopteris of Japan. Harashobo, Tokyo. Amer. Fern J. 6 1975. Notes on the floral biology of ci sai ane negundo). Michigan Bot. 14:73-82. 1973 | [1975]. Plant species and ecosystems: a botanist’s viewpoint. Trans. Missouri Acad. 7/8:40. - 1975. The spoken “x” in hybrid binomials. Taxon 24:296. LELLINGER: BIBLIOGRAPHY OF WARREN HERBERT WAGNER, JR. 57 & F. S. WacneR. 1975. A hybrid polypody from the New World tropics. Fern Gaz. 11:125— 135. Wiptn, C.-J., D. M. Brirron, W. H. WaGNER JR. & F. S. WAGNER. 1975. Chemotaxonomic studies on hybrids of neh sea: in eastern North America. Canad. J. Bot. 53:1554—1567. WAGNER, ke H. Jr. 1975. A bunchberry ‘‘Last Rose of Summer.”’ gen Bot. 14:201-—202. 5. Sex he the angiosperms—another proposition. Sida 6 ; rp Review of: G. L. Stebbins. 1974. Flow st Plants: ae Above the Species Level. Belknap Press, Harvard Univ., Cambridge, MA. Amer. Sci. 702-703 oi J. D., W. H. WaGNeER Jr., D. R. FARRAR & S. W. LEONARD. 1975. Asano sea in the nga Biological Station area, southern Appalachians. reve 40 Wacner, W. H. Jr. 1976. How to find the rare grapeferns and moonworts. thee e oaks 3(2):2. & F. S. WacNER. 1976. The role of foliar "ial in the systematics of ferns. Bot. Soc. Amer. Abstr., Tulane Univ., New Orleans. L. D. Gomez & W. H. WAGNER JR. 1976. An uk triploid vie ae polypody from the vicinity of Cartago, Costa Rica. Bot. Soc. Amer. Abstrs., Tulane Univ., New Orleans. 41-42 Wacner, W. H. Jr. & F. S. WacNneR. 1976. Asplenium ee een Sim ‘rom gues Gorge, Greene County, Ohio—a second North American record. Ohio J. Sci. 76:9 & D. J. SCHOEN. ape yi Oak (Quercus imbricaria) and its hybride in Michigan. Michigan Bot. 15:14 — & W. C. TAyLor. api yaa x leedsii and its westernmost station. Sida 6:224—234. A. H. SHowa ter. 1976. Ecological notes on Celastrina ebenina (Lycaenidae). J. Lepi- ie ong Soc. 30:310-312. . Systematic implications of the Psilotaceae. Brittonia 29:54—-63. NER. 1977. A diminutive grapefern (Botrychium) of the maple—basswood for- ests in 1 the Wpper Great Lakes region. Bot. Soc. Amer. Misc. Ser. Publ. 1 & D. . 1976 [1977]. a Central American fern genus Hyaloticha aa its family Seeneikieeks Syst. Bot. 1:348-36 Voss, J. H. BEAMAN, E. A. es F. W. Case, J. A. CHURCHILL & P. W. THOMPSON. 1977. Bnduneoced, threatened, and rare vascular plants in Michigan. Michigan Bot. 16:99-110. ——., F. S. Wacner & L. ~ — P. 1977. An enigmatic polypody fern from Cartago, Costa Rica. igegs 12/13:8 ——. 1977. A distinctive ae form of be Marbled White Butterfly, Euchloe ae (Lepi- dope bern in the Great Lakes area. Great Lakes Entomologist 10:107-1 . S. Wacner. 1977. Fertile- amet leaf dimorphy in ferns. Gard. Bull. ai Settlem. sosi-267 , F. S. Wacner, C. N. MILLER Jr. & D. H. WacNeR. 1978. New observations on the Royal Hern iii Osmunda x ruggii. Rhodora 80:92-106. — & T. L. MELLICHAMP. 1978. Foo — habitat, and range of Celastrina ebenina (Lycaeni- — - Lepidopterst Soc. 32:20— eer Great om anita Pieris virginiensis nae Pieridae) in com- prion siti outhern counterpart. Great Lakes Entomologist 11 ——. 1978. Venlo idioblasts in Pteris and their systematic Sag Acta Phytotax. ane 29:3 . 1978. ea of eo veins in modern ferns. Bot. Soc. Amer. Misc. Ser. 156:30 — & K. E. Boypston. 1978. warf coastal variety of maidenhair fern, Adiantum pedatum. Canad. J. Bot. oe . 1978. A probable natural ap tic of ao eurymedon and P. rutulus (Papilionidae) from {dale J. ass eee Soc. 32 . WaGNER & L. D. GO es 978. The singular origin of a Central American fern, eae bee mic aes ee aes 10:254-264. 1978 [1979]. Hyalotrichopteris, a new generic name for a Central American polypodioid fern. Taxon 27:548. M. BEITEL. 1979. oreo variation in clones of Running Pine, Lycopodium flabelli- forme. Michigan Bot. 18:19— AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) - 1979. Reticulate veins in the systematics of modern ferns. Taxon 28:87—95. , L. D. Gomez P. & F. S. Wacner. 1979. Coenosori and foliar nen in New World Pleo- peltis and Marginariopsis. Bot. Soc. Amer. Misc. Ser. Publ. 157:48—49. & F. . 1979. [Accounts of i silenare or pre ae or extirpated Senet phytes.] $0: 35-41, 108-110 in D. W. L , LMAN, sa how Ferns (Dryopteris celsa) and their relatives in the Dismal Swany Pp. 127-139 in P. W. Kirk Jr., ed. The Great Dismal Swamp. Univ. Press of Virginia, Oostinc, D. P., D. J. HARVEY & W. H. WAGNER JR. 1979. Euristrymon ontario (Lycaenidae): first re- port in Michigan. J. Lepidopterists’ Soc. 33:151-152. Wacner, W. H. Jr. 1979. New herbarium building in Costa Rica. Taxon 28:358. » F. S. Wacner & J. M. BerreL. 1979. An unusual occurrence of Asplenium heterochroum. Canlele a 174-177. SmiTH, A. R., W. H. WaGNeR Jr. & T. DUNCAN. 1980. Noteworthy Collections. sae ape lusita- nicum L. ee i a Apr, Clausen (Ophioglossaceae). Madrofi Wacner, W. H. & F. S. Wacner. 1980. Polyploidy in Pteridophytes. Pp. mean in an H. aes ed. ae Popo. Biolog Relevance. Plenum Press, New York & London. 1980. Review of: R. L 8. A acne ‘eae of Equisetum subg. Equi- peteati: Nova Hedwigia 30: eae Michigan Bot. 1 - 1980. A probable new hybrid central pera Michigan Bot. 1 , T. F. DANIEL & M. “ HANSEN. vel o Michigan Bot. ie 37-4 Wa ter, K. S., W. H. meet jr. & F — cesta matricariifolium x simplex, from ybridizing Verbascum population in Michigan. - S. Wacner. 1980. Ecological, biosystematic, and nomencla- tural oo on Scott’s Splocawact, x Asplenosorus ebenoides. Bot. Soc. Amer. Misc. Ser. Publ. 158:123. Wacner, W. a jn & M. K. Hansen. 1980. Size reduction southward in Michigan’s Mustard White Butterfly, Pieris napi (Lepidoptera: Pieridae). Great Lakes Entomologist 13:77-88 ——. 1980 . Dyer, ed. 1979. The Experimental Biology of Pan. Academic Press, — York. Econ. Bot. 34:305-306. . MAYFIELD. 1980. Foodplants and cocoon construction i (Lepidoptera: Saturniidae) i in southern Michi ussion pp. 222, 224, 227, 229. Reprinted as pp. . F. eds. 1985. Chadinte Theory and Methodology. Van Nostrand eSnips New York.] Barnes, B. V. & W. H. WAGNER Jr. 1981. Michigan Trees. A ned to vad of Nein and the Great Ladue Region. Univ. of Michigan THOMPSON, J. W. & W. WAGNER, w. H. a a F. Dani a "vs EITEL. 1980 [1981]. Studies on Populus heterophylla in southern Michigan. Michigan Bot. 19:269— Duncan, T., R. B. PHitups & W. H. AGNER, JR. ti [1981]. A comparison es —e diagrams derived by various — and cladistic methods. Syst. Bot. 5:264— Wacner, W. H. GNER. 1981. New s chium (phiogonsicon) from —) Ameri D.M SON. .. W. H. Wacner Jr. & K. S. Wa ter. 1981. Unu usual frond developmen in the Sensi- tive Fern Onoclea sensibilis L.. Amer. Midl. Naturalist 105:396—400 LELLINGER: BIBLIOGRAPHY OF WARREN HERBERT WAGNER, JR. 59 Wacner, W. H. Jr., M. K. HANSEN & M. R. MayYriELD. 1981. True and false foodplants of Callosamia paced (Lepidoptera: Saturniidae) in southern Michigan. Great Lakes Entomologist 14:159-16 fee Fems in the Hawaiian Islands. i wniengce Forum 8:43—4 ——, F. S. Wacner, S. W. LEONARD & M. R. 981. A ts ai of Ophioglossum Bsn “ 'p. St. John. Castanea 46: ne on ———. 1982. Ferns, Clubmosses, Spikemosses, Quillworts, and Horsetail. Pp. phe gt in -S.']. yeaa pe — North American Wildlife. Readers’ Digest Assn., Pleasantville, N —-. F. AGNER. 1982. The taxonomy of Dryopteris x poyseri Wherry. pt Bot. 21; a : Funk, V. A. & W. H. WaGNER Jr. 1982. A bibliography of botanical cladistics: I. 1981. Brittonia sages ere Wacner, W. H. Jr. . S. PEIGLER. 1981 [1982]. Two notable range extensions for Callosamia securifera lovers in Georgia and ane Carolina. J. Lepidopterists’ Soc. 35:247 F. Wacner & C. HaAuFLeR. 1982. A hybrid population hae the ‘‘all- fertile” Paths paradoxum and the hemidimerphic B. hesperium (Ophioglossaceae). Bot. Soc. Wa ter, K. S., W. H. WaGNER JR. & F. S. WAGNER. 1982. erie biosystematic, and nomencla- tural notes on Scott’s Spleenwort, 0.934) for both Rogers’ (1972) and Nei’s (1978) genetic coefficient (Table 4). In population JD, only one genotype was found at each of the locus-pairs. In population CH, Lap and Mdh-3 each had two genotypes, combining to form three multilocus genotypes. In population FS, Lap and Pgm-3 each contained three genotypes and the variable locus 6Pgd-2 contained two geno- types, combining to form six multilocus genotypes (Table 5). In combination, the three populations form 10 multilocus genotypes. Of 30 plants tested, only one displayed recombinational heterozygosity at one locus (Lap 11/23). In P. aureum, 22 putative loci were scored among the 13 enzyme systems. There was no variability within or among the three populations for 21 locus pairs. Eleven (Ald, Fbp-1, Fbp-2, Idh-1, Idh-3, Mdh-1, Mdh-3, 6Pgd-2, Pgi-1, Skdh-1, and Skdh-2) were fixed at the same allele, and ten (Aco-1, Aco-2, Hk, Lap, Mdh-2, 6Pgd-3, Pgi-2, Pgm-1, Pgm-2, and Tpi-2) had fixed heterozygos- TABLE 4. Matrix of Roger’s genetic similarity (above diagonal) and Nei’s unbiased genetic identity (below diagonal) among three populations of Campyloneurum phyllitidis in Florida. Population’ JD CH FS JD ae 0.938 0.934 CCH 0.946 ae 0.951 FS 0.956 0.981 a a 1jD= — Dickson State Park, CH=Castellow Hammock, FS= Fakahatchee Strand State Preserv v2 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) TasLe 5. Description of multilocus genotypes of two populations of Campyloneurum phyllitidis in Florida. Locus CH! FS! Lap-a/b 11/22 11/33 11/33 11/22 41/22 11/33 11/33 11/23 11/22 Mdh-3a/b 29/99 11/11 29/22 Pgm-3a/b 47/11 11/22 ss | 22/22 as a 114/91 6Pgd-2a/b 41/22 22/22 11/22 11/22 22/22 22/22 Observed 10% 10% 80% 30% 10% 20% 10% 10% 20% ‘CH = Castellow Hammock, FS = Fakahatchee Strand State Preserve. ity. Got displayed fixed heterozygosity for the same alleles in TS and JD pop- ulations and in 90% of the samples in FS, but had fixed heterozygosity with a different allele for Got-b in the other 10%. The genetic similarity between population FS and the two identical populations TS and JD was high, 0.998 in Rogers’ (1972) genetic similarity and 1.0 in Nei’s (1978) genetic identity. In populations TS and JD, there was only one genotype for all the locus- pairs. In population FS, only two genotypes appeared, Got-a!'/b"™ (10%) and Got-a''/b** (90%). No recombinational heterozygous individuals were de- tected among the 30 plants tested. DISCUSSION In multi-gametophyte cultures of all species, male and bisexual plants ap- peared only after a significant number of female plants had differentiated. This is consistent with the presence of antheridiogen systems in these spe- cies, as has been previously demonstrated (Chiou & Farrar, 1997a). Anther- idia on bisexual gametophytes grown as isolated plants were probably induced by their own gametophyte’s antheridiogen secretions. This could occur either by absence of or delayed attainment of insensitivity of a gameto- phyte to antheridiogen, or by generation of secondary lobes with reduced physiological connection to the principal antheridiogen-producing apex. A tion from antheridiogen. Plants that remained unisexual males at eight months generally were relatively slow-growing and small. In isolated-plant cultures, a significant difference in sexual expression between gametophytes grown from isolated spores and those grown from gametophytes isolated from multi-gametophyte cultures when they were one pellucidum is that plants in multi-gametophyte cultures were subjected to antheridiogen before individual gametophytes were transferred to the iso- CHIOU ET AL: MATING SYSTEMS 73 lated-gametophyte cultures. It is also possible that transplanting actively growing gametophytes might disrupt their normal developmental pattern. Neither hypothesis explains the slow rate at which bisexuality is attained by spore-isolated gametophytes of these species relative to the other species tested. Genetic load can be estimated by the ability of isolated bisexual plants to produce sporophytes. Isolated gametophytes of Polypodium pellucidum, Microgramma heterophylla, and the B source of Campyloneurum angustifo- lium produced virtually no sporophytes through syngamy, whereas isolated gametopytes of C. phyllitidis, Phlebodium aureum, Phymatosorus scolopen- dira and the A source of C. angustifolium produced abundant sporophytes through intragametophytic selfing. Genetic load was not significantly different between isolated-spore and isolated-gametophyte cultures, except in Campyloneurum phyllitidis where the apparent genetic load of isolated-spore cultures was much greater. The reason for this apparent elevation of genetic load in isolated-spore cultures in C. phyllitidis is not clear, but from the delayed production of bisexual plants evident in isolated spore cultures (Table 1), it is possible that anther- idia in some bisexual plants may still have been functionally immature even at eight months, falsely indicating a high genetic load. No similar delay in production of antheridia by this species was evident in multispore cultures (Chiou & Farrar, 1997a) or in isolated-gametophyte cultures. However, pair- ing of gametophytes from different sporophytes in C. phyllitidis also failed to fully relieve the sporophyte suppression observed in isolated gametophyte cultures. This suggests that some genetic load in this species is perhaps being expressed in gamete development (Klekowski, 1971), or that there was very little genetic difference between the two sporophytes, as might well be the case in a population reproducing primarily by intragametophytic selfing. In fact, most sporophytes of this species were homozygous at all loci tested. Paired-spore and paired-gametophyte cultures allowed intergametophytic mating to relieve inbreeding depression that might prevent intragameto- phytic selfing in either gametophyte. Thus any increase in sporophyte pro- duction in paired cultures relative to isolated cultures of the same species is assumed to have resulted from intergametophytic mating. Two spore sources were used in studies of Campyloneurum angustifolium, C. phyllitidis, and Phlebodium aureum. Genetic load estimates obtained from the two sources were not significantly different for C. phyllitidis and P. aur- eum. Estimated genetic load did differ significantly between the two sources of C. angustifolium in which it was low for source A, but very high for source B. This suggests that the B sporophyte of C. angustifolium was from a highly outcrossing population, whereas the A sporophyte was derived from a population with a higher level of inbreeding. Both diploid and tetraploid chromosome counts have been reported for C. angustifolium (n=37, Evans, 1963, from Peru; n=74, Evans, 1963, from Costa Rica; Sorsa, 1966, from Costa Rica; Knobloch, 1967, from Jamaica). Thus, judging from a correlation 74 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) 143) ¥ 43. * TABLE 6. Description of mult genoty} percentage in a combination of three Florida populations of Campyloneurum phyllitidis. Lap-a/b Mdh-3a/b Pgm-2a/b Pgm-3a/b 6Pgd-2a/b Percentage 11/22 22/22 11/22 11/11 1/22 33.33 11/33 22/22 11/11 TN/41 11/22 26.67 11/33 14/11 14/41 14/11 11/22 3.33 11/22 22/22 71/41 11/11 41/22 3.33 41/33 22/22 11/14 11/11 11/22 6.67 11/22 APs Lies a jg 11/22 10.00 11/23 22122. 11/11 a ala 22/22 3.33 11/33 22122 11/11 22/22 Ti/22 3.33 11/22 22/22 11/11 14/22 22/22 3.33 11/22 22/22 11/11 ag 22/22 6.67 of polyploidy with low genetic load, it is possible that the A sporophyte and B sporophyte of C. angustifolium were tetraploid and diploid respectively. Both diploid and tetraploid forms have also been reported in Phlebodium aureum (n= 37, n=74, Evans, 1963), and Phymatosorus scolopendria (2n=72, Léve et al., 1977; n=72, Tsai and Shieh, 1983). In Campyloneurum phyllitidis, only tetraploids have been found (n=74, 2n=148, Evans 1963, Nauman 1993), whereas in Polypodium pellucidum, only diploid numbers have been reported (Manton, 1951). Thus genetic load estimates indicate that the sample sporophyte spore sources of C. phyllitidis, P. aureum and P. scolopendria were probably tetraploids, whereas those of Microgramma heterophylla, and P. pellucidum were diploid. In fact, isozyme evidence also indicates that C. phyllitidis and P. aureum are polyploids. Strong evidence for describing the mating system of diploid species in the wild can be obtained from analysis of isozyme electrophoretic patterns. Elec- trophoretic patterns for the species tested in this study, Campyloneurum phyllitidis and Phlebodium aureum, revealed a high level of fixed hetero- zygosity for both, indicating that these samples from Florida are polyploid, probably allopolyploid. Because of this, accurate counts of heterozygous in- dividuals and estimates of outcrossing from isozyme evidence are not possi- ble, although we can state that at least one putatively outcrossed individual (Lap-a'*/b**) was among the 10 samples of C. phyllitidis in the FS popula- tion. No evidence of outcrossing was present in the Florida populations of P. aureum but the extremely low level of genetic variability among sampled plants of Pz aureum (29 of 30 plants were genetically identical) would preclude isozymic detection of most recombinational heterozygotes if they were formed. However, the considerable variability among sporophytes of C. phyllitidis (10 multilocus genotypes among 30 plants tested) allows ample opportunity for detection of unbalanced heterozygotes and three-allele hetero- zygotes (at Lap) that would be produced if outcrossing was frequent (Table 6). Assuming allopolyploidy, intragametophytic selfing, and no mutations, the maximum number of genotypes per locus-pair provides an estimate of the CHIOU ET AL: MATING SYSTEMS 75 minimum number of hybridization events involved in producing an allote- traploid species (Ranker et al., 1994). Thus in Campyloneurum phyllitidis, the single genotype in the population JD suggests that only one hybridization event occurred in the ancestry of this population in Jonathan Dickson State Park. Two genotypes at each of Lap and Mdh-3 of population CH indicate that that population originated from at least two hybridization events. Three genotypes at Pgm-3a/b implies that at least three hybridization events pro- duced population FS. However, one of these genotypes (11/22) could have been generated (11/11 x 22/22) by the low level of outcrossing demonstrated by presence of the Lap 11/23 genotype. Because there are not great distances separating these populations in Florida and because of the high similarity of genotypes among the three populations, we can also consider them as a sin- gle population. In that case the number of hybridization events responsible for the Florida plants is at least three and may be as high as ten. In Phlebodium aureum, the genetic similarity among sampled populations is very high. Isozyme evidence showed no genetic variation in the samples of TS and JD populations and only a single plant representing a distinct genotype in the FS population. Because Campyloneurum phyllitidis and P. aureum are widespread in tropical America and because likely parental diploids are not present in Florida, these Florida genotypes probably repre- sent three to ten separate spore introductions for C. phyllitidis and possibly only one or two for P. aureum. In an electrophoretic study of Polypodium pellucidum, Li and Haufler (1999) found a small but significant (mean fixation index of 0.169 across populations of epiphytic P. pellucidum var. pellucidum) excess of homozy- gous individuals in most (but not all) populations sampled in the Hawaiian archipelago. Since intragametophytic selfing as a dominant breeding system would lead to much higher fixation values in relatively few generations (Hartl & Clark, 1997), the observed level of fixation likely results from a mixed mating system where intergametophytic selfing among gametophytes in populations derived from the same sporophyte is occurring. Thus electro- phoretic analysis of population structure in P. pellucidum is not inconsistent with the implications of our results that intragametophytic selfing in this species is rare, probably curtailed by genetic load. This constraint on repro- duction via single isolated spores likely contributes to the genetic differen- tiation and low values of gene flow between populations estimated by Li and Haufler (1999). The fact that diploidy favors intergametophytic mating whereas tetraploidy favors intragametophytic selfing has been demonstrated (e.g., Masuyama and Watano, 1990). The fixation of different alleles from the different parent spe- cies of allopolyploids may mitigate the expression of recessive lethal genes caused by intragametophytic selfing. In our study, low levels of genetic load, as evidenced by production of sporophytes through intragametophytic self- ing, was well correlated with polyploidy (Table 7). Isolated gametophytes of the diploid species (Polypodium pellucidum, Microgramma heterophylla, and the B source of Campyloneurum angustifolium) produced virtually no 76 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) TaBLE 7. Genetic load, ploidy level and probable mating systems in Polypodiaceae species. Species Genetic load’ Ploidy” Mating system?* C. angustifolium A* 14 P (Evans 1963) Intragam. selfing C. angustifolium B* 100 D (Evans 1963) Intergam. mating C. phyllitidis 35 P (Isozyme; Evans 1963) Intergam. selfing M. heterophylla 99 D Intergam. mating hl. aureum 10 P (Isozyme; Evans 1963) Intergam. selfing Phy. scolopendria 11 P (Tsai & Shieh 1983) Intergam. selfing Pol. pellucidum 98 D (Manton 1951) Intergam. mating ‘Estimated from average sporophyte production by single gametophytes isolated as spores and as young gametophytes, except for C. phyllitidis in which the estimate is from plants isolated as young gametophytes. * Determined from isozyme patterns (isozyme) and/or chromosome counts (reference). D=dip- loid, P= polyploid. * Determined from isozymes and/or genetic load. * C. angustifolium A and B plants are from two different sources. sporophytes through Ssyngamy, whereas isolated gametopytes of the tetra- ploid species (C. phyllitidis, Phlebodium aureum, Phymatosorus scolopen- dria and the A source of C. angustifolium) produced abundant sporophytes through intragametophytic selfing. Gametophytes derived from isolated spores resulting from long-distance spores can produce sporophytes through either intragametophytic selfing or intergametophytic mating, the latter being augmented by antheridiogen- stimulated production of male gametophytes. Species predominantly repro- ducing by intergametophytic mating can maintain high levels of genetic variability, including a high genetic load, but may be very limited in their ability to migrate by long-distance spore dispersal (Peck et al., 1990). Thus a trade-off exists between maintenance of genetic variability on one hand and ease of migration on the other. CHIOU ET AL: MATING SYSTEMS 77 Schneller & Hess, 1995). But, the existence of antheridiogens here could also function to promote bisexuality in isolated plants if individual thalli have reduced ability to attain bisexuality. Gametophytes of these species propa- gate vegetatively by branching (Chiou & Farrar, 1997b). Maintaining an an- theridiogen system may be adaptive in promoting antheridium formation on new vegetatively produced thalli regardless of whether the species is in- breeding or outbreeding. Whether species are outbreeders or inbreeders, breeding behavior must be adaptive to establishment and survival of individuals of those species. In general, inbreeding may be advantageous for initiating new populations fol- lowing long-range spore dispersal where gametophytes are likely to be de- rived from single isolated spores. The opposite strategy, outbreeding, has the advantage of generating and retaining genetic diversity, and high genetic loads carried by outbreeders tend to maintain that mating system. Previous studies have demonstrated that gametophytes of the epiphytic species studied here grow perennially through branching and vegetative prolifera- tions which increase their life span and effective gametophyte size (Chiou & Farrar, 1997b) and produce antheridiogen that facilitates the production of male gametophytes (Chiou & Farrar, 1997a). Both of these characteristics have been proposed to increase the probability of intergametophytic mating (Chiou & Farrar, 1997a; b; Chiou et al., 1998). Here we demonstrate that sporophytes of these species may be produced through either outcrossing or inbreeding, as evidenced by high or low level of genetic load, respectively. Their mating systems may be controlled principally by genetic load and generally correlated with ploidy level. Interwoven with these factors are gametophyte morphology, growth habit, antheridiogen production, and envi- ronmental parameters which together maintain successful reproduction of these species. For outbreeding diploid species, genetic load is possibly the driving force leading to morphological and physiological adaptations promo- ting outcrossing. ACKNOWLEDGMENTS The authors thank Drs. D. B. Lellinger, C. H. Haufler, and L. Hickok for their useful comments, Dr. Paul N. Hinz for help with statistical analysis of gametophyte culture data, the Marie Selby Botanical Garden for allowing collection of spores of Campyloneurum angustifolium, Dr. H. Lu- Rivero for assisting collection of Campyloneurum angustifolium, the Florida ia?) | 5 jor ns) tellow Hammock, Dr. J. B. Miller and Mr. R. E. Roberts for help with collecting Campyloneurum phyllitidis and Phelebodium aureum in Jonathan Dickinson State Park, Mr. M. Owen for assis- tance in collecting Campyloneurum phyllitidis and Phelebodium aureum in Fakahatchee Strand State Preserve, Mr. Keith Fisher for help with locating Phelebodium aureum in Tosohatchee State Reserve, and Mrs. Ming-Ren Huang for help with plant cultures and data recording. This research was supported by the Taiwan Forestry Research Institute (Contribution No. 196). 78 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) REFERENCES Cuiou, W.-L. and D. R. Farrar. 1997a. Antheridiogen production and response in Polypodiaceae species. Amer. J. Bot. 84:633-640. and D. R. Farrar. 1997b. — a morphology of selected species of the wong? Polypodiaceae. Amer. Fern J. 87: , D. R. Farrar, and T. A. cai a eae eee and reproductive biol- ogy in Elaphoglossum Schott. Canad. J. Bot. 76:1967-19 CRIsT, ish Gs dees _R. Farrar. 1983. Genetic load and ue dispersal in Asplenium platy- Canad. J. Bot. 61: ab ip Dass, ee L 1995. Significance of gamatophyte form in tropical, epiphytic ferns. PhD diss. lowa ate University, Ames. —— pe D. R. Farrar. 2001. Significance . eae form in long-distance colonization by mele bigaie ferns. Brittonia 53:352-369. M. Evans, A. 3. New peace a. in the Polypodiaceae and Grammitidaceae. Carolan i. 671-6 Farrar, D. R. 1990. heat and evolution in asexually reproducing independent fern gameto- phytes. ak Bot. 15:98-111. Hartt, D. L. and A. G. CLark. 1997. Principles of Population Genetics. Sinauer Associates, Sunder- and, Massachusetts. Haurier, C. H. and G. J. Gasrony. 1978. Antheridiogen and breeding system in the fern genus Bommeria. Canad. J. Bot. 56:1594—1601. and D. E. Souris. 1984. Obligate mies von in a homosporous fern: field confirmation of a laboratory prediction. Amer. J. Bot. 71:878—-881. , M. D, WinpHaM, and T. A. Ranker. 1990. oo analysis of the Cystopteris tennes- seensis complex. Ann. Missouri Bot. Gard. 77:314—-329. Hooper, E. A. and C. H. HAuFer. 1997. Genetic Hii and ues system in a group of neo- tropical Kae ferns (Pleopeltis; Polypodiaceae). Amer. J. Bot. 84:1664—-1674 K.exowsKkI, E. J. JR. 1968. Reproductive biology and evolution in the pteridophyta. PhD diss. Uni- versity of Sateen Berkeley. ———. 1971. Ferns and genetics. BioScience 21:317-322. . 1979. The genetics and reproductive biology of ferns. Pp 133-170 in A. F. Dyer, ed. The Experimental ain of Ferns. Academic Press, Lon 966. Evolutionary significance of polyploidy i in the pteridophyta. Sci- ence 153: 305-307. Knosiocu, I. W. 1967. Chromosome numbers in Cheilanthes and Polypodium. Amer. J. Bot. 54:461 ae Li, J., and C . 1999. Genetic variation, breeding apres and patterns of diversifica- tion in poantonanp: Polypodinan (Polypodiaceae). Syst. Bot. 2 Love, A., D, Love, and R. E. G. oo SERMOLLI. 1977. eset renait Atlas of Pteridophyta. mer, Vaduz, Liechtenstei Cra Manton, I. 1951. ee of Pabeodieae’ in America. Nature 167:3 Masuyama, S Y. Watano. 1990. Trends for inbreeding in cas pteridophytes. Pl. Sp. Antheridiogens and antheridia development. Pp 435-470, in A. F. Dyer, ed. The enc ldesit regen of ferns. Academic Press, London. , C. E. 1993. Campyloneurum. Pp 328 in Flora of North America Editorial Committee, add. Flora of eet America. Oxford Univ. Press, New York, NY. Nei, M. 1978. Estimation of seu anes and genetic distance from a small number of individuals. Genetics 89:583— Peck, J. H., C. J. Peck, and D. R. viciieae 90. Influences bs life history attributes on formation of local and distant fern populations. Amer Fern. J. 8 Ranker, T. A., LER, P. S. Soitis, and D, E aay he 1989. Genetic evidence for allopoly- ploidy in the Neotropical fern esis eee (Adiantaceae) and the reconstruction of an ancestral genome. Syst. Bot. 14:439-447 CHIOU ET AL: MATING SYSTEMS 79 a, : K. FLoyp, and P. G. Trapp. 1994. Multiple colonization of Asplenium adiantum-nigrum o the ea eget Archipelago. Evolution 48:1364—1370. Roc, i S. 1972. Measures of genetic similarity and genetic distance. Studies in Genetics, + Toss ate 7213:145-153 Petia J. J., C. H. Haurter, and T. A. Ranker. 1990. Antheridiogen and natural gametophyte populations. Amer. Fern J. 80:143-15 and A. Hess, 1995. Antheridiogen syatem in the fern Asplenium ruta-muraria (Asplenia- Soitis, D. E. and P. S. Soitis. 1986a. Electrophoretic evidence for inbreeding in the fern Botrychium asl (Ophioglossaceae). Amer. J. Bot. 74:504—509. and P. S. TIs. 1986b. Polyploidy and breeding systems in homosporous pteridophyta: a eeecnhs preci Naturalist 130:219- d, P.:S, oo 1990. Genetic i pa and among populations of ferns. Amer. Fern J. 84:161— Sorsa, V. 1966. ae studies in the rcs come Amer. Fern J. 56:113-119. Sworrorp, D. L. and R. B. SELANDER. 1989 -1. A computer program for the analysis of allelic variation in acenuanl cites or biochemical systematics, release 1.7. Illinois Natural History Survey. Urban Tsal, J.-L. and W.-C. SHIEH. 1983. A ptenicoienhe survey of the pie e se in Taiwan. (1) Material collection and chromosome observation. J. Sci. Engin. 20:1 VoELLER, B. 1971. Developmental adehes of fern gametophytes: enitian hr biology. Bio- Science 5 267-270. Wertu, C. R. 1989. The use of — data for inferring ancestry of polyploid pteridophytes. 2 lid eo Ecol. 17:117— ——., and M. I. Cousens. 1990. yen the contribution of population studies on ferns. Amer. Fern J. 80:183-190. American Fern Journal 92(2):80—92 (2002) Belowground Distribution and Abundance of Botrychium Gametophytes and Juvenile Sporophytes CINDY JOHNSON-GROH'’, CHANDRA RIEDEL?, LAURA SCHOESSLER®, and KrissA SKOGEN* Biology Department, Gustavus Adolphus College, 800 W. College Ave., St. Peter, MN 56082 phytes m_ * respectively. Botrychium hesperium also has a relatively high density of 478 gameto- [ ith ar The importance of propagule banks (also referred to as seed, spore, or dia- spore banks) in community dynamics has long been recognized (Leck et al., 1989) for flowering plants. Propagules may persist belowground for many years, creating a secure reservoir from which aboveground plants can be density of fern gametophytes resulting from cultivation of spore banks is high, ranging from 57,000 spores m ? (Milberg, 1991) to 5,000,000 m~2 y Hall, Iowa State University, Ames, IA 50011. Evolution and Behavior, University of Minnesota, 100 Ecology Bldg., 1987 Upper Buford Circle, St. Paul, MN 55108 Current address: Argonne Na | ry, ’ . ° ’ a . JOHNSON-GROH ET AL.: BELOWGROUND DISTRIBUTION 81 of propagules (represented by germinated gametophytes) to the aboveground density of sporophytes, all previous studies have reported the remarkable lack of ferns present in the aboveground vegetation despite a belowground reservoir of spores. During and ter Horst (1983), Leck and Simpson (1987), Milberg (1991), Raffaele (1996), Rydgren and Hestmark (1997) and Strickler and Edgerton (1976) all found evidence of high densities of fern propagules where ferns were entirely absent from the aboveground vegetation. Botrychium presents two problems in applying conventional propagule bank techniques. First is the difficulty and length of time required to culture Botrychium beyond germination (Whittier and Thomas, 1993). Botrychium spores require darkness for germination, and in nature Botrychium gameto- phytes require mycorrhizae for growth beyond the two- or three-celled stage. Therefore, it is not possible to culture and quantify the propagules using standard seed bank techniques. The second and more significant problem relates to the different life cycle of Botrychium, relatively little of which is visible aboveground (Fig. 1). Plants produce spores that filter into the soil and germinate in darkness. Fol- lowing germination, a belowground achlorophyllous, fleshy gametophyte is produced. The gametophyte produces gametangia and sexual reproduction occurs, resulting in a belowground juvenile sporophyte. The belowground rhizome is upright and short with mycorrhizal roots (no root hairs) and a single leaf-producing bud at the apex. It takes several years for this juvenile sporophyte to produce a leaf-bearing apex and emerge aboveground (Johnson- Groh et al., 1998). The plants generally produce one leaf annually, but it is common for Botrychium plants to remain dormant belowground in a given year and produce no aboveground leaf (Johnson-Groh, 1997). In addition to these belowground stages, some species reproduce asexually via belowground gemmae, small (0.5—-1 mm) propagules that can indepen- dently give rise to a new plant once detached from the parent plant (Farrar and Johnson-Groh, 1990). Gemmae form on the rhizome and abscise at ma- turity. Upon germination, gemmae develop 4 or 5 short roots prior to the dif- ferentiation of a shoot apex and production of leaves (Farrar and Johnson- Groh, 1990). The first leaves formed are short and slender and do not reach the soil surface. The presence of vegetative reproduction greatly influences the population dynamics of these gemmiferous species. It is common in the field to see two or more leaves of gemmiferous Botrychium emerging in close proximity. Excavation of these clusters usually reveals a large number of belowground sporophytes in various stages of development. The terms propagule bank and diaspore bank as used in other studies imply banks of structures that have been disseminated (e.g. seeds, spores). Except for gemmae, Botrychium gametophytes and sporophytes are not struc- tures designed specifically to propagate or disseminate. Consequently the use of existing terms (propagule, diaspore, spore or seed bank) does not accurately describe the belowground structures of Botrychium. The term belowground structure bank, as used in this paper, includes all belowground structures: gemmae, gametophytes, juvenile plants and spores. 82 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) sporophyte (2n) } spore ground juvenile belowground sporophyte (2n) belowground gametophyte sporophyte (2n) with shoot and root (with gametangia) gemmae development gametophyte fertilization Fic. 1. Botrychium life cycle illustrating the belowground developmental stages. Long-term demographic studies (15 years) of Botrychium reveal that popu- lation numbers are quite variable (Johnson-Groh, 1997). Aboveground Botry- chium population numbers fluctuate independently within and between populations, as well as between years and between different sites. Fire, herbi- vory, herbicides, and timber harvests may have an immediate impact on the aboveground sporophytes (Johnson-Groh and Farrar, 1996). However, the above- ground populations are fairly resilient and rebound following perturbations, although recovery may take several years. Many species of Botrychium are considered rare. Several are listed as criti- cal, threatened, or endangered (Minnesota Department of Natural Resources, 2002). Understanding Botrychium population dynamics, including their be- lowground biology, is necessary to effectively manage these species. A more complete understanding of the belowground structure bank will allow pre- diction of the impact of various management regimes, such as fire or grazing, on these rare species. The goal of this research was to investigate the belowground structure bank of several species of Botrychium. It seems likely that the belowground structures (gametophytes, juvenile sporophytes, gemmae and spores) play a JOHNSON-GROH ET AL.: BELOWGROUND DISTRIBUTION 83 key role in the population dynamics of Botrychium, and it is anticipated that the number of belowground structures should be at least as many as the number of aboveground plants. MATERIALS AND METHODS Sites for sampling the belowground populations of Botrychium were selected in areas that contained a typical density of aboveground plants (~50—400 plants in an area 200 m*). A sampling grid resembling spokes of a wheel was established within each population, and samples were collected with a bulb planter at one-meter intervals along the six 8-meter spokes (Fig. 2). In addition to the 48 spoke samples, a sample was collected from the middle. A bulb planter (5-cm diameter) assured a uniform sample volume of approximately 200 cm® for each sample. Distance from each sample to the nearest aboveground plant, aboveground population density, and notes on the vegetation and general ecology were recorded. Soil samples were collected during the summers of 1988 to 1999 from a variety of sites; these were processed the following academic years (Table 1). Seven species of Botrychium subg. Botrychium were investigated: Botry- chium campestre W. H. Wagner & Farrar, B. hesperium (Maxon & R. T. Clausen) W. H. Wagner & Lellinger, B. gallicomontanum Farrar & Johnson-Groh, B. lanceolatum (S. G. Gmel.) Angstr., B. montanum W. H. Wagner, B. mormo . H. Wagner, B. yaaxudakeit Stensvold & Farrar, B. virginianum (L.) Sw. An eighth species, Botrychium virginianum (L.) Sw. from subg. Osmundopteris, was surveyed from two different sites. Because Botrychium typically grows in mixed species assemblages, it is difficult to locate populations of only one species. The sites sampled for B. campestre, B. hesperium, and B. virginianum consisted only of those species, respectively. All other sites contained more than one species. In all cases, the named species was the dominant species (>75% of aboveground Botrychium). Soil samples were processed using centrifugation, which allows the lighter plant material to be extracted for examination under the microscope (Mason and Farrar, 1989). The soil sample was broken up and washed through a series of soil sieves where larger roots and debris were removed. The sediment was collected and root segments dyed in neutral red and cut to the approxi- mate size of Botrychium gemmae (0.5 mm) were added to the sediment to gauge the success of the procedure. Samples were sieved in successively finer sieves and then centrifuged first in water and then in sucrose solutions. Centrifugation separated the belowground structures from the denser soil particles. The first centrifugation caused dead organic matter to float to the top of the tube, leaving living organic matter and stained roots in the pellet. A second centrifugation in sucrose caused the living material to float. The decanted liquid containing Botrychium plants was examined under the microscope for gemmae, gametophytes, sporophytes, and stained root seg- ments. Gametophytes were usually small (<1 mm) and irregularly shaped, and 84 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) N NW NE section SW SE $s Fic. 2. Sampling design for collecting 49 soil samples for analysis. Circles represent soil sam- ples collected on one of 6 radii. could be identified by the presence of rhizoids, antheridia, and archegonia. Juvenile sporophytes represented a continuum of development following the formation of the embryo (still attached to the gametophyte) through the development of a mature plant. Except for very young sporophytes, all juve- niles had roots that were succulent and lacked root hairs. Very young sporo- phytes, here referred to as embryos, were small (<1 mm), spherical to irregular in shape, lacked rhizoids, and, when detached from the gameto- phyte, had a relatively large scar where they had been attached. Gemmae likewise represented a continuum of development following their formation. At abscission, gemmae were spherical and small (0.5-1 mm). As gemmae began to elongate and form roots, it was impossible to distinguish them from juvenile sporophytes originating from a gametophyte. Examination of rhi- zomes of mature leaf-bearing plants determined which species regularly pro- duce gemmae. RESULTS AND DISCUSSION Results of the belowground analysis are shown in Table 2. The density of aboveground sporophytes ranged from 0.4 m ~~ for B. hesperium and B. virginianum to 16.1 m * for B. gallicomontanum. The proportion of samples that contained belowground structures ranged from 4% for B. campestre to 65% for B. hesperium, with an overall average of 30%. Botrychium campestre and 8B. virginianum, had notably low frequencies. The density of below- ground gametophytes ranged from 10 m * for B. gallicomontanum and one JOHNSON-GROH ET AL.: BELOWGROUND DISTRIBUTION 85 TaBLE 1. Botrychium species surveyed for belowground analysis, their sampling location, and year of collection. Species Location Year B. campestre IA, Plymouth County, 5-Ridge Prairie 1988 B. hesperium WA, Colville Nat. Forest, Big Muddy 1998 B. gallicomontanum MN, Norman County, Frenchman’s Bluff 1995 B. lanceolatum OR, Wallowa-Whitman Nat. Forest, Lostine Creek 1999 B. montanum WA, Colville Nat. Forest, Kerry Creek Seep 1998 B. mormo MN, Chippewa Nat. Forest, Ottertail Peninsula 1995 B. yaaxudakeit AK, Tongass Nat. Forest, Yakutat 1999 B. virginianum (So. MN) MN, Nicolett County, St. Peter 1997 B. virginianum (No. MN) MN, Superior Nat. Forest, Pike Mountain 1999 population of B. virginianum to 738 m-° for B. montanum. The density of belowground juvenile sporophytes ranged from 0 m ° (B. gallicomontanum, B. montanum, and B. virginianum) to 281 m ” (B. hesperium). The average stained root recovery was 84%. Three of the surveyed species are known to produce gemmae (B. cam- pestre, B. hesperium, and B. gallicomontanum). For those jag that pro- duce gemmae, the density ranged from 5,907 gemmae m * for B. campestre to 21 m ° for B. hesperium (Table 3). Gemmae, or gemma-like structures, were also found in samples in which the dominant species does not produce gemmae. For example, B. lanceolatum does not produce gemmae, but gem- mae were found in the samples. Two associated species, B. minganense and B. pedunculosum, were also found at the B. lanceolatum site and are known to produce sparse gemmae, so it is possible that the gemmae extracted be- long to these associated species. (It is impossible to distinguish belowground structures of species except sometimes through absence or presence of gemmae.) Botrychium yaaxudakeit also does not produce gemmae, and it is likely that the gemmae found in this survey were from B. ascendens, which produces numerous gemmae, or even B. minganense, which produces few ae, The high number of gemmae in B. /anceolatum may also be due to a pecu- liar behavior observed in this species of expelling the embryo from the ga- metophyte. Embryos appear to be “ejected” through the degeneration of surrounding gametophyte tissue at a stage very similar in size and morphol- ogy to gemmae produced in other species. This behavior may allow produc- tion of multiple embryos per gametophyte, increasing the total production of sporophytes. Further investigations of this behavior are underway. There is large variation (4-65%) among species with regard to their fre- quency in the soil samples. Whether this is due to natural variation among species in response to environmental differences (e.g., Botrychium campestre may naturally have a lower frequency than B. hesperium) or to chance is not easy to resolve. Two technical problems prohibit easy resolution of this. First, centrifugation and microscopic techniques used for this investigation TasLe 2. Result of Botrychium belowground analysis, with the maximum and minimum values for each column in boldface. sity of ensity of belowground Ratio of below- aboveground Frequency of Density of juvenile Total density of ground to sporophytes belowground gametaphytes sporophytes mere ty o belo ih cir aboveground Species (m*) structures (%) (m~*) (m*) mae (m~*) structures (m™*) plants B. campestre (sig 4 21 198 5907 6126 914 B. hesperium 0.4 65 478 281 21 780 1950 B. gallicomontanum 16,1 27 10 0 4170 4180 260 B. lanceolatum 3.1 61 135 10 520' 665 215 B. mo 12 41 738 0 0 738 615 B. 12.8 46 728 104 0 832 65 B. akeit 1.4 29 281 42 312" 635 454 B. virginianum (So. MN) 0.6 8 73 0 0) 73 122 B. virginianum (No. MN) 0.4 6 10 0 0 10 25 Average (all) 4.7 30 261 fa 1228! 1560 332 Average (moonworts) 6.0 40 324 91 1579 1994 332 ' See discussion for explanation of “gemmae” in soil samples for non-gemmiferous species. (2002) 2 MAAINNN 26 ANN'TIOA *TVNUNOl NYA NVOMANV JOHNSON-GROH ET AL.: BELOWGROUND DISTRIBUTION 87 TABLE 3. Soil surveys that contained gemmae produced either by the dominant primary species or by associated species. Associated species Dominant species found at the sampling Density of known to produce site and known to Dominant species gemmae (m*) gemmae produce gemmae B. campestre 5907 Yes None B. hesperium 21 Yes None B. gallicomontanum 4170 Yes None B. lanceolatum 520 No B. minganense, B. pedunculosum B. yaaxudakeit 312 No B. ascendens are extremely labor-intensive, thereby precluding multiple repetitions of the same species. Second, although the sampling technique used for this study is nominally disruptive, it is prudent to minimize excessive disturbance of rare Botrychium species. For the latter reason, B. virginianum (which is not rare) was included in this survey (despite its position in a different sub- genus) and was sampled twice. These samples reveal a high degree of con- sistency, with nearly identical belowground structure banks, even though the sites surveyed were more than 300 miles apart. It is also likely that the frequency differences among species are increased by the patchy distribution of Botrychium and the probability of sampling those patches. Distributions of plants result from factors of dispersal, survi- vorship, and habitat heterogeneity. Because the distribution of aboveground plants is patchy the same might be expected belowground. The random chance of sampling a dense patch could account for the variation between species. Spores.—Although it is difficult to determine the presence of Botrychium spores in the soil, other spore bank studies have shown high diversity and abundance of fern spores that persist in the soil for many years (Dyer, 1994; Milberg, 1991). It is reasonable to assume that there is a long-lived bank of Botrychium spores. This bank is in constant turnover, receiving a variable annual input of spores and losing spores to predation, loss of viability, and other environmental perturbations. Annual spore production is the primary means of restocking the spore bank. Spore production, however, can be expected to vary from year to year, depending on the number of aboveground plants and their development in any given year. Johnson-Groh and Lee (in press) found that only 55% of B. gallicomontanum and 39% of B. mormo produced spores in 1996. The re- mainder of the plants senesced prior to spore production. Similar results have been found for other species (Cabin and Marshall, 2000; Houle, 1998; Kalisz, 1991: Raffaele, 1996). Houle (1998) found temporal differences in the seed rain for Betula alleghaniensis and seedling establishment, but a con- 88 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) stant seed bank. Houle noted that the uncoupling of the seed bank, rain, and seedling establishment further contributes to the spatiotemporal heterogene- ity. It seems probable that similar complex spatiotemporal relationships influence Botrychium distribution. Compounding the variable numbers of spores added annually to the pro- pagule bank is the limited dispersal of spores. It is unknown how widespread moonwort spores disperse but based on the work of Peck et al. (1990) on Botrychium virginianum we can conclude that most spores disperse within 5 m or less. Dyer (1994) found that the largest spore banks occurred in samples taken immediately below ferns and that at a distance of 2 m away from the spore source, the spore bank was notably smaller. It seems likely that a few spores may become airborne and disperse farther. Because of the ability of Botrychium gametophytes to self-fertilize, it is reasonable to expect that a single spore is capable of dispersing and establishing a new population (Farrar, 1998). Over time, a sizeable Botrychium spore bank is established in the soil. Botrychium spores likely remain viable for long periods of time, as do many other ferns (Lloyd and Klekowski, 1970: Miller, 1968; Sussman, 1965; Windham et al., 1986). These spores are probably dormant until conditions (moisture and mycorrhizae) are adequate, at which time many or all the spores in that localized area germinate and develop. densities of 15,000 spores m2: have the lowest at 100 m ”. This tion, mortality at this stage j expect high spore densities within Botrychium populations. JOHNSON-GROH ET AL.: BELOWGROUND DISTRIBUTION 89 Species with gemmae (B. campestre, B. gallicomontanum) have a higher total belowground density than those without gemmae. Like spores, gemmae, once detached from the parent plant, require mycorrhizae for further deve- lopment. Farrar and Johnson-Groh (1990) found relatively few gemmae that contained mycorrhizae, which could explain the low number of developing gemmae relative to the number of gemmae produced. (Gemmae obtain myce- lia through their connection with the parent rhizome; if unsuccessful, they remain dormant.) The primary role of gemmae may be to maintain the popu- lation in a microsite that has already proven successful. The frequent occur- rence of multiple stems within a small-localized area (1-4 cm”) suggests that gemmae are effective in local propagation. Dispersal beyond a short distance is limited, as evidenced by the low frequency of the highly gemmiferous spe- cies (B. campestre, B. gallicomontanum). Species that produce profuse gemmae produce the lowest number of game- tophytes (B. campestre, B. gallicomontanum). Gemmae, a form of asexual reproduction, produce essentially the same genetic product that a selfing gametophyte produces. The advantage of gemma production is the position- ing for immediate success (mycorrhizae present). A greater reliance on repro- duction via spores and gametophytes by most species and the higher disperability of spores undoubtedly accounts for the higher frequencies in soil samples of the non-gemmiferous species. The advantage of spore—game- tophyte production allows dispersal to new sites, thereby insuring that “assets are diversified,” which may provide a long-term advantage to the species. To draw further from investment analogies, gemmae are short-term investments with immediate returns, whereas spores are long-term invest- ments with greater evolutionary payback. Som. HETEROGENEITY.—Variations in microtopography, microclimate, parent material, mycorrhizae, and microorganisms all influence soil heterogeneity (Stark, 1994), creating a patchy environment. Because the spatial scale is very small, conditions merely a few centimeters away may not be sufficient to induce germination, thus creating a patchy distribution of plants. Of these variable factors, mycorrhizae are probably the most important for Botrychium. Moonworts require endophytic mycorrhizae for gametophyte and sporophyte development and are dependent on mycorrhizae as a signifi- cant source of carbohydrate, minerals, and water. This observation is based on several peculiar behaviors. First, similar to orchids, moonworts do not emerge every year. They frequently fail to emerge for one to three consecu- tive years, with no subsequent decrease in size or other negative effects (Johnson-Groh, 1998). Second, “albino” botrychiums have been observed Another indication that Botrychium depends relatively little on its own leaves for photosynthesis is the observation that these leaves frequently do not emerge above the litter. In fact only a small proportion of the total popu- lation of some species actually emerges from the litter (Johnson-Groh, 1998). Herbivory and loss of leaves through fire do not affect the size and vigor of plants in the subsequent year (Johnson-Groh, 1998). Finally, if roots and 90 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) leaves of juvenile plants are produced one per year, as in adults, 5-8 years may be required for development from gametophyte to a mature sporophyte with an emergent photosynthetic leaf (Johnson-Groh, 1998). Juvenile plants must rely on mycorrhizae for carbohydrates. Thus, although there has been no physiological studies to confirm this, it seems certain that moonworts (Botrychium subg. Botrychium) may depend largely on mycorrhizae for carbon from other plants, in addition to that produced by their own photosynthesis. If photosynthesis is not critical for this subgenus and mycorrhizae are pri- marily responsible for overall energy budget, then understanding the below- ground biology of Botrychium is imperative. Indeed, assumptions made about the population biology of other ferns may be irrelevant to Botrychium. Health of the mycorrhizal connection may determine juvenile recruitment and survivorship, and Botrychium populations may appear or disappear in accordance with mycorrhizal health. Mycorrhizae play an important role in nutrient acquisition. This may be especially important for Botrychium because of the inability of its roots to forage. Root-foraging has been observed in flowering plants (Caldwell, 1994). It allows them to respond to small-scale nutrient patches. However, Botrychium roots are relatively few (5-30/plant), do not have root hairs, and do not appear to have the morphological plasticity to forage for small-scale patches of soil nutrients. Typically roots extend almost perfectly horizontally for their entire length (3-20 cm). Only occasionally are roots observed to abruptly bend in another direction. Tibbet (2000) argued that mycorrhizae are especially important for roots that do not have the morphological plasti- city to respond to small-scale nutrient patches. Mycelia rapidly colonize patches of soil nutrients, making them ideal foraging instruments of the au- totroph. In Botrychium it seems highly probable that its mycorrhizal mycelia are more important than root proliferation in nutrient acquisition. Botrychium species that have high belowground densities generally have high aboveground population densities. Botrychium campestre, B. gallico- montanum, and B. mormo have the highest below- and aboveground den- sities. Botrychium virginianum has a relatively low below- and aboveground density. The ratio of belowground to aboveground plants ranges from 65:1 in B. mormo to 914:1 in the gemmiferous B. campestre. For each species in this study, a relatively small volume of soil (962 cm”) was sampled across a large spatial grid (201 m?*). Estimations of density are derived from these results. The patchy distribution and inability to sample large volumes of soil make it difficult to determine precise belowground populations. The opposite approach of intensively and completely surveying both above- and belowground in a small area is currently being investigated for B. virginianum and B. campestre. Preliminary results indicate that the belowground density in a small patch is several times the aboveground density. Despite the difficulty of making direct comparisons below- and above- ground, in all cases it is readily apparent that belowground structures are much more abundant than aboveground plants. Sizeable reservoirs of JOHNSON-GROH ET AL.: BELOWGROUND DISTRIBUTION 91 belowground structures are extremely important to the population dynamics of Botrychium and serve to replenish the populations following environmen- tal perturbations. Because moonworts often remain dormant in any given season, they are essentially protected a aacoreh pe and can easily withstand dry years, fires, herbivory, or other aboveground disturbances. Despite highly variable aboveground populations, belowground stages provide Botrychium populations with a high degree of buffering against local extinction. The belowground structure bank is a reserve of plants in various stages that can eventually produce an aboveground population, regardless of past above- ground perturbations. Dyer (1994) noted the importance of fern propagule banks to the conserva- tion of fern species. Reintroducing or augmenting populations from the spore bank broadens the options available for many species. Whereas it is difficult to manipulate Botrychium belowground structure banks, it is important to recognize the importance of these banks to the overall population dynamics. Modeled Botrychium populations (Johnson-Groh et al., 1998) predict greater stability of populations than would be concluded from monitoring only aboveground plants. This resiliency is a result of the large belowground re- serve of gametophytes and juvenile sporophytes capable of regenerating the population. The long-term impact of environmental perturbations on popula- tions is buffered by a large bank of belowground structures. AACKNOWLEDGMENTS We acknowledge partial funding provided by the Colville and Tongass National Forests and Gustavus Adolphus College. We are grateful to Kathy Ahlenslager and Mary Stensvold for their interest, support, and help with this study. We thank Don Farrar for discussions and for provid- ing data on species examined in his laboratory. LITERATURE CITED CaLDWELL, M. M. 1994. ae nutrients in fertile soil microsites. Pp. 325-347 in M. Caldwell and R. earcy, eds. Exploitation of environmental heterogeneity by ae Academic es as mi ego. Canin, R. J. and D. L. MarsHALL. 2000. The demographic role of soil seed banks. I. Spatial and temporal sien of below- and aboveground populations of the desert mustard Les- querella eee J. Ecol. 88:283-—292. Durinc, H. J. and B. TER ip 1983. The diaspore bank of bryophytes and ferns in chalk grass- land. Lindbergia 9:5 Dyer, A. F. 1994. Natural es spore banks—can they be used to retrieve lost ferns? Biodiversity and Conservation 3:1 and S. Linpsay. 1992. Soil spore banks of temperate ferns. Amer. Fern J. 82:89-123. Farrar, D. 3 1998. dh trai ones of moonwort Botrychiums. Pp. 109-113 in N. Berlin P. Miller, J. Borovansky, U. S. Seal, and O. Byers, eds. Population and habitat abba seni (PHVA) for hee ae fern anual mormo) Final Report. Conservation Biology Specialist Group, Apple Valley, M and OHNSON-GROH. 1990. rissa he i gemmae in moonwort ferns, Bo- trychiu umm subgenus Botrychium. Amer. J. Bot. 77:1168-1175. HamiLTon, R. G. 1988. The significance of spore { {) Zi OO a elite, SOMITE) sss Z, Os ; ea LL Pe Ose sp. LM. part of fertile frond; and part of stipe; B. Distal ules. Adiantum squamulosum. A. Rhizome C. Rachis scales; D. Abaxial view of fertile pinn Pas. 1. PRADO & SMITH: NOVELTIES IN PTERIDACEAE 107 pinna conform, 1—1.5 times longer than the subtending pinnae, 0.8—1 times as long as medial pinnae; indument of costae like that of stipes and rachises; pinnules (11)41-46 pairs per pinna, ca. 2 times longer than wide, charta- ceous, not articulate, free-veined, without an evident midrib, the proximal pinnules reduced, somewhat rounded or rhombic, the medial pinnules di- midiate, trapeziform to oblong and with the acroscopic base truncate, the sterile apices obtuse to acute, sterile margins denticulate, fertile apices angu- lar, distal pinnules ca. 1/2 or less as long as the medial pinnules; both pin- nule surfaces scaly, the scales sparse, ferrugineous, 1.0-1.5 mm long but otherwise similar to those of the stipes, glands absent, veins slightly promi- nulous, idioblasts present between the veins; sori oblong, up to 7(9) per pin- nule; indusia scaly, scales with filiform apices and pectinate bases, entire to erose, the cells well evident; spores yellowish, trilete, tetrahedral-globose, with prolonged angles, the surface rugulate. Adiantum squamulosum is distinguished by having densely scaly stipes and rachises, large scales on both surfaces of the pinnules, a relatively large number of pinnule pairs per pinna (up to 46), and scaly indusia. This spe- cies occurs in partially disturbed, primary, non-inundated forest at low ele- vations (ca. 230 m) along road margins. Adiantum diogoanum Glaz. ex Baker is probably the most closely related species but differs in having stipes, ra- chises, and pinnules glabrescent, fewer pinnule pairs per pinna (up to 24), and indusia with reddish hairs, rather than scales. It grows in drier, inland forests along rivers and slopes. Adiantum squamulosum is known from a single locality in Bolivia, whereas A. diogoanum has a wide range in Brazil, occurring in the states of Pernam- buco, Alagoas, Minas Gerais, So Paulo, and Parana. The type was cited by Smith et al. (1999, p.247) as questionably A. humile Kunze. Pteris boliviensis Prado & A. R. Sm., sp. nov. (Fig. 2A—C).—Type: BOLIVIA: Depto. Cochabamba: Prov. Chapare, 22 Feb 1929, 1700 m, J. Steinbach 9327 (holotype UC). Ad P. lividam Mett. affinis, a qua imprimis frondibus pinnatis (vs. triparti- tis) et 2 vel 3 areolae inter contiguas costulas (vs. 1 areola) differt. Plants terrestrial. Rhizomes short-creeping, densely clothed at apex with light brown ovate-lanceate scales 2-3 mm long, the scale margins entire and glabrous. Fronds erect, to ca. 52 cm long; stipes ca. 1/2 of frond length, up to 3 mm in diam., straw-colored, dark brown at base, grooved adaxially, smooth, glabrous except for a few ovate-lanceate scales at base; laminae chartaceous, 2-pinnate-pinnatifid at base, with 1-4 pinna pairs per frond, 1- pinnate-pinnatifid above the base, ending in a broadly based, pinnatifid api- cal pinna; rachises similar to stipes, alate distally, glabrous; proximal pinna pair stalked (stalks ca. 0.5 cm long), opposite, deeply pinnatifid, furcate, the basal basiscopic pinnules 12-16 x 4.5—7.0 cm, the remainder of the pinnae 20-22 x 5-7 cm; median pinnae sessile, subopposite, deeply pinnatifid, de- current on the rachis, 13-14 5.5—-6.5 cm; distal pinnae sessile, alternate, AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) PRADO & SMITH: NOVELTIES IN PTERIDACEAE 109 decurrent on the rachis, 8-9 X 2.5—3.0 cm; costae not sulcate on both sides, glabrous, awns absent adaxially; pinna segments subopposite to alternate, oblong, the margins entire at the middle to crenate-serrate at the apex, the terminal segment of each pinna narrowly deltate, acute or acuminate, si- nuses between segments roundish to biangulate; venation copiously areolate but free near margins and segment apices, with 2 or 3 costal areoles between adjacent costules, the costules slightly prominent especially on abaxial sur- faces. Sori interrupted at sinuses and absent at apex of segments; indusia greenish when young, margins entire; spores brown to tan, trilete, tetrahe- dral, the surfaces rugulate with roundish tubercles and a smooth equatorial flange. PaRATYPE.—BOLIVIA. Cochabamba: Pcia. José Carrasco Torrico: 137 km antigua carretera Cocha- bamba-Villa Tunari, 17°06’S, 65°35’W, 1600 m, 18 July 1996, M. Kessler et al. 7393 (paratype UC; isoparatype LPB not seen). The two or three areoles between adjacent costules distinguish this species from Pteris livida Mett., which has only one areole between adjacent costules. In addition, the fronds are ternate in P. livida and pinnate in P. boliviensis. Pteris boliviensis grows in wet forests and cleared forests at ca. 1600— 1700 m. Pteris krameri Prado & A. R. Sm., sp. nov. (Fig. 2D, E).—Type: GUYANA: Upper Essequibo Region: Rewa River, near camp 2 at foot of Spider Moun- tain, forest on light brown sand, 03°08’N, 0°58’W, 16 Sept 1999, M. J. Jansen-Jacobs et al. 5923 (holotype UC; isotype U on 2 sheets). A P. altissima Poir. costis aristis adaxialiter carentibus differt. Plants terrestrial. Rhizomes stout, woody, creeping, densely clothed at apex with linear-attenuate, bicolorous scales, 3-5 mm long, these with a blackish to reddish brown central portion and entire to more or less erose, glabrous, pale margins. Fronds erect, ca. 1.3 m long; stipes ca. 2/3 of frond length, up to 6 mm in diam., straw-colored, grooved adaxially, smooth, glabrous except for a few bicolorous scales at base; laminae chartaceous, 2: pinnate-pinnatifid at base, 3-7 pinna pairs per frond, the median portion 1-pinnate-pinnatifid, fronds ending in a broadly based pinnatifid apical pin- na; rachises similar to stipes, narrowly alate distally, glabrous or with sparse, whitish hairs ca. 1 mm long; proximal pinna pair stalked (stalks ca. 1.0 cm long), opposite, deeply pinnatifid, furcate, the basiscopic pinnules (12-22 x 3.7-6.0 cm, the remainder of the pinnae 29-35 x7.5—10.0 cm; median pin- nae sessile, subopposite, deeply pinnatifid, 22-30 x 7.5-8.5 cm; distal pinnae — Fic. 2. Pteris species. veins; C. Detail of fertile segments. D-E lamina showing veins, costa, and costules. A-C. Pteris boliviensis. A. Habit; B. Detail of a sterile pinnae showing _ Pteris krameri. D. Median pinnae; E. Adaxial view of 110 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) sessile, alternate, 13-14 x 3.9-4.5 cm; costae not sulcate adaxially, glabrous or with sparse, minute, whitish hairs ca. 1 mm long, awns absent; segments subopposite to alternate, long-lanceate, the margins entire to conspicuously serrate at apex in fertile fronds (sterile fronds not seen), the terminal seg- ment of each pinna elongate-deltate, acute to long-acuminate or sometimes caudate, the sinuses between the segments acute to roundish; venation copi- ously areolate but free near margins and segment apices, with 2 or 3 areoles between adjacent costules, veins slightly prominent especially on abaxial surfaces. Sori interrupted at sinuses and absent at segment apices; indusia greenish or brownish, the margin entire; spores tan, trilete, tetrahedral, the surfaces rugulate with roundish tubercles and smooth equatorial flange. Pteris krameri is apparently most similar morphologically to P. altissima Poir., and both species occur in similar habitats: near river and stream mar- gins, moist ravines, in rocky soils or silt from big rocks, or on humus-rich soils in forests. However, P. altissima differs in having an awn at the base of each costule on the adaxial side of the lamina. Although it has been inad- equately assessed, the character of presence or absence of awns along the pen- ultimate axes seems of fundamental importance in assessing relationships in Pteris, so the apparent similarity of P. krameri to P. altissima may be the result of convergent evolution. Other New World species that lack awns adaxially, and that have pinnatifid pinnae or pinnules, are P. angustata (Fée) C. V. Morton, P. boliviensis Prado & A. R. Sm., P. brasiliensis Raddi, P. congesta Prado, P. decurrens C. Presl, P. denticulata Sw., P. fraseri Mett. ex Kuhn, P. lechleri Mett., P. leptophylla Sw., P. limae Brade, and P. pearcei Baker, all species of South America and especially Brazil (Prado and Wind- isch, 2000). None of these species seems closely related or similar morpho- logically to P. krameri. Whether the presence or absence of adaxial awns on the lamina is a character that circumscribes monophyletic groups is an intrigu- ing but unanswered question. Although Pteris krameri is known only from the type gathering, it may have a wider range in northern South America. The species epithet honors the late Karl Kramer, who contributed much to the study of fern systematics and made Pteris one of his specialties; Dr. Kramer also focused especially on the ferns of the Guianas, the source of this new species. Brade 13494 (holotype RB, not seen, frag. HB!). This rather uncommon species can be distinguished by its delicate, terete stipes with uniseriate, jointed hairs, laminae proximally with 1-3 pairs of free pinnae and distally pinnatifid. It is endemic to southeastern Brazil PRADO & SMITH: NOVELTIES IN PTERIDACEAE 111 (Minas Gerais). The new name proposed below honors Alexander Curt Brade, who described this species in Notholaena (Brade, 1940). SPECIMENS STUDIED.—BRAZIL. Est. Minas Gerais: Estrada Diamantina—Curvelo, Williams & Anderson 8466 (UB); Santana do Riacho, Serra do Cip6, Parque Nacional da Serra do Cip6, Cachoeira da Farofa, CFSC 10240, Prado et al. (SPF); Idem, Serra do Cipé, Sena s.n. (RB, HB); Idem, Schwacke 14520 (BHCB); Ouro Preto, Itacolomi, Schwacke 9906 (RB); Lima Duarte, Krieger s.n. (BHCB); Mun. Gouveia: Rib. do Tigre, Hatschbach 27841 (MBM, UC). Pteris bakeri C. Chr. According to Tryon and Stolze (1989. p. 80), Pteris bakeri C. Chr. is en- demic to Peru and occurs in forests at middle to high elevations, 2300-3000 m. This is the second collection outside of Peru; Arbeléez (1996) also re- ported this species from the Colombian Andes. Pteris bakeri can be distinguished by fronds ca. 1 m long, laminae decom- pound (5-pinnate at the base), very small ultimate segments 1-1.5 mm wide, free veins, and spiculate, scaly axes abaxially. SPECIMENS STUDIED.—BOLIVIA. Depto. Cochabamba: Pcia. José Carrasco Torrico: 116 km antigua carretera Cochabamba-Villa Tunari, 17°08’S, 65°38’W, 2350 m, Kessler et al. 7064 (LPB not seen, UC); Idem, 123 km antigua carretera Cochabamba-Villa Tunari, 17°08’S, 65°37’W, 2100 m, Kess- ler et al. 7117 (LPB not seen, UC). ACKNOWLEDGMENTS We thank curators of the Missouri Botanical Garden (MO) and Utrecht (U) Herbaria for loans of types of Adiantum squamulosum and Pteris krameri, respectively. We also thank Sra. Emiko Naruto for preparing the illustrations. LITERATURE CITED ARBELAEZ, A. A. L. 1996. La tribu Pterideae (Pteridaceae). Flora de Colombia 18:10—106. BRaADE, A. C. he Filices novae Brasilianae VI. Anais Reuniaéo Sul-Amer. Bot. 2:5—-11. PRADO, J. a _G. Wmnoiscu. 2000. The genus Pteris L. (Pteridaceae) in Brazil. Bol. Inst. Bot. (Sao ee 13:103-199. SmitH, A. R., M. Kessier, and J. Gonzates. 1999. New records of pteridophytes from Bolivia. Amer. Fern J. o 244-266. Tryon, R. M. and R. G. STOLzE. 1989. er ele of Peru. Part II. 13. Pteridaceae-15. Denn- staedtiaceae. an Bot., n.s. 2 American Fern Journal 92(2):112—118 (2002) Is Gametophyte Sexuality in the Laboratory a Good Predictor of Sexuality in Nature? ToM A. RANKER and HEATHER A. Houston University Museum and Department of Environmental, Population, and Organismic Biology, 218 UCB, University of Colorado, Boulder, CO 80309-0350 USA than what is found among natural populations, although this may be true for any lab-based study regardless of growth medium. Thus, we suggest caution in the use of agar as a growth medium, and the use of laboratory conditions in general, for studies of fern gametophyte sexual development. The gametophytes of homosporous ferns possess the ability to be either unisexual (antheridiate or archegoniate) or bisexual. In addition, they may undergo developmental sequences involving changes in sexuality over time due to genetic and/or environmental factors (Klekowski, 1969a; Greer & McCarthy, 1999). Because gametophytes of most species are easy to grow in culture, many studies have been conducted on cultured gametophytes to antheridiogen production and response (e.g., Stevens & Werth, 1999), and populational genetic load/isolate potential (Peck et al., 1990). Numerous studies have employed data from lab-cultured gametophytes to make inferen- ces about mating systems operating in nature (e.g., Soltis & Soltis, 1990, and references therein; Chiou et al., 1998: Li & Haufler, 1999). Although many studies have examined gametophytes that were cultured on mineral-enriched agar, several have employed various natural substrates (Rubin & Paolillo, 1983; tion in common: that the patterns and processes observed in the laboratory are indicative of what is occurring in nature. At least one study, however, has demonstrated significant differences in the sexual expression of gametophytes RANKER & HOUSTON: GAMETOPHYTE SEXUALITY—GOOD PREDICTOR 113 collected from nature versus those cultured on mineral-enriched agar. Schneller (1979) found that 79.9% of gametophytes of Athyrium filix-femina (Woodsiaceae) collected from the wild were sexual, whereas those cultured were only 60.1% sexual. The proportions of gametophytes that were asexual, antheridiate, archegoniate, or bisexual differed between the two populations: wild-collected were 21.1% asexual, 49.6% antheridiate, 20.6% archegoniate, and 8.6% bisexual; lab-cultured were 39.9%, 30.5%, 27.3%, and 2.3%, respec- tively. These two distributions differ statistically (2 x 4 contingency table, lab vs. field by sexual category, x7 = 100.55, df = 3, p < 0.001; our analy- ses). That study demonstrates that, at least for gametophyte populations of A. filix-femina, the sexual expression of gametophytes cultured on mineral- enriched agar was not a good indicator of the sexual expression of gameto- phytes occurring in the wild population studied. Such comparisons are crucial to assess the veracity of data collected from artificially cultured gametophytes for inferring patterns and processes occurring in nature. Ranker et al. (1996) studied the sexual expression of lab-cultured gameto- phytes of two species of the endemic Hawaiian genus Sadleria, S. cyatheoides Kaulf. and S. pallida Hook. & Arn., grown on mineral-enriched agar. That substrate was chosen primarily for the sake of convenience and the ease of gametophyte manipulation. Observations were made on a total of 5,749 gametophytes from 26 sibships of S. cyatheoides and 3,032 gameto- phytes from 15 sibships of S. pallida (see Ranker et al., 1996, for explicit methods). All gametophytes were sampled from multigametophyte cultures. Although the two species differed statistically in the exact proportions of gametophytes that were antheridiate, archegoniate, or bisexual, the overall pattern among sexual gametophytes was the same in the two species: a small proportion was antheridiate (grand means of 9.7% and 10.0% from S. cyatheoides and S. pallida, respectively), a larger proportion was arche- goniate (84.7% and 81.2%), and a small proportion was bisexual (5.6% and 8.8%). The predominance of unisexual gametophytes was consistent with analyses of sporophyte-population surveys of allozymic variation from which Ranker et al. (1996) inferred that sporophytes of both species primarily arose via intergametophytic matings. Another aspect of the study of Ranker et al. (1996) involved assessments of the ability of Sadleria spp. gametophytes to produce and respond to an- theridiogen (Dépp, 1950; Naf, 1979). Based on an excess of antheridiate ga- metophytes in treatment vs. control cultures, Ranker et al. (1996) concluded that an antheridiogen system was operating in these species. These results are at least superficially incongruent with the data from the sexual expres- sion studies described above; that is, one might expect to find a greater pro- portion of antheridiate gametophytes from multi-gametophyte cultures, due to the action of antheridiogen, than was observed in either species. We conducted field studies on the sexual expression of natural popula- tions of Sadleria spp. gametophytes. Our primary goal was to compare field- collected data to the lab-based data of Ranker et al. (1996) to determine if the laboratory data were an accurate reflection of what is occurring in nature. 114 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) In particular, we were interested in comparing gametophyte populations cultured in the laboratory to those sampled in the field for: 1) the ratio of asexual to sexual gametophytes; 2) the ratio of unisexual to bisexual gametophytes, among sexual gametophytes and, 3) the ratio of antheridiate to archegoniate gametophytes, among unisexual gametophytes. MATERIALS AND METHODS Five populations of Sadleria spp. sporophytes were located on the Hawaiian Island of Kauai in November, 1999. Three populations were mixtures of S. cyatheoides and S. pallida (populations CP-1, CP-2, and CP-3), one popu- lation was pure S. squarrosa (population S-1), and one population was pure S. unisora (population U-1). Populations CP-1, CP-2, CP-3, and U-1 were from the Hanalei District and population S-1 was from the Waimea District; exact localities are available from the first author. Samples of earthen or bryophyte-covered substrate approximately 10 cm by 10 cm by 1-2 cm (deep) were collected with a kitchen spatula from the vicinity of Sadleria sporophytes. Five to 10 of such samples were collected at each site. Samples were transported in plastic bags to the herbarium of the National Tropical Botanical Garden. Each sample was then inspected under a binocular dissecting microscope for the presence of Sadleria gametophytes, which could be distinguished from those of other species of ferns by the presence of distinctive glandular trichomes that are present on both gametophytes and sporophytes. Sadleria gametophytes were gently removed from the substrate, rinsed in water, and categorized under a compound microscope as belonging to one of four classes of sexuality: asexual, antheridiate, archegoniate, or bisexual. Only gametophytes with mature archegonia and antheridia were scored as bisexual. RESULTS We found and recorded observations from a total of 200 gametophytes across the five field populations surveyed, with individual sample sizes ranging from 26 to 67 (Table 1). All gametophytes sampled possessed notch meristems. ASEXUAL vs. SEXUAL GAMETOPHYTES.—The percent ratio of asexual to sexual ga- metophytes in individual field populations ranged from 4:96 to 41:59 and, summed across populations, that ratio was 20:80 (Table 1). The results summed across populations were significantly different from what was observed in the laboratory, where nearly the opposite percent ratio (74:26, asexual:sexual) was obtained across all sibships of both species (2 x 2 contingency table of asexuality/sexuality vs. lab/field, y° = 286.6, df = 1, p < 0.001). Similarly, the results summed across just those field populations likely to be mixtures of gametophytes that were conspecific with those cul- tured (CP populations) also exhibited a percent asexual:sexual ratio (18:82) RANKER & HOUSTON: GAMETOPHYTE SEXUALITY—GOOD PREDICTOR 115 TaBLE 1. Sexuality of sampled gametophytes. Population N' Asexual? —- Sexual”_~— Antheridiate* Archegoniate* Bisexual” Field results: 49 10 (20) 39 (80) 34 (87) 1 (3) 4 (10) CP-2 26 1 (4) 25 (97) 9 (36) 13 (52) 3)(72) CP-3 67 14 (21) 53 (79) 41 (77) 9 (17) 3 (6) S-1 27 11 (41) 16 (59) 15 (94) 1 (6) 0 U-1 31 3 (10) 28 (90) 27 (96) 0 1 (4) Total 200 39 (20) 161 (80) 126 (78) 24 (15) 14 (7) Laboratory results’: a oides 5749 4375(76) 1374 (24) 133 (10) 1164 (85) 77 (5) S. pallida 3032 2089 (69) 943 (31) 94 (10) 766 (81) 83 (9) Total 8781 6464(74) 2317 (26) 227 (10) 1930 (83) 160 (7) * Sample size. 2 Number (percent of total rounded to nearest whole percent). 3 Number (percent of sexual gametophytes rounded to nearest whole percent). 4 Data from Ranker et al. (1996), summed across sibships. that was significantly different from that of laboratory gametophytes (y* = 220.9, df = 1, p < 0.001). The percent ratio of asexual:sexual summed across laboratory sibships of S. cyatheoides was 76:24 and that in S. pallida was 69:31 UNISEXUAL vs. BISEXUAL GAMETOPHYTES.—Among the sexual gametophytes sampled, all field populations exhibited a predominance of unisexual game- tophytes (Table 1). Field populations did not differ statistically from labora- tory populations in the ratio of unisexual to bisexual gametophytes, both summed across all field populations (y” = 0.001, p > 0.05) and summed across just the CP populations (%* = 0.462, p > 0.05). ANTHERIDIATE VS. ARCHEGONIATE GAMETOPHYTES.—As with the ratios of asexual to sexual gametophytes, field-collected and lab-cultured gametophytes ex- hibited nearly opposite percent ratios of antheridiate to archegoniate gameto- phytes. The overall percent ratio in the field was 84% antheridiate to 16% archegoniate, among CP populations that ratio was 79:21, and that in the lab- oratory was 11:89. These field and laboratory proportions were significantly different from each other in total (x2 = 584.2, df = 1, p < 0.001) and when comparing only CP populations to laboratory populations (y* = 397.6, df = 1, p < 0.001). DISCUSSION ted populations of gametophytes d interpretation of data from lab- of other taxa as well. The results of our study of field-collec have important implications for the use an cultured gametophytes of Sadleria spp. and, possibly, ASrxuAL vs. SEXUAL GAMETOPHYTES.—The striking difference between labora- tory and field gametophytes in the relative proportions of asexual-to-sexual 116 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) gametophytes suggests that the laboratory conditions employed by Ranker et al. (1996) were a poor mimic of the natural environment in terms of simply allowing or forcing gametophytes to become sexual. Those laboratory condi- tions were similar to those that have been employed by numerous other investigators in studies of fern gametophytes. Thus, some factor or combina- tion of factors in our laboratory cultures appears to have been inhibiting the development of sexual organs. One potential impact on inferences resulting from the laboratory data is that they would lead to an underestimate of the sexual reproductive potential of a population or species relative to those based on field observations, although Ranker et al. (1996) did not explicitly make any such inferences. UNISEXUAL vs. BISEXUAL GAMETOPHYTES.—In both laboratory and field popula- tions of gametophytes there was a predominance of unisexual gametophytes among those that were sexual. These results are similar to those from A filix-femina (Schneller, 1979) and from Blechnum spicant (Blechnaceae; Cousens, 1979, 1981). Among sexual field-collected gametophytes of B. spicant, 21% were bisexual and 79% were unisexual (Cousens, 1981). Cousens (1979) agar-cultured gametophytes of B. spicant as isolated gameto- phytes and at moderate and high densities. Among sexual isolates, only 8% were bisexual and among both groups of multi-gametophyte cultures, 20% exhibited bisexuality. In terms of assessing the likelihood of unisexuality vs. bisexuality and, thus, the relative likelihood of intergametophytic mating vs. selfing, gametophytes cultured on mineral-enriched agar appear to behave in a manner consistent with development of gametophytes in nature. ANTHERIDIATE VS. ARCHEGONIATE GAMETOPHYTES.—Among gametophytes of Sadleria spp. and A. filix-femina (Schneller, 1979), there were significant differences in the ratio of antheridiate-to-archegoniate gametophytes in the laboratory vs. in the field [using Schneller’s data from A. filix-femina, we analyzed a 2 x 2 contingency table of antheridiate vs. archegoniate by lab vs. field; y7 = 29.07, df = 1, p < 0.001]. In both cases, laboratory popula- tions were generally mostly archegoniate and field populations were mostly antheridiate. These results are similar to those of Rubin and Paolillo (1983) where unisexual gametophytes of Onoclea sensibilis (Woodsiaceae) that were agar-grown were disproportionately archegoniate whereas those that were soil-grown were disproportionately antheridiate. Also, in field populations of B. spicant unisexual gametophytes were predominantly antheridiate (Cousens, 1981). The relative proportions of antheridiate vs. archegoniate gametophytes in laboratory populations of B. spicant, however, varied with density (Cousens, 1979), Among isolated gametophytes, all were archegoniate and all but one (98%) were archegoniate when cultured in moderate density (only one was antheridiate). By contrast, when cultured at high density, 96% of unisexual gametophytes were antheridiate and only 4% were archegoniate, similar to gametophytes collected from nature, Cousens attributed these differential patterns to a greater likelihood of antheridiogen effects at higher densities in the laboratory. RANKER & HOUSTON: GAMETOPHYTE SEXUALITY—GOOD PREDICTOR 117 Laboratory studies of S. pallida, S. cyatheoides (Ranker et al., 1996), A. filix-femina (Schneller, 1979), and B. spicant (Cousens, 1979) have demon- strated that gametophytes of all of these species can produce and respond to the male-inducing pheromone antheridiogen, as has been shown for many other species of ferns (e.g., Naf, 1979; Schneller et al., 1990). Thus, as was suggested for B. spicant, it is possible that antheridiogen has a greater effect in natural populations of Sadleria spp. and A. filix-femina than on agar- cultured populations. The possibility that an agar-based medium directly in- hibits antheridia formation and/or promotes archegonia formation, however, cannot be ruled out (see Rubin & Paolillo, 1983). Similarly, the impact of potential differences in the age structure of natural and cultured populations on sexual expression should also be considered. In laboratory cultures, gametophyte populations typically are established from single sowings of spores, resulting in nearly even-aged populations. In nature, gametophyte populations would presumably be composed of mixed-age individuals. Because gametophytes do not respond to antheridio- gen after they reach the notch-meristem stage and because most gametophytes in an even-aged stand would reach that stage essentially simultaneously, few cultured gametophytes would become antheridiate due to antheridiogen response. That would be particularly true at low to moderate densities. By con- trast, in a mixed-aged natural stand, antheridiogen-producing gametophytes may often already be present in a location when new spores arrive, thus causing newly produced gametophytes to become antheridiate. Thus for ex- perimental studies of the mating systems of species in which an antheridio- gen system is controlling sexuality in nature, and depending on the goals of a study, it is important to establish laboratory conditions that allow for ef- fective antheridiogen-mediated interactions. In conclusion, we found that for generally inferring mating systems operating in nature (i.e., relative likelihood of intra- vs. intergametophytic mating), agar-based laboratory studies of gametophytes would lead to the same conclusions as would observations of field-collected gametophytes in Sadleria spp. For detailed studies of gametophyte sexuality and develop- ment, however, an agar-based medium produced significantly different results than what we found among natural populations of gametophytes. Whenever an even-aged population is created in the laboratory, however, this may be true on any growth medium. Thus, as have others before us, we suggest caution in the use of agar as a growth medium, and laboratory condi- tions in general, for studies of fern gametophyte sexual development. ACKNOWLEDGMENTS We dedicate this paper to the memory of Herb Wagner for his constant inspiration, encourage- ment, support, and friendship. This study was funded by a grant from the University of Colorado Graduate School. Permission to collect specimens was granted by the State of Hawaii, Division of Land and Natural Resources. We are grateful to Tim Flynn, Steve Perlman, Ken Wood, and Diane Ragone of the National Tropical Botanical Garden for generously providing assistance in the field and logistical support. 118 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) LITERATURE CITED Cuiou, W.-L., D. R. FARRAR, and T. A. RANKER. 1998. peri be morphology and reproductive biology in i ean Schott. Canad. J. Bot. 76:1967—-19 Cousens, M. I. 1979. Gametophyte ontogeny, sex expression, and sco load as measures of merch Pie divergence in Blechnum spicant. Amer. J. Bot. 66:116—132. . 1981. Blechnum spicant: habitat and vigour of optimal, argina and disjunct popula- tions and field observations of gametophytes. Bot. Gaz. 142:25- Dopp, W. 1950. Eine die Antheridienbildung bei Farnen fordernde ees in den Prothallien 9—159. Dyer, A. ng pies The culture of fern gametophytes for experimental investigation. Pp. 253-305 in A. F, Dyer, ed. The Experimental Biology of Ferns. Academic Press, London. a ee K. 1993. The influence of soil topography and spore-rain density on gender expression in ee populations of the homosporous fern Aspidotis densa. Amer. Fern ie mans . on McCartuy. 1999. Gametophytic plasticity among four species of ferns with con- trasting ecological iy cotinine Int. J 9-886 Haurter, C. H. and D. E. Soxtis. 1984. Obligate outcrossing in a s easeeuoiorciia fern: field confir- mation ofa a laboratory sear a Amer. J. Bot. 71:878-881. KLEKOWSKI, JR., E. J. 1969a. oe biology of the pteridophyta. II. Theoretical considera- tions. Bot. J. Linn. Soc. 62:347-359. . 1969b. ecole sd of the pteridophyta. III. A study of the Blechnaceae. Bot. J. Linn. vps Li, J. W. an os pale 1999. Genetic variation, breeding systems, and cis of diversifi- cation in Hawai sige ogg (Polypodiaceae). Syst. Bot. 24:339-35 Linpsay, S. . F. Dyer. 1996. Investigating the phenology of secre development: an alt ae rien Pp. 633-650 in J. M. Camus, M. Gibby, and R. J. Johns, eds. Pteri- ology in Perspective. Royal Botanic Gardens NAr, U. 1979. Antheridiogens and antheridial adtalcnsaides: Pp. 435-470 in A. F. Dyer, ed. The Experimental Biology of Ferns. Academic Press, London. Pancua, E., S. Linpsay, and A. Dyer. 1994. Spore germination and gametophyte development in three species of Asplenium. Ann. Bot. 73:587—-593. Peck, J. H., C. J. Peck, and D. R. Farrar. 1990. ct ea se Ppa studies and the distribu- tion of pteridophyte populations. Amer. Fern J. 8 Ranker, T. A., C. E. C. GeEMMiLL, P. G. Trapp, A. HAMBL LETON, ate ss Ha. 1996. Population genetics and reproductive ee! of lava-flow colonising species of Hawaiian Sadleria (Blechna- ceae). Pp. 581-598 in J. M. Camus, M. Gibby, and R. J. Johns, eds. Pteridology in Perspec- tive. Royal Botanic “Na a Kew Rusin, G. and D. J. PAoLi1o, Jr. 1983. Scud development of Onoclea sensibilis on agar and soil media without the addition of antheridiogen. Amer. J. Bot. 70:811-815. SCHNELLER, J. J. 1979. cn a investigations on the Lady Fern (Athyrium filix-femina). PI. Syst. Evol. 132:255-2 . H. HAUvFLEr, vied “ A. Ranker. 1990. Antheridiogen and natural gametophyte popula- ian Amer. Fern J. 80:143-152, Soxtis, P. S. and D. E. =e 1990. Evolution of i lies. Pl. Sp. Biol. 5:1- TEVENS, R. D. and C. R. sina 1999. Interpopulational comparison of dose-mediated antheridi- ogen response in Onoclea sensibilis. Amer. Fern J. 89:221-231. nbreeding and outcrossing in ferns and fern-al- American Fern Journal 92(2):119-130 (2002) Additional Support for Two Subgenera of Anemia (Schizaeaceae) from Data for the Chloroplast Intergenic Spacer Region trnL-F and Morphology J. E. SkoG Biology Department, George Mason University, Fairfax, VA 22030 E. A. ZIMMER Laboratory for Molecular Systematics, Smithsonian Institution, Washington, DC 20560 J. T. MICKEL New York Botanical Garden, Bronx, NY 10458 Asstract.—An analysis of morphological data for 13 species with 33 characters and molecular data for 14 species from the chloroplast DNA intergenic spacer region trnL-F indicates that spe- cies of the genus Anemia fall into two well-supported subgenera, Anemiorrhiza and Anemia. In addition, one species of the genus Mohria appears to belong within Anemia. Although further study is required, these data support the relationships suggested by a previous study of fossil and extant representatives of the genus. The fern genus Anemia Sw. (Schizaeaceae) comprises about 120 species distributed mainly within the tropics and subtropics. Most of the species are found within the New World, with only about 12 in Africa and one in India. No monograph of the whole genus has been produced, although subgenus Coptophyllum (Mickel, 1962), subgenus Anemiorrhiza (Mickel, 1981), and spores of subgenera Coptophyllum and Anemia (Hill, 1977, 1979) have been studied. Since those works were produced, several new species have been described (Mickel, 1982, 1984, 1985), and a study of some of the Cretaceous fossils within the genus completed (Skog, 1992). In addition, the spores of the family Schizaeaceae have been described for modern and fossil represen- tatives (van Konijnenburg-van Cittert, 1991, 1992). Other fossil representa- tives of the family have been described from Mesozoic and Cenozoic time periods and are summarized in Skog (2001) and Collinson (2001). There is clearly a need for a new revision of the genus and integration of all morpho- logical data from fossil and living species with data from new molecular studies. We began a collaborative study in 1999. This paper reports some re- sults indicating that the chloroplast sequence data are consistent with the fossil phylogeny reported earlier by Skog (1992). The Schizaeaceae, considered to be a basal family of leptosporangiate ferns, includes the genera Lygodium, Schizaea, Actinostachys, Mohria, and Anemia. Fossil records of the family extend back to the Jurassic (Skog, 2001). The position of this family is not clear; it generally falls among sev- eral clades, including the Hymenophyllaceae, Cyatheaceae, Schizaeaceae, Matoniaceae, aquatic ferns, and more derived ferns (Raubeson & Stein, 1995; Pryer et al., 1995; Pryer et al., 2001). However, there is strong support from 120 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) the morphological and molecular evidence to date that the genera within the Schizaeaceae form a well-supported clade (Pryer et al., 1995; Pryer, 1999; Wikstrém et al., 2000, pp. 149-150). A molecular study using the chloroplast gene rbcL has been done for the Schizaeaceae and this study included sev- eral species of Anemia (Wikstrém et al., 2000). MATERIAL AND METHODS Species from which DNA was isolated and sequenced, as well as the voucher information are presented in Table 1. Morphological characters (Table 2, 3) for the preliminary analysis were derived from the descriptions and characters cited in the literature (Mickel, 1962, 1967, 1981, 1982; Skog, 1992; Tryon & Tryon, 1982). Three subgenera (Anemiorrhiza, Coptophyllum, and Anemia) were repre- sented in the samples. Both herbarium specimens and living material dried in silica gel were sources for the DNA. More recent herbarium specimens provided sequence data, but many older specimens yielded little, if any, DNA. Other taxa included were one species of Mohria, which is often sug- gested as congeneric with Anemia (Mickel, 1962), two species of Lygodium, one species of Schizaea, and a species in the Hymenophyllaceae (Cardio- manes reniforme) as the outgroup. The Hymenophyllaceae is also considered fairly basal among the ferns and has been shown to be more basal than the Schizaeaceae in analyses using morphology and rbcL chloroplast data (Pryer et al., 1995, 2001). Total DNA was extracted from 15-20 mg of leaf tissue dried in silica gel (then frozen in liquid nitrogen) with the DNEasy Plant Mini kit from Qiagen, following the manufacturer’s protocol. Double-stranded DNA amplifications were performed in a 50 ul volume containing 5 pl MgCl, 4 ul dNTP, 5 ul buffer, 2 ul primers, 1 ul BSA, 5 ul DNA, 0.5 pl TAQ, and 25.5 ul distilled water to volume. Following an activation step of 10 min at 94°C for the en- nutes before being held at 4°C. A few species were sequenced for the cpDNA gene rbcL using the primers previously published for fern sequences (Hasebe et al., 1995). The trnL (UAA)-trnF (GAA) intergenic spacer was amplified Sequence data were generated for both strands of PEG (polyethylene gly- col) purified PCR product using the ABI PRISM dye terminator cycle se- quencer (Applied Biosystems, Inc.): 4 min at 96°C; then 25 cycles of 10 sec at 95°C, 0.5 sec at 50°C, 4 min at 60°C; followed by 1 min at 96°C; and then SKOG ET AL.: ADDITIONAL SUPPORT FOR TWO SUBGENERA 121 T . List of species (with subgenera of Anemia in paraentheses), voucher specimens, and localities for the material used in the molecular analysis. Species (Subgenus) Specimen Locality Anemia adiantifolia (Anemiorrhiza) Bradley 30582, GMUF Bahamas, Andros Isl. (L.) Sw. A. cicutaria (Anemiorrhiza) Kunze Bradley 30691, GMUF Bahamas, Andros Isl. ex. Spreng. A. jaliscana (Anemia) Maxon Mickel 1689, NY Mexico, Jalisco A. munchii (Anemia) Christ Mickel 6874, NY Mexico, Oaxaca A. semihirsuta (Anemia) Mickel Mickel 1120, NY Mexico, Oaxaca A. hirsuta (Anemia) (L.) Sw. Prado 1064, NY Brazil, Sao Paulo A. phyllitidis (Anemia) (L.) Sw. Mickel 1121, NY Mexico, Oaxaca A. underwoodiana (Anemia) Maxon Greuter & Rankin 24997, B Dominican Republic A. villosa (Coptophyllum) Humb. & Salino 1771, NY Brazil, Sao Paulo Bonpl. ex Willd. Mohria cafforum (L.) Desv. NYBG 136/97 (living coll.) South Africa Lygodium microphyllum (Cav.) R. Br. NYBG 1245/89B (living coll.) unknown L. flexuosum (L.) Sw. NYBG 1281/76B (living coll.) unknown Schizaea elegans (Vahl) Sw. H. Tuomisto 12744, TUR Peru, Loreto Cardiomanes reniforme (Forst.) C. Presl A.R. Smith 2606, UC New Zealand, No. Isl. Long Ranger gel. The resulting chromatograms were edited with Sequencher version 3.1 (Gene Codes, Inc.) and the final consensus sequences were ex- ported for alignment to Clustal X. All sequences were aligned manually with the aid of Se-Al version 1.0a1 (Rambaut, 1996) multiple sequence editor fol- lowing initial alignment in the program Clustal X. Alignment of the rbcL da- ta was guided by the sequences of Pryer et al. (1995). The trnL-F sequences were aligned by eye within the genus and between genera using the editing program Se-Al. Sequences were submitted to GenBank (Accession numbers AF448922-AF448935), and the alignment used in phylogenetic analyses will be posted at www.l|ms.si.edu. All phylogenetic reconstruction analyses were conducted using version 4.08b of PAUP* (Swofford, 2001). Phylogenetic reconstruction under maxi- mum parsimony was conducted using both the heuristic search algorithm with TBR branch-swapping, MULPARS and ACCTRAN options active, as well as the Branch and Bound search option in PAUP*. For molecular data, characters were assigned equal weights at all nucleotide positions; for mor- phological characters, equal weighting was also used. Robustness of cladistic linkages was evaluated with 1000 bootstrap replicates. RESULTS Results from the few species sequenced for rbcL were in agreement with the study done by Wikstrém et al. (2000). After their abstract was published, we dropped this gene from our study and accepted their discussion of a phy- logeny based on that gene. Their topology has not yet been published. The plastid DNA sequence trnL-F was chosen because it has been shown to have AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) TABLE 2. Character names and states used in the morphological analysis. 1. rhizome habit 12. spat cell shape 24. apical plate cells 0: scarab horizontal sodiametric 0: thick-walled 1: creeping [ at ate Ls ein 2: sich te upright 13. stomata 25. spore shap 2. stele 0: attached 0: tetrahedral 0: dictyostele 1: floating 1: round 1: solenostele 14. laminar hair 2: bilateral, elongate 3. axillary pockets 0: uni- or ltl 3: tetraglobose (rounded) 0: absent 1: multic 26. spore ornamentation 1: present 2: ies 0: narrow sulci, wide muri 4. arrangement of leaves 15. laminar trichomes 1: wide sulci, narrow muri 0: polystichous 0: glandular 2: granulate 1: distichous 1: nonglandular, pointed 27. a n =¢ ao) ® Q ° = o = 1: variable or med. brown 6. trichomes on rhizome 2: nonglandular, broad . fertile pinna position rizontal rect fl ne apnaeigg i) fos) groove ornamentation : smooth 1: granulate . ridge ornamentation th 0: smoo 0: scales 17. 1: granulate 1: hairs long (4-10 mm) chains erile 2: spinose 2: hairs short (1-3 mm) Neon ate 29. spore size 7. rhizome trichome color 18. fertile iy length : <50um : 0: shorter than sterile blade 1: 50-60 1: dark brown or maroon 1: exceeding sterile blade 2: 60-80 2: yellowish brown 2: ca. equal to sterile blade 3: 385 8. stipe size 19. fertile pinna differentiation 30. spore ridge — 0: slender (0.6 mm) 0: undifferentiate 0: parallel, 1: medium (1.0 mm) : basal pair differentiated 1: coarsely anne 2: stout (2-3 mm) 2: fronds fully dimorphic 2: finely reticulate frond division 3 22 pairs differentiated 31. fertile “eon on pinna pinnate 20. 0: all fert 4: i all ih sterile 2: at tips Sa ste . Stipe sha N fo —_ ind i") | = w N 10 21. fertile pinna insertion : terete 0: several ni (4-8) 0: sessile 1: somewhat flattened 1: many (10-15 +) 1: stalked 2: flattened 2: none 22. veins 33. indument on stipe 11. pinna rai shape 0: free 0: hairs scattered : truncate 1: anastomosing 1: abundant orange hairs 1: as 23. sporangia shape 2: abundant white hair: 0: oval, oblong 3: abundant stiff black hairs 1: spherical 4: subglabrous relatively high and even substitution rates among the plastid loci within an- giosperms (Richardson et al., 2000) and it has been successfully used for phylogenetic analyses among species within a genus of angiosperms (Molvray et al., 1999). It had not been used for many analyses within the ferns, but has been used for a study in a eusporangiate fern family (Hauk et al., 1996), and we have had some success in a study of the Osmundaceae (in progress). Within that region, only sequences between the “e” and “f” SKOG ET AL.: ADDITIONAL SUPPORT FOR TWO SUBGENERA ria, Actiniostachys, and Schizaea. Character coding is given in Table 2. Tapte 3. Character state matrix for ten species of Anemia and one each of Moh Species abbreviations for Anemia use the subgeneric designations given in Table 1. 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 6010 11 32 13 14 13 16 77 O20 a 1 12 3°4 5 6 Res (9G 0 ee 9 ae Taxa Ar. adiantifolia Oo Oe Ha a A tO ° oO o o o _ oO o © An. phyllitidis ow i=) o o i=) So o o ct o Oo - o ise) i) o o 1 80: 2 3 As. pennula 124 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) primers were amplified successfully. The region amplified between primers “c” and “‘d” in all angiosperms studied in the Laboratory for Molecular Sys- tematics (Smithsonian Institution) to date did not work, even with optimiza- tion attempts using a Stratogene Robocycler. We suspect that either an accelerated rate of evolution in that subregion of the cpDNA has resulted in the primers’ failing to anneal or that the inversion and rearrangements of the chloroplast genome, reported for ferns by Raubeson and Stein (1995), will account for the amplification difficulties. The latter may be more probable, as we successfully amplified the trnL-F “‘c-d’’ region for the Osmundaceae (Bauder, Skog & Zimmer, unpubl.), but not for the higher fern genus Elapho- glossum (Skog et al., 2001, p. 87). Within the Schizaeaceae, the trnL “e-f’ spacers had a GC content from 36-40%; the outgroup Cardiomanes GC content was 32%. These numbers were similar to those obtained in Zimmer’s previous work on the basal angiosperm families Winteraceae and Canellaceae (Karol et al., 2000). For the taxa studied, there was significant length variation among the sequences. Within the Schizaeaceae, the trnL sequences ranged in length from 387 base pairs (bp) (Schizaea elegans) up to 525 bp (Anemia cicutaria); in the outgroup Cardiomanes reniforme, the trnL region sequenced was 483 bp long. Within previously described subgenera of Anemia, the range of spacer lengths was much more limited. For subg. Anemia, Spacers were 494-506 bp long; for subg. Coptophyllum the single trnL region sequenced from Anemia villosa was 515 bp long; for the two species of subg. Anemiorrhiza, A. adianti- folia at 524 bp was a single nucleotide shorter than A. cicutaria. Mohria cafforum, sometimes proposed as being nested within Anemia, had a trnL “e-f” region of 493 bp. The two Lygodium species were 481 and 489 bp long. Given the degree of variation in spacer lengths among the Anemia species and among taxa at the family level in the Schizaeaceae, we had to infer a number of indel events when aligning the sequences. The overall alignment required gaps to be inserted into the raw sequences to a final length of 803 po- sitions. As would be expected from the raw sequence lengths, most of the gaps that were inferred differentiated Anemia (including Mohria) from the other two genera of Schizaeaceae and that family from the outgroup Cardiomanes. alignment used in subsequent phylogenetic inference will benefit from a much broader sampling of taxa across Anemia and within Schizaea and Lygodium. However, the relatively lower degree of length variation within and between Anemia and Lygodium suggests that the general phylogenetic patterns pre- sented below will not be strongly affected by increased taxon sampling. In the raw aligned data set for just the trnL‘‘e-f” region, 372 characters were constant, 180 characters were parsimony-uninformative, and 251 characters gions could be aligned with Canellaceae outgroup genera for a total length of 990 characters and where 913 characters were constant, 60 characters were SKOG ET AL.: ADDITIONAL SUPPORT FOR TWO SUBGENERA 125 parsimony uninformative, and only 17 were parsimony-informative (Karol et al., 2000). The high degree of length variation and large number of parsimony informative characters seen with just the smaller PCR product from the trnL region suggest that additional chloroplast spacer regions will be extremely useful in delineating species relationships in the Schizaeaceae. Both the Heuristic search and the Branch and Bound search options in PAUP* produced one most parsimonius tree of length 704 steps, with a consistency index of 0.875 and a retention index of 0.804. This tree is pre- sented in Figure 1, with branch lengths given in Figure 1a and bootstrap values in Figure 1b. The same topology is obtained when the gap characters were ignored, Cardiomanes was excluded from the analysis, and Schizaea and Lygodium were set as outgroups. In addition, we used our data for trnL from the various genera (set as outgroups) within the Osmundaceae (Osmunda, Todea, Plenasium, Leptopteris) in the analysis and obtained the same topology for the tree. Maximum likelihood analyses of the data, using either the default options in PAUP* (Tn/Tv ratio = 2) or a more complex HKY85 +G+I model, where base frequencies and the proportion of invariant sites were calculated from the data, also yielded a single tree with a topology identical to that obtained in the parsimony analysis (data not shown). The morphological analysis is incomplete. To date, 56 species of the genus have been coded for 64 characters; however, the data were pruned to only 13 species and 33 characters for this paper (Tables 2, 3), which were ob- tained from species descriptions in the literature (Mickel, 1962, 1967, 1981, 1982; Skog, 1992; Tryon & Tryon, 1982). These 13 species were chosen be- cause we also had material for DNA extraction. We present a brief outline of the results from the pruned morphological analysis, as it suggests several in- teresting hypotheses to be tested (Figure 2). A single most parsimonious tree was obtained from this reduced data set, shown in Figure 2a with branch lengths and Figure 2b with bootstrap values. Eighteen characters were syna- pomorphic for various clades. Subgenus Anemiorrhiza forms a distinct clade within the genus (66% bootstrap). The traditional subgenera Anemia and Coptophyllum cluster together with 52% support for a clade of these subge- nera and Mohria. There is strong support for Mohria to be included within the genus Anemia (98%). The one species of Mohria and the one species of subg. Coptophyllum form a clade with 69% support. Obviously the expanded morphological data set is necessary, as more species of subg. Coptophyllum and Mohria are needed to determine if the relationship receives continued support from both morphological and molecular data. At the moment, only morphological data indicate this relationship. If these relationships are to be resolved, additional molecular data is also needed for the rest of the species of Mohria and additional species of Anemia. DISCUSSION In their published abstract, Wikstrém et al. (2000) noted that a maximum parsimony analysis of 30 living species of Schizaeaceae indicated that Schi- 126 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) An. jaliscana An. munchii An. semihirsuta An. hirsuta Ac. villosa An. underwoodiana An. phyllitidis Mo. cafforum Ar. adiantifolia Ar. cicutaria Sc. elegans . Ly. flexuosum - Ly. microphyllum Ca. reniforme 2a > Ar. adiantifolia Ar. cicutaria Ar. wrightii Ac. villosa Mo. cafforum An. hirsuta An. munchii An. phyllitidis An. underwoodiana An. jaliscana An. semihirsuta Sc. elegans As. pennula An. jaliscana An. munchii An. semihirsuta An. hirsuta Ac. villosa An. underwoodiané An. phyllitidis Mo. cafforum Ar. adiantifolia Ar. cicutaria Sc. elegans Ly. flexuosum Ly. microphyllum Ca. reniforme 2b ” Ar. adiantifolia 66 Ar. cicutaria Ar. wrightii An. hirsuta 98 An. munchii 51 An. phyllitidis An. underwoodiana 52 =f Ac. villosa Mo. cafforum An. jaliscana An. semihirsuta eC Sc, elegans cerns f\ pennula SKOG ET AL.: ADDITIONAL SUPPORT FOR TWO SUBGENERA 127 zaea and Lygodium were monophyletic. Anemia was paraphyletic to Mohria, as was subg. Coptophyllum to subg. Anemia. They noted that Anemiorrhiza was a sister-group to a clade with all the other species of Anemia and Moh- ria. They stated that within the family there was a long branch leading to the species of Lygodium. The outgroup used in their analysis was not mentioned. The actual tree topology from their study has not yet been pub- lished and was not included in their abstract. Our data are not completely consistent with their study. The trnL tree agrees with Wikstrém et al. (2000) in the placement of taxa within Anemia (Figure 1). However, Schizaea forms a strongly supported monophyletic clade with Anemia (79% bootstrap), and only the two Lygodium species are n a separate long branch relative to the outgroup Cardiomanes. Mohria clearly falls within Anemia (100% support), and basal to the clade of subge- nera Anemia and Coptophyllum (77% support). The single species of subg. Coptophyllum falls within subg. Anemia, and this clade plus Mohria are sis- ter to subg. Anemiorrhiza, which has 100% support on our tree. We believe that subg. Coptophyllum and subg. Anemia should be combined into a single subg. Anemia. There are no good morphological characters to support the separation of these two subgenera, and the molecular data do not support their separation either. Furthermore, we suspect that when additional spe- cies and characters are included in the analysis, subg. Coptophyllum and Anemia will form a strong single clade within the genus. There is, however, strong support for subg. Anemiorrhiza as a monophyletic taxon within an expanded genus Anemia (including Mohria). As Mickel (1962) noted, Mohria is congeneric with Anemia. Traditionally the separation of these two genera was based mainly on the possession of scales by Mohria, but, as noted by Skog (1992), these scales have filiform tips similar to the trichomes found in Anemia and the trichomes on the leaves of the two genera are identical. Mohria bears sporangia on all pinnae of the fer- tile frond, but the confinement of the fertile pinnae to the basal pair is not always definitive. Some species of Anemia have dimorphic fronds (Mickel, 1984). According to van Konijnenburg-van Cittert (1991, 1992), the spores of Mohria show a hollow area in the ridges of the exospore that is not found in — ious tree derived from DNA sequences trnL ‘‘e-f” region using r maximum parsimony settings in PAUP*. Branch lengths are Bootstrap consensus tree for the molecular bove the lines leading to the lineages. Spe- Fic. 1a. The single most parsimon the branch-and-bound option unde given above the lines leading to each lineage. Fig. 1b. data; bootstrap support values above 50% are given a cies are those listed in Table 1. The abbreviations are: Ac = Anemia subg. Coptophyllum, An = Anemia subg. Anemia, Ar = Anemia subg. Anemiorrhiza, Mo = Mohria, Sc = Schizaea, Ly = Lygodium, and Ca = Cardiomanes. Fic. 2a. The most parsimonious tree based on the preliminary morphological data set, using a. same options as above. Branch lengths are given above the lines leading to each lineage. Fig. 2 : Bootstrap consensus tree for the morphology data set. Bootstrap values are given above the line. The abbreviations are those used in Figure 1, plus As = Actinostachys. 128 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) e spores of Anemia subg. Coptophyllum and Anemia. She also noted (1992) that hollow ridges occur in subg. Anemiorrhiza. Mickel (1962, Plate IV) discussed and illustrated three types of ridge structure: a solid ridge with no internal differentiation, a medulla within the ridge containing a spongy network, or a simple medulla, which may be a small portion of the ridge as in Mohria and in some species of Anemia. We see few substantial arguments to maintain Mohria and Anemia as separate genera, although we await the outcome of expanded analyses that include more than a single species of Mohria. Hill (1977, 1979) suggested that spore morphology might be an important character within Anemia because the spore morphology was a conservative character. Even in our preliminary analysis of morphological characters, we see that the spore morphology of ridges, grooves, and bacculae will help to delimit taxa, and that more species require examination for the critical char- acteristics of the spores. Some characters that should be included are, for ex- ample, elaborations at the angles of the spore outline, orientation of the ridges, width of the striations, cross ridges between the ridges, and ornamen- tation on the ridges and between the ridges. The trnL data support the phylogeny previously inferred from the morpho- logical study of fossil and modern species (Skog, 1992). No outgroup was used in that analysis. At the time, no fossil representatives were known for the Hymenophyllaceae earlier than the appearance of the Schizaeaceae; char- acters from the fossils known were few, leading to a matrix with many miss- ing characters; and relationships within the primitive fern families were not stable or easy to discern. However, the analysis of the species within Ane- mia and the fossils attributed to the genus or closely aligned to it indicates that there is strong support for the subg. Anemiorrhiza, that subg. Coptophyl- lum is paraphyletic to subg. Anemia, that these form a sister group to Ane- miorrhiza, and that Mohria and two fossil species form another clade that is not placed consistently in either of the other two clades and was better sup- ported as a third group (Skog, 1992). Interesting questions remain concerning Anemia. Subgenus Anemiorrhiza is consistently diploid, whereas the other subgenera have various levels of the greatest current morphological diversity for the genus. However, the fam- ily Schizaeaceae first appears in the northern hemisphere in the Mesozoic, as does Anemia (Skog, 2001), suggesting biogeographic questions might be addressed when the phylogeny of the genus is better known. There are also SKOG ET AL.: ADDITIONAL SUPPORT FOR TWO SUBGENERA 129 morphologically intermediate between the fossil species, which had com- plete fertile fronds, interspersed fertile pinnules, or basal fertile pinnules. Whether these extant species with nonextended fertile pinnae form a grou or are dispersed throughout the genus will be of benefit in the understanding of the development of the fertile fronds and the fertile pinnae. In our pre- liminary morphological trees for the expanded data set, species with nonex- tended fertile pinnae are currently scattered throughout the tree. Based upon our data, we support the phylogeny of Anemia suggested pre- yap # by fossil data (Skog, 1992) and rbcL data (Wikstrém et al., 2000, pp. 149-150). We support only two of the three subgenera of the current genus yon p Oe Se and Anemia (including subg. Coptophyllum). We be- lieve that Mohria should be placed in Anemia. ACKNOWLEDGEMENTS This paper is dedicated to the memory of Warren Herb Wagner (1920-2000). Professor Wagner taught Skog the modern ferns and encouraged her study of fossil ferns. He was the mentor, advi- sor, and friend of Mickel. We also thank Youngbae Suh for DNA sequencing instruction and Molly Nepokroeff and Ken Karol for advice on sequence editing and alignment, as well as phy- logenetic analyses. We are indebted to reviewers Alan Smith, Don Farrar, and Tom Ranker for their helpful improvements to the manuscript. LITERATURE CITED CoLLINSON, M. 2001. Cainozoic ferns and their distribution. Brittonia 53:173-235. Hasese, M., P. G. Wotr, K. M. Pryer, K. Uepa, M. Ito, R. Sano, G. J. Gastony, J. YOKOYAMA, J. R. Manuart, N. Murakami, E. H. CRANE, C. H. HAUFLER, and W. D. Hauk. 1995. Fern phylogeny based on rbcL nucleotide sequences. Amer. Fern J. 85:134—181. Haux, W. D., C. R. Parks, and M. W. Cuase. 1996. A comparison between trnL-F intergenic nai and rbcL DNA sequence data: an example from Ophioglossaceae. Amer. J. Bot. 7126. Hi, 5 R. 1977. Spore morphology of Anemia subgenus pein Amer. Fern J. 67:11-17. . 1979. Spore morphology of iat subgenus Anemia. Amer. Fern J. 69:71-79. Karot, K., Y. Sux, G. Scuatz, and E. A. Zimmer. 2000. Molecular evidence for the position of iit in the Winteracess: paca from nuclear ribosomal and chloroplast gene spacers. Ann. Missouri Bot. Gar d. 87:414—-432. Kovac Crrrert, J. H. A. vAN, 1991. Diversification of spores in fossil and extant Schi- ae. Publ. Syst. Assoc., spec. vol. 44:103—-118 ie ; "1992 The evolutionary development of schizaeaceous spores in situ. Courier Forschung- sinst. Senckenberg. 147:109-117. MicxzL, J. T. 1962. A monographic study of the fern genus Anemia, subgenus Coptophyllum. Iowa State Coll. J. Sci. 36:349-482. ———. 1967. The phylogenetic position of Anemia colimensis. Amer. J. Bot. 54:432—-4 —. 1981. Revision of Anemia subgenus Anemiorrhiza a Brittonia 33: eicue ———. 1982. The genus pom Veins) in Mexico. Brittonia ———. 1984. New Tropical American Ferns. Amer. Fern L 74: 111-119. —. 1985. Three new anemias from northern South America. Amer. Fern : 75:33-37. : Mo vray, M., P. J. Kores, and M. W. Cuase. 1999. Phylogenetic relationships within Korthalsella (Viscaceae) based on nuclear ITS and plastid trnL-F sequence data. Amer. J. Bot 86:249- 260 130 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) Pryer, K. P. 1999. Phylogeny of marsileaceous ferns and relationships of the fossil Hydropteris pinnata reconsidered. Int. J. Pl. Sci. 160:931-954 SMITH, and J. E. Skoc. 1995. Phyiopunetia apse ge et ma extant ferns based on hie are from morphology and rbcL sequences. Am ee —282. . SCHNEIDER, A. R. SMITH, R. CRANFILL, P. G. Wore, 18 ae and S. R. Sires. 2001. Sete and een ts are a monophyletic group and the oaest living hens to seed plants. Nature — 618-621 — A. 1996. Se-Al Sécuimics Alignment Editor Version 1. 0 alpha 1. Department of Zoology, Papin of ace South Parks Road, Oxford OX1 4JD, England. Rauseson, L, and D. B. STEIN. 1995. Insights into fern evolution from mapping chloroplast genomes. Tle. Fern J. a 193- 204. RICHARDSON, J. E., M. F. Fay, Q. C. B. Cronk, D. Bownam, and M. W. Cuase. 2000. A a eon analysis Rhamnacoae using rbcL and trnL-F plastid DNA sequences. Amer. J. Bo 87:1309-13 Skoc, J. E. 1992 ann Lower Cretaceous ferns in the genus Anemia (Schizaeaceae), Potomac Group of Virginia, and relationships within the genus. Rev. Palaeobot. Palynol. 70:279- 2 —.. 2001. Pag cad of Mesozoic leptosporangiate ferns related to extant ferns. Brittonia 53:236— 24) Ens em R. Moran, and E. A. Zimmer. 2001. Phylogeny of the fern genus Elaphoglos- sum based on two chloroplast Program/Botany 2001, Botanical Society of America. ruman State pit Sapa: Sworrorp, D. 2001. PAU Lh analysis using parsimony, version 4. 08b. Laboratory of Molecular ce Smithsonian Institution, Washington, D.C., and Sinauer, Sunder- land, TABERLET, P., L GrELLy, G. Panton, and OUVET. 1991. Universal primers for amplification of three non-coding regions of cloopiast DNA. PI. Molec. Biol. 17:1105-1109. Tryon, R. M. and A. F. Tryon. 1982. Ferns and allied plants with special reference to tropical merica. Springer-Verag ey Yo WikstROM, N., P. KENRICK, and J. C. VocEL. 2000. Schizaeaceae: a phylogenetic approach. Abstr 6th Conf. Int’l. Ong Palaeobot., july 31—-August 3, 2000, Hebei, China. see taeear ae Committees of Palaeobotanical Society and Botanical Society, China. American Fern Journal 92(2):131-149 (2002) Intrafamilial Relationships of the Thelypteroid Ferns (Thelypteridaceae) ALAN R. SmiTH and RAYMOND B. CRANFILL University Herbarium, 1001 Valley Life Sciences Building #2465, University of California, Berkeley, California 94720-2465 ABSTRACT.—Data from three chloroplast genes (rps¢4 + trnS spacer, + trnL spacer; 1350 base pairs) for 27 of the recognized segregates show the Thelypteridaceae to be monophyletic and sis- ter to an unresolved alliance of blechnoid, athyrioid, onocleoid, and woodsioid ferns. The family comprises two primary lineages, one phegopteroid, the other thelypteroid (including cyclosor- oid). The phegopteroid lineage (Macrothelypteris, Pseudophegopteris, and Phegopteris) includes those elements that are the most dissected, lack adaxial grooves on the frond axes, and are gen- erally morphologically the most distinct elements within the family. Within the thelypteroid- cyclosoroid lineage, three predominantly north-temperate subgroups, including Thelypteris s.s., form a free-veined clade that is in turn sister to the rest of the family. All segregates possessing x=36 (Cyclosorus sensu Smith, with predominantly anastomosing veins) form a strongly sup- ported clade. Those groups with dysploid base chromosome numbers (x= 27, 29 St, 32,33; 34, 35) form a series of smaller clades basal to Cyclosorus s.]. Although our sampling is not yet sufficient to favor one classification over another, recognition of an intermediate number of genera may be the most reasonable taxonomic course. ww 2 Since its taxonomic separation from the dryopteroid ferns as a distinct group, about 60 years ago, Thelypteridaceae has been treated as a natural group comprising nearly 1000 mostly tropical species. Although generally recognized as a natural monophyletic group, there is a wide divergence of views about generic circumscription. Morton (1963) placed all species in a single genus Thelypteris. Ching (1963) outlined a classification that accepted 18 Asian genera, including Hypodematium, which is generally excluded from the family by other workers, even by Ching (1978). Ching (1978) later added two newly described genera to Thelypteridaceae: Craspedosorus (which we regard as a synonym of Leptogramma, often included in Stegno- gramma sensu Iwatsuki, 1963); and Trichoneuron (regarded by us as belonging to Lastreopsis, a dryopteroid, definitely not thelypteroid). Both of these genera are monotypic and poorly known. Still later, Shing (1999), subsumed Amphi- neuron under Cyclosorus s.J. and removed Trichoneuron from the family. Iwatsuki (1964) recognized three genera in the family, Stegnogramma and Meniscium (each with four sections), and Thelypteris, the last comprising 14 subgenera and several additional sections; Iwatsuki (1964:23) regarded two of his subgenera of Thelypteris (Haplodictyum and Cyrtomiopsis) as probably ‘generically distinct.’ Holttum (1971, 1982) characterized 25 genera in the Old World but did not explicitly address New World groups. Pichi Sermolli (1977:335-337), largely following Holttum, accepted 32 genera. In the most re- cent classification, Smith (1990) adopted an intermediate view, recognizing 132 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) five genera. Most of the paleotropical segregate genera (Holttum, 1969, 1972, 1973a, 1974, 1975, 1976a, 1976b, 1977, 1981; Holttum & Grimes, 1979; Iwatsuki, 1963), and several of the neotropical ones (Maxon & Morton, 1938; Smith, 1971, 1980), have been recently revised or monographed, making this one of the best known fern families morphologically, cytologically, and distributionally. Little, however, is known about relationships among these segregates. The goals of this work are to provide a phylogenetic hypothesis for the Thelypteridaceae based on molecular evidence from the 25-30 groups com- monly recognized within the family. A second goal is to confirm or refute the hypothesis of monophyly for the family. We make no attempt in this pa- per to incorporate a rigorous morphological data set to contrast with the mo- lecular data set, although we find it useful, and we hope informative, to comment on certain characters that are often used to distinguish genera or groups of genera, in light of the molecular results. In this paper, we choose to concentrate on the higher-level relationships in the family, and this initial approach necessarily involves only one or, in a few cases, two species per group (genus, subgenus, section). A study directed to the generic subdivision of the family would need to incorporate a minimum of two and ideally at least three species per group, so as to address the monophyly of the individ- ual genera, subgenera, or sections (depending on classification employed). MATERIALS AND METHODS TAXON SAMPLING.—As a basis for sampling, we used the classification of Holt- tum (1971), as modified by Holttum (1982) for Old World groups. For New World groups, we sampled from groups recognized by Morton (1963). A total of 30 ingroup taxa were sampled (Table 1). These represent 19 of the 22 Malesian genera recognized by Holttum (1982; only Ampelopteris, Amphi- neuron, and Cyclogramma are missing from our analysis), 25 of the 32 genera of Thelypteridaceae recognized by Pichi Sermolli (1977: only Ampelopteris, Amphineuron, Cyclogramma, Glaphyropteris, Haplodictyum, Menisorus, and Stegnogramma s.s. are missing), and 13 of the 20 Chinese genera recognized by Ching (1978; only Ampelopteris, Amphineuron, Cras- pedosorus (regarded by us as a synonym of Leptogramma), Cyclogramma, Mesopteris (regarded by us as a synonym of Sphaerostephanos), Stegno- gramma s.s., and Trichoneuron are missing). All neotropical groups, with the exception of Glaphyropteris s.s. (which we regard as a subgroup of Steiropteris; Smith, 1980), were also sampled. Recent phylogenetic analyses of the Polypodiales (higher leptosporangiate ferns) by Hasebe et al. (1995) and by Cranfill (unpubl. data) indicate that the thelypteroid ferns belong to this order, and that the most closely related groups (families in some classifi- cations) are an unresolved alliance of blechnoid, athyrioid, deparioid, ono- cleoid, and woodsioid ferns. Sixteen representatives from these families were used as outgroups. SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 133 DNA ExtRACTION, AMPLIFICATION, AND SEQUENCING.—We utilized silica gel-dried leaf material and, in a few cases, leaf material removed from herbarium specimens as sources for genomic DNA, which was extracted using DNEasy Plant Mini DNA extraction kits from Qiagen Corporation. Amplification was performed using Amplitaq Gold DNA taq polymerase produced by Perkin Elmer Corporation. Ongoing work by Cranfill previously demonstrated the phylogenetic utility of the chloroplast gene rps4 (Cranfill, 2000a, 2000b) and the trnS and trnL-F intergenic spacer regions (Cranfill, unpubl.), and so data from these three markers were collected for phylogenetic analysis. We also considered the inclusion of rbcL, but chose to exclude it from our study for reasons of time and economics. A preliminary analysis showed that rbcL data provided more or less the same phylogenetic information at roughly the same taxonomic levels as rps4. The rps4 region was amplified using the primers provided by Nadot et al. (1995), yielding an amplicon of approximately 650 bp. The intergenic spacer region between rps4 and trnS was amplified using the reverse compliment of Nadot et al. (1995), reverse rps4 primer R1 and the novel reverse primer trnS R (5’-TAC-CGA-GGG-TTC-GAA-TC-3’), yielding an amplicon of approxi- mately 400 bp. The intergenic spacer at the 3” end of trnL-F was amplified using primers e and f from Taberlet et al. (1991), yielding an amplicon of ap- proximately 350 bp. Amplification was accomplished using standard thermo- cycling protocols, with annealing temperatures of 48°C-54°C for rps4 and 45°C-48°C for the two intergenic spacer regions, using an MJ Research PTC- 200 Peltier thermal cycler. Direct sequencing of each PCR product was con- ducted in both directions for each marker using the BigDye cycle sequencing kits produced by Applied Biosystems Inc. (ABI) using an ABI PRISM 377 DNA automated sequencer in the Molecular Phylogenetics Laboratory of the University of California, Berkeley. DNA sequences were manually aligned. Alignment of the coding rps4 se- quences was unambiguous. The two intergenic spacer regions were aligned with more difficulty, and all ambiguously aligned regions were removed from the analyses. Gaps were treated as missing, while unambiguous and phylogenetically informative indels were treated as binary characters and coded at the end of the data matrix. Maximum parsimony analyses were performed with PAUP* (Swofford, 1999). Combinability of data from the three markers was investigated using the partition homogeneity test (Farris et al., 1995), as implemented in PAUP*, the results of which supported data combination. Heuristic searches were sultant trees were rooted with appropriate outgroups identified previously in the rbcL phylogeny by Hasebe et al. (1995), and confirmed by Cranfill from an unpublished five-gene molecular phylogeny of the Polypodiales. In addi- tion to computing a strict concensus of trees obtained from our searches, we TABLE 1. Sources of material of ingroup and outgroup species providing rps4, trnL spacer, and trnS a. sequences for this study. All seq data reported in this paper are newly generated. Parenthetical RBC numbers are DNA extraction num Species Locality Voucher/Source Herb. GenBank rps4-trnS' GenBank trnL spacer Amauropelta oe Costa Rica: San José: Las U. C. Bot. Gard. 57.0002 UC AF425162 AF425125 (Willd.) er ubes Chingia Jonglestinn Fiji: Viti Levu: Tomaniivi Game 99/270 (RBC 818) UC AF425163 AF425126 (Brack.) Holttum (Mt. Victoria) pei arida (D. Don) Malaysia Cranfill BF-22 (RBC 718) UC AF425164 ~ Holttu Christel hispidula cult., unknown source N. Y. Bot. Gard. 477/77A UC AF425165 AF425127 ne.) C. F. Reed = Cranfill s.n. (RBC 577) Christella (Pelazoneuron) _U.S.A.: Florida: Grey- U. C. Bot. Gard. 78.0268 UC AF425166 AF425128 augescens (Link) Pic nolds Park, S end West = Cranfill s.n. (RBC e, Miami 785) Coryphopteris seemannii Fiji: Viti Lev Tomaniivi Game 95/147B (RBC 817) UC AF427096/ AF425129 Holttum (Mt. Victori AF427097 Cyclosorus Agree cult. by B. Parris, New Cranfill s.n. (RBC 707) UC AF425167 AF425130 (Willd. Zealand Picton ae Taiwan Cranfill TW-004 (RBC UC AF425168 AF425131 6 clphyroptridops Taiwan Cranfill TW-197 (RBC UC AF425169 AF425132 erubescens (Hook.) 640) Ching Goniopteris Naa Mexico: Oaxaca: near Mickel 5799 = U. C. Bot NY AF425170 AF425133 (Bory) Chin xtepec Gard. 78.0266 (RBC oo tottoides Taiwan Cranfill TW-79 (RBC 638) UC AF425171 AF425134 Bers edie New Zealand Cranfill s.n. (RBC 673) UC AF425172 - torresiana re hing Meniscium sp. Panama: Canal Zone: A. Smith 2633 (from P. UC AF425173 AF425135 across street from Sum- mit Zoo Hammond) (RBC 786) (2002) Z MHEIWON 26 ANNTIOA “TYNYNOl NYAA NVOMANV TABLE 1. Continued. Species Locality Voucher/Source Herb. GenBank rps4-trnS' GenBank trnL spacer Mesophlebion Malaysia Cranfill BF-59 (RBC 703) UC AF425174 AF425136 —cesiaai (Blume) Mtatholyptrs dayi Malaysia Cranfill CH-35 (RBC 745) UC AF425175 AF425137 (Bedd.) Holttu Nannothelypteris Philippines: Mt. Makil- Price 670 (RBC 826) UC AF425176 - aoristisora (Harr.) ing, Laguna Holttum Oreopteris limbosperma Great Britain: South Crabbe et al. 11830 (RBC UC AF425177 = (All.) Holub Somerset, 4 km SW of Nether Stowey Parathelypteris U.S.A.: California: Plu- U. C. Bot. Gard. 60.0707 JEPS AF425178 AF425138 nevadensis (Baker mas Co., ca. 7 mi SE of (RBC 464) Holttum eddia Phegopteris connectilis cult., Mickel garden, New — Cranfill s.n. (RBC 575) UC AF425179 AF425139 (Michx t York Phegopt cult., Unknown source N. Y. Bot. Gard. 685/76 UC AF425180 - decursivepinat (H. C. = Cranfill s.n. (RBC Hall) F 576) hie cambio archboldiae Fiji: Viti Levu: Serua, A. C. Smith 9348 (RBC UG AF425181 - (Copel.) Holttum between Waininggere 829) and Waisese Creeks Pneumatopteris ecallosa Malaysia Cranfill CH-25 (RBC 639) UE AF425182 AF425140 (Holttum) Holttum Pronephrium simplex China: Hong Kong: U2, = Gard. 79.0293 UC AF425183 AF425141 (Hook.) Holttum Victoria Peak, (RBC 580) Victoria Isl. Pseudocyclosorus iwan Cranfill TW-29 (RBC 713) UC AF425184 AF425142 esquirolii (H. Christ) shing SdIHSNOILV' Tae TVITINVAVALLNI “THANVYO 8 HLIDNS SEL TABLE 1. Continued. Species Locality Voucher/Source Herb GenBank rps4-trnS' GenBank trnL spacer Pseudophegopteris aurita —_cult., Parris garden, New Cranfill s.n. (RBC 238) UC AF425185 - (Hook.) Ching Zealand, orig. from Ja- pan Sphaerostephanos Malaysia: Bangi Fern Cranfill BF-24 (RBC 641) UC AF425186 AF425143 penniger (Hook.) Garden Holttum Sphaerostephanos Taiwan: Taibei Botanical Cranfill TW-231 (RBC UC AF425187 - taiwanensis (C. Chr.) Garden 709) Holttum ex Kuo Steiropteris leprieurii French Guiana: Granville 13264 (RBC UC AF425188 ~ (Hook.) Pic. Serm. Montagnes de la Tri- 828) nité, Bassin de la Mana Thelypteris palustris N. Y. Bot. Gard. (wild) Cranfill s.n. (RBC 574) UC AF425189 AF425144 Schott Trigonospora ciliata cult., unknown source N. Y. Bot. Gard. acc. = UC AF425190 AF425145 (Benth.) Holttum Cranfill s.n. (RBC 582) Outgroups Acystopteris japonica Taiwan Cranfill s.n. (RBC 590) UC AF425150 AF425121 (Luerss.) Nakai Asplenium cristatum Costa Rica: Puntarenas: U. C. Bot. Gard. 90.2234 UC AF425146 - Las Alturas, 31 km = Cranfill s.n. (RBC from San Vito 042) Asplenium nidus L. cult. origin Cranfill s.n. (RBC 309) UC - AF425118 Athyrium filix-femina cult., Mickel garden, New — Cranfill s.n. (RBC 356) UC AF425152 (L.) Roth ex York Mert. s./. Cystopteris protrusa cult., Mickel garden, New — Cranfill s.n. (RBC 369) UC AF425148 AF425120 (Weath.) Blasdell York Deparia lancea (Thunb. cult., source unknown, U. C, Bot. Gard. 71.0038 UC AF425153 AF425123 ex Murray) Fraser-Jenk. native to SE Asia = Cranfill s.n. (RBC 1 Didymochlaena cult., unknown source Cranfill s.n. (RBC 432) UC AF425161 _ truncatula (Sw.) J. Sm. 9ET (Z00z) Z WAAIWAN 26 AWN'TOA “TVNUNOL NAA NVORIANV TABLE 1. Continued. Species Locality Voucher/Source Herb GenBank rps4-trnS! GenBank trnL spacer Gymnocarpium cult., Parris garden, New Cranfill s.n. (RBC 240) UC AF425149 - oyamense (Baker) Zealand, origin from in apan Homalosorus cult., Mickel garden, New Cranfill s.n. (RBC 597) UC AF425154 AF425124 pycnocarpos (Spreng.) York Pic. Serm. Hypodematium crenatum —_ Japan, sent by M. Kato Hyashi s.n. (RBC 768) TI AF425151 AF425122 (Forssk.) Kuhn Lorinseria areolata (L.) C. U.S.A.: S. Carolina: U. C. Bot. Gard. 82.2087 - AF425155 - Presl Orange Co., along = Cranfill s.n. (RBC Rte. 176 170) Matteuccia struthiopteris _—_cult., original source North Carolina Bot. Gard UC AF425158 - (L.) Tod. unknown = Cranfill s.n. (RBC 460) Onoclea sensibilis L. U.S.A.: Massachusetts Weed s.n Bot. UC AF425159 - Gard. 80.0321 (RBC 005 Onocleopsis hintonii F. Mexico: Oaxaca, sent by Mickel s.n. IND AF425160 - Ballard Gaston Sadleria cyatheoides U. S.A.: Hawaii: Maui: Ornduff 8506 = U. C. UC AF425156 - Kaulf. between Olinda and Bot. Gard. 78.0380 Wai Kamoi stream (RBC 003) Stenochlaena milnei Philippines (“New U.C. Bot. Gard. 55.0076 UC AF425157 - uinea”’ on label): Sa- (RBC 0.37) mar, from spores Woodsia polystichoides cult., Mickel garden, New — Cranfill s.n. (RBC 551) UC AF425147 AF425119 York ' rps4—trnS sequence data repre sent one contiguous sequence with rps4 at the 5’ end and the trnS spacer at the 3’ end. Sequences for Cory- phopteris are for rps4 alone (AF427096), trnS spacer alone (AF427097), and trnL alone (AF425129). SdIHSNOILV Tau TVITINVAVELLNI “THANVYD 8 H.LINS ZEl 138 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) also explored clade stability using bootstrap resampling (Felsenstein, 1985), as implemented in PAUP*. RESULTS support, including a clade comprising Dictyocline and Leptogramma (100%) and a clade comprising Goniopteris and Christella augescens (78%) (Fig. 1). Hasebe et al. (1995), who sampled four representatives in their global analy- sis: Amphineuron opulentum (Kaulf.) Holttum, Christella acuminata (Houtt.) Holttum, Parathelypteris beddomei (Baker) Ching, and Thelypteris palustris SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 139 PSEUDOCYCL ei ORUS 75 Sirintenbedtratnd P SPHAEROSTEPHANOS 56 cian CHRISTELLA 67 pepsin elaine be PLEISIONEURON PRONEPHRIUM C 78 CHRISTELLA AU 93 GONIOPTERIS MENISCIUM 89 AMAUROPELTA 98 PARATHELYPTERIS 69| 76 CORYPHOPTERIS ETATHELYPTERIS 68 HOMALOSORUS ce TIUM ae — ASPLENIUM 10 nious trees drawn as a phylogram, combined analysis utilizing acer, and frnL spacer. Branch lengths are proportional to the h. a me for clades, if specter than 50%, is indi- teps; CI = 5; RC = 0.63. See Table 1 for full names of species and additional voucher data. Fic. 1. One of 26 most parsimo the chloroplast genes rps4, trnS sp 140 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) Schott (listed here under their segregate generic names). The first two of these species belong to cyclosoroid genera, the last two species to amauro- peltoid genera or Thelypteris s.s. (Fig. 2). No representatives of the phegopte- roid clade have been included in previous analyses, and no other global or family level molecular surveys have provided any additional evidence on re- lationships either outside or within the family. Our study clearly shows that Thelypteridaceae, in the sense used by nearly all modern authors, is mono- phyletic, with high (98%) bootstrap support (Fig. 1). Holttum (1971) hypothesized a relationship between Thelypteridaceae and Cyatheaceae, a kinship that is now effectively refuted on the basis of ample morphological (Smith, 1990) and molecular evidence (Hasebe et al., 1995; Pryer et al., 1995; Wolf et al., 1999). Pichi Sermolli (1977) postulated a close relationship of Thelypteridaceae to Aspleniaceae, and indeed this relation- ship appears relatively close, although more remote than with several other large clades. Our analysis, and work by Cranfill (unpubl.), clearly show that the thelypteroid ferns are most closely related to a large terrestrial clade comprising the athyrioid, woodsioid, blechnoid, and onocleoid ferns (Fig. 1). The exact interrelationships of these immediate outgroups are still not well supported in our analysis or any other published molecular analysis. INTRAFAMILIAL RELATIONSHIPS.—Using morphological and cytological evidence, several students of the family have postulated intrafamiliar relationships within Thelypteridaceae, most notably Loyal (1963), Smith (1971), and Pichi Sermolli (1977). All of these studies invoked morphological and cytological evidence, and offered somewhat different and conflicting pictures of rela- tionships. Loyal (1963) depicted two main evolutionary lines within the fam- ily, one line with species groups having a chromosome base number of 36, with species having one or more pairs of veins from adjacent pinna segments united below the sinus. The second evolutionary line in Loyal’s scheme comprised species mostly with dysploid (stepwise progression in the basic chromosome set) chromosome numbers (27 to 35) and free veins. Smith’s (1971:46) scheme was an attempt to update Loyal’s tree, and showed basi- cally a free-veined evolutionary line, with the phegopteroid ferns terminat- ing this branch, and three shorter anastomosing- or connivent-veined lines having x=36. These included an Old World line, a New World line, and Stegnogramma s.]. Pichi Sermolli (1977:441) hypothesized a basal dichotomy in the family leading, along one line, to the evolution of several free-veined genera (phegopteroid ferns plus Metathelypteris), followed by other free-veined elements (Parathelypteris, Coryphopteris, Oreopteris, and Thelypteris), and ultimately giving rise to several free-veined and connivent-veined New World species groups (Amauropelta, Steiropteris, and Glaphyropteris). As a second evolutionary line, Pichi Sermolli derived all of the anastomosing- veined Old World genera. Although Holttum (1971, 1982) revised nearly all of the Old World genera, and commented extensively on relationships, he never presented a formal phylogenetic treatment. Holttum (1971) believed SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 141 that certain Old World cyclosoroid genera (Pronephrium, Nannothelypteris, Stegnogramma, Sphaerostephanos, and Pneumatopteris) formed a natural group. He also postulated that Mesophlebion, Chingia, and Glaphyropteri- dopsis were related, and that Coryphopteris and Parathelypteris formed an isolated group. According to Holttum (1982:386), Cyclosorus s.s. and Ampel- opteris (with united veins) and Thelypteris s.s.(with free veins), all with wide distribution and similar aquatic habitat, formed another closely related group. One of Holttum’s most important contributions was the recognition that the three phegopteroid genera formed a related group (Holttum, 1969, 1971). Although elements of all of these precladistic and largely intuitive trees have some credence, the molecular data provide a unique new hypothesis of relationships. The phegopteroid genera (Phegopteris, Macrothelypteris, and Pseudophegopteris) appear to be the sister group to all other taxa in the fam- ily (Figs. 1, 2). Within the phegopteroid clade, Macrothelypteris appears as the sister group to Pseudophegopteris and the two Phegopteris species. All nodes within the phegopteroid clade have high bootstrap support (Fig. 1). Following the origin of Thelypteris s.s., which is shown to be basal in the thelypteroid—-cyclosoroid clade, the free-veined thelypteroids arise, with greater or lesser bootstrap support for various nodes, and no significant sup- port at all for some of the nodes at the crown of one of the main branches (Fig. 1). Thelypteris s.s. is shown to be only distantly related to Cyclosorus s.s., thus refuting the close relationship suggested by Holttum (1982:386). The cyclosoroids, from Steiropteris on up, form a large, mostly unresolved or poorly resolved clade toward the tip of the tree (Figs. 1, 2). This topology supports the contention of Pichi Sermolli that, in general, those elements in the family with anastomosing veins and x= 36 are evolutionarily derived in the family. In the tree shown, Dictyocline and Leptogramma appear as sister genera, a result that supports all recent classifications of the family. Most authors, including Iwatsuki (1963), Holttum (1982), and Smith (1990) unite these two segregates in the same genus or subgenus. The lack of resolution of various genera within the cyclosoroid clade (Figs. 1, 2) may be an artifact of inadequate sampling or an indication of in- sufficient variation in the genes sampled; more likely, it also reflects weak distinctions between and among genera within this clade. Several hybrids are known between species in different cyclosoroid genera (Viane, 1985; Quansah & Edwards, 1986; Sledge, 1981; J. Game, unpubl. data). Many of the cyclosoroid generic segregates seem little more than one- or few-character genera, often containing exceptional species that do not fully conform in their diagnostic features (Smith, 1990). We note in this regard that the three species of Christella sampled in our study do not form a monophyletic group, and in fact the single New World species sampled, Christella auges- cens (Link) Pic. Serm., is well separated from the other two species. This ecies (Smith, 1971; Smith, 1990) has sometimes been treated in a different subgenus, subg. Pelazoneuron, in con- trast to the largely Asian subg. Christella. 142 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) CHARACTER EvoLuTION.—Below we discuss briefly several characters often ap- plied in circumscription of groups (genera, subgenera, sections) of Thelypter- idaceae, expecially those characters that appear to correlate to a significant extent with the results presented here. Venation.—All students of the family have recognized the importance of venation in the classification and subdivision of the thelypteroid ferns and have utilized that character, whether free, connivent at the sinus, or anasto- mosing in various ways, as a primary key character and feature delimiting genera or infrageneric taxa within the family (e.g., Christensen, 1907, 1913, 1920; Holttum, 1971, 1982; Tryon & Tryon, 1982:432—453; Smith, 1990). Our results show that the basalmost clades in the family, from the phegopteroid genera through Amauropelta and Parathelypteris (Fig. 3) are all free-veined or at least have veins that run to the sinus (but do not anastomose). In con- trast, most of the groups belonging to the cyclosoroid clade (Fig. 3, shaded boxes) have veins that are either connivent at the sinus, or anastomose at an acute angle below the sinus, or unite at an oblique angle below the sinus with an excurrent vein to the sinus. In some groups, there are multiple series of anastomoses between the costa and pinna margin, as in the neotropical groups Meniscium and Goniopteris, or in the paleotropical groups Prone- phrium, Sphaerostephanos, and Dictyocline. Vein fusion is known to have arisen independently in many leptosporangiate fern lineages, and no doubt has also evolved many times within Thelypteridaceae. Smith (1990:271) pointed out that half of the 20 subgenera of Cyclosorus s.]. have both free- veined and anastomosing-veined members. The evolutionary trend toward increased vein anastomosing, however, is unmistakable, and lends support to the overall topology of the tree. Many of the outgroups we used in this study are also free-veined, but certain ones, such as Onoclea and Lorinseria, also have evolved anastomosing venation. Anastomosing or reticulate venation in Thelypteridaceae is generally of a simple or very repetitive nature: at most there are simple (unforked), excur- rent free veins within areoles, as in Meniscium. This pattern of vein reticula- tion, termed ‘“‘intersegmental” by Wagner (1979), is so common in man thelypteroid groups that Wagner termed the venation pattern “‘thelypteroid”’; if the pattern is repeated several times, Wagner (and others) have termed it “meniscioid” (after the genus Meniscium) or “goniophlebioid” (similar to a venation pattern in the genus Goniophlebium, a member of the Polypodia- ceae). The various venation types in the family have been discussed and il- forked veinlets in the areoles. Adaxial sulcation.—Holttum (1960) was one of the first to discuss in depth the importance of sulcation, or adaxial grooving, as applied to the system- atics of ferns. He soon recognized that all three phegopteroid genera, as well SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 143 2 PSEUDOCYCLOSORUS AISUNOBO1DAD AA TAHL AION LAA ISHS. x=34 ACN. AOOsHa rooted on Asplenium, and Asplenium through Cystopteris represent outgroups used in the anal- ysis. Familial and subfamilial groups are indicated at the right. See Table 1 for full names of species and additional voucher data. Fic. 3. Strict concensus tree of the 28 most parsimonious trees found in the combined parsi- mony analysis, based on the chloroplast genes rps4, trnS spacer, and trnL spacer. Thelypteroid taxa having connivent or anastomosing veins are indicated by shaded boxes; chromosome base numbers for various groups are shown along the right. See Table 1 for full names of species and additional voucher data. as Metathelypteris, lacked adaxial grooves along the costae (Holttum, 1969, 1982). This contrasts with all other thelypteroid ferns, which have the costae adaxially sulcate. As Holttum (1960) noted, some genera of higher leptospor- angiate ferns (e.g., Dryopteris and Polystichum) have the grooves of one axis continuous with the grooves of the axes of greater (and often lesser) orders, but continuous grooving is unknown in Thelypteridaceae. The grooving of many of the outgroups used in this study has not been sufficiently studied, but most appear to be adaxially grooved, including both Cystopteris and Gymnocarpium, the apparent closest relatives. . Blade dissection—With rare exceptions, blades of all cyclosoroid and thelypteroid genera in Figure 2 are pinnate-pinnatifid or less divided. —. few species that have more divided fronds (e.g., Thelypteris (Amaurope ta) pteroidea (Klotzsch) R. M. Tryon) are clearly derived, and barely 2-pinnate. A few members of some cyclosoroid clades have simple or merely pinnatifid 144 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) blades, e.g., certain species of Meniscium, Goniopteris, Pronephrium, and Sphaerostephanos. In contrast, the phegopteroid genera have blades that are bipinnate-pinnatifid (or even more divided), or the blades are bipinnatifid or tripinnatifid (Phegopteris). Dimorphism.—Frond (blade) dimorphism is nowhere strongly pronounced in the family, as it is in many other fern families, but does occur on a subtle to moderate scale (subdimorphism) in many (but not all) members of the cy- closoroid clade, e.g., Christella, Goniopteris, Meniscium, Pronephrium, and Sphaerostephanos. Dimorphism is nearly totally lacking among members of the phegopteroid clade, as well as most members of the thelypteroid clade; exceptions to this are found in the subdimorphic group Thelypteris s.s., and subtly in Coryphopteris and Parathelypteris. There appears to be a positive correlation between degree of vein anastomosing and blade dimorphy in the family, i.e., those groups that have significant vein reticulation also are more likely to be subdimorphic. Hemidimorphism (fertile-sterile differentiation on separate parts of the same leaf; Wagner, 1977) is essentially unknown in Thelypteridaceae. As pointed out by Wagner (1979), dimorphy appears to be a character that is valuable mainly at the species level, not at generic rank, and the evolution of dimorphy (or subdimorphy) has occurred indepen- dently many times in the ferns and in the Thelypteridaceae. It is doubtful that this character is of any help in delimiting clades within this family. Among outgroups used in this study, most have monomorphic fronds, but members of the Blechnaceae (e.g., Lorinseria and Stenochlaena) are often, but not always (e.g., Doodia, Sadleria, a few Blechnum species), strongly dimorphic, as are members of the onocleoid ferns (Onoclea and Matteuccia in our sample). Spore morphology.—Spores of thelypteroid ferns are generally monolete and kidney-bean-shaped, with a relatively thick, folded, cristate, reticulate, or echinate perispore. Spore surveys of thelypteroid genera suggest that spore ornamentation is, in some cases, correlated with the segregate taxonomy in Amauropelta. Species of Stegnogramma, in the cyclosoroid alliance, have distinctive echinate spores, but these are not unlike many species of Sphaer- pteris and Meniscium, as well as in the paleotropical group Mesophlebion (Tryon & Lugardon, 1991:401), and these similarities (as well as others) sug- gest a possible close relationship. Spores of Trigonospora are trilete, the only SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 145 such spores in the family; this is unquestionably a derived condition in Thelypteridaceae, and not likely an indication of relationship to the Jurassic fossil Aspidistes, as suggested by Holttum (1982) and others. No doubt, there is much more evidence of a taxonomic nature to be gleaned from a thorough and comparative study of spore morphologies in the family. Biogeography.—Most of the segregate genera recognized in thelypteroid ferns are restricted either to the New World or to the Old World. In the tropics, exceptions are Cyclosorus s.s., Christella (predominantly Old World) and Amauropelta (predominantly New World). In north-temperate areas, Oreopteris, Thelypteris, and Parathelypteris are amphi-oceanic, occurring in both North America and Eurasia. The cyclosoroid clade (Fig. 2) is entirely tropical or subtropical in its cur- rent distribution; few members extend above 25° north latitude. Contrast- ingly, of the basal thelypteroid segregates, Thelypteris s.s. (except the segregate T. confluens (Thunb.) C. V. Morton also occurs in south temperate regions), Oreopteris, and Parathelypteris are mostly found north of 25° north latitude, and Metathelypteris also has many temperate representatives. Amauropelta and Coryphopteris are the only largely tropical segregates. In the phegopteroid alliance, Phegopteris itself is north-temperate in distribu- tion, in both eastern Asia and North America, while Macrothelypteris and Pseudophegopteris are restricted essentially to the Old World tropics and subtropics (M. torresiana (Gaudich.) Ching is widely naturalized in the New World). Several outgroups sampled are cosmopolitan (e.g., Asplenium, Cys- topteris, Woodsia); others are more restricted (e.g., Lorinseria and Sadleria). We note that the phegopteroid clade and also the basal clades of the the- lypteroid lineage are, for the most part, Laurasian in distribution. This distri- bution perhaps indicates an east Asian origin, north of the Tethys sea, for the family. Although both Pseudophegopteris and Macrothelypteris are con- fined to the Old World (in their native distribution), the other phegopteroid and most of the basal thelypteroid genera are found in both the Old World and New World, and most are decidedly northern in their general distribu- tion. Only higher in the thelypteroid lineage, in particular the cyclosoroid nearly confined to the Cen the historic inclusion of thelypteroid ferns in the distantly related dryopte- roid assemblage. The best known and earliest examples of thelypteroid fos- sils appear to be from Eocene strata of North America, and Europe (Collinson, 2001:209-212). Further study and re-evaluation of identifications in light of modern systematic knowledge of extant members is needed to de- termine whether the fossil evidence adds support to a Laurasian origin for the family. Chromosome numbers.— are characterized by having a tatives of all of the cyclosoroid genera s on x = 36, and many species in most 0 Most of the segregate genera of Thelypteridaceae single base chromosome number. All represen- ampled (Fig. 3) are known to be based f these segregates have been sampled. 146 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) Chromosome base numbers are known for all groups except Nannothelyp- teris, which has been combined with Pronephrium sect. Dimorphopteris by Smith (1990) as suggested by Holttum (1982:538). Of the phegopteroid gen- era (Fig. 3), Phegopteris is consistently x = 30, with all three species counted; Pseudophegopteris is x = 31, with eight species counted; and Macrothelypteris is also x = 31, with four species counted. The situation in the free-veined thelypteroid segregates is more complex. There appears to be a dysploid series of numbers that characterizes the various groups (Fig. 3: Lovis 1977). Thelypteris s.s. is consistently x = 35; Oreopteris is x = 34 (all three species counted); Coryphopteris species are x = 32 and 33 (Smith, unpubl.), but with few counts available; reports for Metathelypteris are mostly x = 31, 34, 35, and 36 (five species counted); Parathelypteris is variously x = 27, 31, 32, 34, with the first two numbers being most commonly reported; and Amauropelta is consistently x = 29, with ca. 25 species counted. Whether the multiple numbers for some of the segregates indicate unnaturalness in the existing classifications or accurately reflect chromosomal variation among (and even within) closely related species remains to be determined. It does appear, however, that there is some chromosomal instability within the basal thelypteroid group of genera. SUMMARY AND CONCLUSIONS Although taxon sampling is still only preliminary, we conclude that the family Thelypteridaceae is monophyletic, with a high degree of certainty. It excludes a few genera that have sometimes been included in the family by Ching (e.g., Hypodematium; Ching, 1963; Trichoneuron; Ching, 1978) but ex- actly coincides with the circumscription given by Iwatsuki (1964), Holttum (1971), Pichi Sermolli (1977), and Smith (1990). Although our sampling is not yet sufficient to favor one of the many Classifications (Morton, 1963; Iwatsuki, 1964; Holttum, 1971, 1982; Ching, 1963, 1978; Pichi Sermolli, 1977; Smith (1990) of the family over another, our analysis suggests that rec- ognition of an intermediate number of genera may be a reasonable taxonomic course. The three phegopteroid genera appear to be basal in the family and clearly sister to the remaining genera and subgenera. ACKNOWLEDGMENTS ial of Cyclosorus, material which was to serve as SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 147 We thank the following individuals for help in procurring either living or silica-gel-dried materi- al for sequencing: John Game, Philip Hammond, Barbara Joe Hoshizaki, Masahiro Kato, John T Mickel, and Barbara Parris. Gerald Gastony kindly provided DNA sample of Onocleopsis. We also thank individuals and staff at the New York Botanical Garden and the University of Califor- collection data, and providing for the collection of vouchers. Thanks also to numerous individu- als who provided field assistance in Malaysia and Taiwan, where Cranfill collected many of the Asian vouchers and DNA material cited in Table 1. The National Science Foundation provided support to both Smith (DEB-9616260) and Cranfill (DEB- 0073036). LITERATURE CITED CHING, : C. 1940. On natural classification of the family ‘‘Polypodiaceae”. Sunyatsenia 5: pre A coasted of the family Thelypteridaceae from the mainland of Asia. Acta Phytotax. Sin. 8:289— . 1978. The Chinese gel families and genera: ‘e paeecr arrangement and historical ori- Acta Phytotax. Sin. 16(3):1-19 and 16(4):16— Cont C. 1907. Revision of the American species - Dryopteris of oe group a 5 opposita. el. Danske Vidensk. Selsk. Skr., Naturvidensk. Math. Afd., ser. 7, 4:247— ——. aats, A monograph of the genus Dryopteris. Part. I. The rae pone pinata bipinnatifid species. Kongel. Danske Vidensk. Selsk. eg Naturvidensk. Math. Afd., ser. 7, . 1920. A enol of the genus Dryopteris. Part. II. The tropical American bipinnate- ap compound species. Kongel. Danske Vidensk. Selsk. Skr., Naturvidensk. Math. Afd., ser. 8, 6:1-13 COLLINSON, “4 E. 2001. Cainozoic ferns and their distribution. Brittonia 53:173-235. CranFiLL, R. 2000a. Has ecological > aemimaignn driven a fundamental phylogenetic split in the Polypodiales? Amer. J. Bot. 87(6, Suppl.):9 . 2000. Phylogenetic utility of plastid acta protein S4 (rps4) in land plants. Amer. J. Bot. 87(6, Suppl.):121. sear a S., M. KALLeRsJO, A. KLUGE, — C. Butt, 1995. Constructing a significance test for in- ongruence. Syst. Biol. 44:570-57 PaLsenstet, J. 1985. Confidence aes on n phylogenies: An approach using the bootstrap. Evolu- 783-791 Hasese, M., P. G. Siem K. M. Pryer, K. UEDA, — R. Sano, G. J. GasTONY, : YOKOYAMA, J. R. MANHART, N. MURAKAMI, Ec H: sunt _H. Haurier, and W. D. Haux. 1995. Fern phylogeny based on rbcL nucleotide sequences. Amer. Fern J. 85:134—181. Houtrum, R. E. 1960. Vegetative characters ; eal the various groups of ferns included in soars in — s ai Filicum, and other ferns of similar habit and sori. 9. Studies in the se vaiskepisadlioee The genera Phegopteris, Pseudophegop- aia por Macrothelypteris. Blumea 17:5—32. . 1971. Studies in = family Thelypteridaceae III. A new system of genera in the Old World. Blumea 19: . 1972. Studies in ie fa 20:105—126 . 1973. Studies in the family Blumea 21:293-325. . 1973b. The family Thelypteri Crabbe & B. A. Thomas, eds. Press, London . 1974. Studies in the family mily Thelypteridaceae IV. The genus Pronephrium Presl. Blumea Thelypteridaceae V. The genus Pneumatopteris Nakai. daceae in the Old World. Pp. 173-189 in A. C. Jerm The Phylogeny and Classification of the Ferns. Arcane Thelypteridaceae VII. The genus Chingia. Kalikasan 3:13-28. 148 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) - 1975. Studies in the family Thelypteridaceae VIII. The genera Mesophlebion and Plesio- neuron. Blumea 22:223-250. 1976a. bein in the family Thelypteridaceae X. The genus Coryphopteris. . 197 eb. satis in ve arene Thelypteridaceae XI. The genus Christella Léville, sect. agora Kew Bull. 31: —.. 7. Studies in Pet family Thelypteridaceae XII. The genus Amphineuron shee Blumea 23:205-218. - 1981. The genus Oreopteris (Thelypteridaceae). Kew Bull. 36:223-226. . 1982. oo Flora Malesiana, ser. II. Pteridophyta 1(5):331-560. Martinus Nijhoff, The Hague —, = J. W. Grimes, 1979. The genus Pseudocyclosorus Ching (Thelypteridaceae). Kew Bull. 499-516. Iwatsuxi, K. 1962. On the venation of the cana ferns. Acta Phytotax. Geobot. 20:219-227, ——. 1963. Taxonomic studies of Pteridophyta VII. A revision of the genus Stegnogramma emend. Acta Phytotax. Geobot. 19:112-126. . 1964. Taxonomy of co thelypteroid ferns, with special reference to the species of Japan and adjacent regions. III. Classification. Mem. Coll. Sci. Kyoto Imp. Univ., ser. B, Biol.: 31:11—40. Lovis, ie D. 1977. Evolutionary patterns and processes in ferns. Pp. 239-440 in R. D. Preston & H. W. Woolhouse, eds. Advances in Botanical Research. Academic Press, London LoyaL, D. S. 1963. Some evolutionary trends in family Thelypteridaceae with partionlar refer- ence to lene yee species. Mem. Indian Bot. Soc. 4:22—29, Maxon, W. R. and C. V. Morton. 1938. The American species of Dryopteris, subgenus Menis- cium. Bull. Tetrey Bot. Club 65:347-376. Morton, C. V. 1963. The classification of Thelypteris. Amer. Fern J. 53:149-154. Napot, S., G. Birrar, L. Carrer, R. LaCroix, and B. LEJEUNE, 1995. A phylogenetic analysis of monocotyledons based on the ‘deraplaat gene hed using parsimony and a new numerical henetics method. Molec. Phylogen. Evol. 4:25 PICHI SERMOLLI, R. E. G. 1977. Tentamen Se genera in taxonomicum ordinem redi- gendi. Webbia 31:313-512 Pryer, K. M., A. R. Soir, an dj. E. Skoc. 1995. Phylogenetic yeaa of extant ferns based on evidence from morphology and rbcl. sequences. Amer. Fern J. 85:205—282 Quansak, N. and D. S. Epwarps. 1986. A natural bigeneric fern hybrid in Thelypteridaceae from Ghana. Kew Bull. 41:805-809. SHING, K. 1999. Flora Reipublicae Popularis Sinicae, vol. 4(1). Pteridophyta: Hypodematiaceae, Thelypteridaceae. Science Press. SLEDGE, W. A apo The Thelypteridaceae of Ceylon. Bull. Brit. Mus. (Nat. Hist.), Bot. 8:1-54 SmitH, A. R. 19 = ess “ si neotropical species of Thelypteris section Cyclosorus. Univ. Calif. pre Bot. - 1980. Taxonomy of a spters os Steiropteris, including Glaphyropteris (Pterido- phyta). Univ. Calif. Publ. Bot. . 1990. ed econ Pp. pores n K. U. Kramer & P. S. Green, vol. eds., The Fami- of Vascular Plants. ar" I. Pteridophytes and Gymnosperms. Springer- Verlag, hatin Sworrorp, D. L. 1999. Phylogenetic analysis using parsimony (*and other methods). Version 4, eta 8. Sinauer Associates, Sunderland, MA - GIELLY, ou, and J. Bouver. 1991. Universal primers for amplification of three non-coding regions of chloroplast DNA. Pl. Molec, Biol. 17:1105—1109. ‘ B. LuGAaRDON. 1991. Spores of the Pteridophyta. Springer- -Verlag, New Y and A. F. Tryon. 1982. Ferns and Allied Plants with Special Reference to Tropical America. pgelleane: New York. ViANE, R. L. L. 1985. A n W species and a new Sa hag of Thelypteris (Pteridophyta) from the Ivory Coast. Bull. hog Roy. Bot. Belg. 118:4 SMITH & CRANFILL: INTRAFAMILIAL RELATIONSHIPS 149 WAGNER, us - o 1977. Fertile-sterile leaf dimorphy in ferns. Gard. Bull. Straits Settlem. 30:2 os Retculate veins in the systematics of modern ferns. Taxon 28:87-95. WoLr, P. G., S. D. Swes, M. R. Wuite, M. L. Martines, K. M. Pryer, A. R. SmirH, and K. UEDA 1999. Baraat ee of the enigmatic fern families Hymenophyllopsidaceae and ears: Epa vidence from rbcL nucleotide sequences. PI. Syst. Evol. 219:263-270. Woop, C. C. 1973. Spore frneren in the Thelypteridaceae. Pp. 191-202 in A. C. Jermy, J. A. Crabbe & : A. Thomas, eds. The Phylogeny and Classification of be Ferns. hectic Press, London. American Fern Journal 92(2):150-160 (2002) Two New Species of Moonworts (Botrychium subg. Botrychium) from Alaska Mary Ciay STENSVOLD USDA Forest Service, Alaska Region, 204 Siginaka Way, Sitka AK 99835 DONALD R. FARRAR Department of Botany, Iowa State University, Ames, IA 50011 CINDY JOHNSON-GROH Department of Biology, Gustavus Adolphus College, St. Peter, MN 56082 AssTRact.—Botrychium tunux and Botrychium yaaxudakeit, new species of moonworts currently known only from southern Alaska, are described and illustrated. These ferns are distinguished B. lunaria, with which they have been confused, by allozyme data and their morphological characteristics. Ploidy levels of B. tunux (diploid) and B. yaaxudakeit (tetraploid) are inferred from allozyme patterns. A key to Alaskan moonworts is presented. Before 1995, little attention was paid to Alaskan moonworts, Botrychium subg. Botrychium. Information gained from rare plant surveys since then has improved our knowledge of moonwort abundance, distribution, and relation- ships. Seven species of moonworts are now known in southern Alaska: Botrychium ascendens W.H. Wagner, B. lanceolatum (S.G. Gmel.) Angstrém, B. lunaria (L.) Sw., B. minganense Vict., B. pinnatum H. St. John, the new species B. alaskense W.H. Wagner & J. R. Grant, and two additional new species described here, B. tunux and B. yaaxudakeit. During a 1980 USDA Forest Service rare plant survey, several Botrychium identified as B. lunaria (M. C. Muller 3806, ALA, TNFS) were collected in sandy upper beach meadows near Yakutat, in southeastern Alaska. In 1986, . H. Wagner and F. S. Wagner (1986) described a new moonwort, Botrychium ascendens, and re-identified some of the 1980 specimens as B. ascendens; other specimens on the herbarium sheets remained B. Junaria. The inclusion of Alaska in the range of B. ascendens in Volume 2 of the Flora of North America (W. H. Wagner, 1993) was based on the 1980 Muller collection. STENSVOLD ET AL: TWO NEW SPECIES OF MOONWORTS 151 Farrar’s analysis identified B. minganense, B. lunaria, and two moonworts previously unknown to science. These new moonworts, a diploid and an allotetraploid, resemble B. Junaria most closely in morphology, but are distinct genetically and morphologically from all other known moonworts. To our knowledge the diploid has not been previously collected. Previous collections of the tetraploid had been included in B. Junaria. Genetic comparison of B. tunux with B. lunaria at 18 allozyme loci yielded a value for Nei’s (1978) genetic identity (GI) of 0.5102 (Farrar, un- publ. data). This represents a level of genetic similarity only slightly greater than that between B. Junaria and B. simplex (GI=0.4646) and is significantly less than the similarity between B. Junaria and its sister taxon, B. crenula- tum (GI=0.7015). Absence of fixed heterozygosity at any enzyme locus implies that plants of B. tunux are diploid. Presence of fixed heterozygos- ity and a spore size significantly larger than that of B. Junaria implies that plants of B. yaaxudakeit are tetraploid. Allozyme patterns also show that one ancestral diploid parent of B. yaaxudakeit is North American B. lunaria; the other ancestral diploid is probably not among the known North Ameri- can diploid species. Preliminary investigation of European B. Junaria indi- cates that B. yaaxudakeit may be the result of ancient hybridization between genetically distinct North American and Eurasian genotypes of B. Junaria. Key TO THE SPECIES OF Botrychium Susc. Botrychium OF SOUTHERN ALASKA iy rhs on ca cans hk es ae he ER EE PPS Oe OERET CED RY Has os 3(2). Pinna bases cuneate; pinna apices acute .....---+++++- see rr rrr rt eects 3. Pinna bases obtuse to truncate; pinna apices rounded : . pinnatum 4(1). Basal pinnae narrowly fan-shaped to oblong with the blade spanning an arc less than 90°; pinnae remo 4. Basal pinnae broadly fan FS ko oisclind scl css SEK Gilg Te OUR OAT LE ES eh iS entel an veie tes -shaped with the blade spanning an arc greater than 90°; pin- n ere Se. er Seat Oh OF eee MAE Fe 6 5(4). Trophophores sessile; pinnae strongly ascending, their margins dentate to lacerate; spor- TTC sc scans soe oe es ascendens cending, oblong to narrowly fan-shaped, lly lacking on the trophophore pinnae by rilece alegre eeuaeanrs B. minganense 6(4). Basal pinnae asymmetrical with the basal portion expanded; mature sporophore stalks shorter than or equal to the trophophores; basal sporophore branches spread- Seip and twined: plants Q-idiene Bill. ove «1s onrss nits dats seee es TS Ree 6. Basal pinnae generally symmetrical; mature sporophore stalks longer than the trophophores; basal sporophore branches ascending and straight; plants 8-25 cm i Bey en a ee ee ee a falls 2 Ree ade Ginna Sees 7(6). Pinnae strongly overlapping one another an greater than 180°; basiscopic inner margins 0 45 (43-48) pm in longest diameter ...----+-+-++errrrrrrss d the rachis; basal pinnae spanning an arc f the basal pinnae strongly recurved; spores B. yaaxudakeit 152 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 2 (2002) 7. Pinnae approximate to somewhat overlapping, not overlapping the rachis; basal pinnae spanning an arc of less than 180°; basiscopic inner margins of the basal pinnae linear to moderately concave; spores 36 (34-39) um in longest diameter ............. B. lunaria Botrychium tunux Stensvold & Farrar, sp. nov.—Type: U.S.A.: Alaska: Yakutat area, ca. 5 km W of Yakutat, on the Phipps Peninsula ca. 1.5 km SE of Point Carrew, just N of the mouth of the Ankau River, ca. 1.5 m elev., 25 Jul 2001, M. C. Stensvold & D. R. Farrar 7936 (holotype ISC; isotypes ALA, NY, TNFS, US, WTU). Figs. 1c, 2c, 3b, 3c. Plantae supraterraneae 6-12 (ave. 9) cm altae; stipites communes 0-3 (ave. 1.5) cm longi. Trophophora flavovirentes coriaceae oblongae. Paria pinnarum plus minusve perpendicularia ad rhachim, separata vel leviter imbricata; pinnae basales 7-20 (ave. 13) mm longae, 7-18 (ave. 12) mm latae, sessiles, flabellatae, angulo ad basim 120-180°, saepe asymmetricae, parte basali expanso, margine externa integra vel interdum lobata, sinibus rotundatis. Stipites sporophora 2.5-5 (ave. 4) cm longi, trophophoria aequantes vel breviores. Sporae 38—42 jum diametro. 12 (ave. 9) cm tall with a common stalk 0-3 (ave. 1.5) cm long. Tropho- phores yellow-green, leathery; stalks 0-1 cm long; blades 2.5-7 (ave. 4) cm long, 2-4 (ave. 3) cm wide, narrowly ovate to ovate, once pinnate. Pinna pairs 4-6, more or less perpendicular to the rachis, separated to slightly overlapping, not overlapping the rachis. Basal pinnae 7-20 (ave. 13) mm long, 7-18 (ave. 12) mm wide, sessile, fan-shaped spanning an arc of 120— 180°, often asymmetrical, with the basal portion expanded; basiscopic inner major veins entering the pinna base, 45-55 veins ending at the margins. Sporophores 5-10 (ave. 7) cm long; sporophore stalks 2.5-5 (ave. 4) cm long and shorter than or equaling the length of the trophophore; sporangia- bearing portion erect, 1—2-pinnate, broadly ovate in outline, branches 4-6, ascending to spreading, especially the lowermost, which are often twisted such that the sporangia project downward; sporangia partially embedded in the distally thickened sporangiophore branches. Spores 38—42 (ave. 40) tm in longest diameter, released later than those of B. lunaria. Apparently diploid. This species is morphologically most similar to B. lunaria, from which it can be distinguished by its short stature, short common stalk, frequently stalked, ovate trophophore, asymmetrical pinnae with their basiscopic side expanded, entire margins commonly cleft by shallow incisions with rounded sinuses, and sporophore stalks seldom exceeding the height of the tropho- . > generally short-stalked or unstalked trophophore. Its fan-shaped pinnae are 153 STENSVOLD ET AL: TWO NEW SPECIES OF MOONWORTS Ayal b yy beh 7 variable Athyrium angustum and , id. A. Array of spore B. Close-up of ‘“‘fla Fic. 4. Spores of putative hybrids between SEM. A-G, spores of putative first-generation hybr lote morphology and ap exomean of “flaky’’ intermediates. ky” spore morphotype. C-G. Examples . spore morphotypes occurring in putative F-1 hybrid, ing from A mecaestics tee: ‘e to; = oo s-like (G). H— spores of putative backcross rt stween Fs hybrid and A. asple ots 2s. H. Array of spore bie Sa exhibited by the puta- ve backcross. Note prevalent of A. asple aaa s-like spores. I. Close-up of A. asplenioides-like spore from backcross. Note irresolute nature of inflated folds and iach ncy tomeed ‘flakiness’ spore morphotypes putative backcross. Note variation from ) A. asplenioides- under intermediate vary- J-M. Examples of occurring in “flaky” intermediate (J) tc ike (M). june KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 197 Mdh-3, Mdh-4, Pgi-2, Pgm-2, 6-Pgd-1, 6Pgd-2, Skdh, Tpi-1, and Tpi-2) were variable in at least one population, and the frequencies of the alleles were computed (Table 3). There was a strong tendency for all populations to share the more common alleles and some of the less frequent ones as well. Most of the loci were weakly polymorphic, a single allele predominating in all popu- lations of both A. angustum and A. asplenioides. In contrast, at the four most polymorphic loci (Idh-1, Pgi-2, Pgm-2, and Tpi-2), population allele fre- quencies were similar within A. angustum and A. asplenioides respectively, but strikingly different between the two taxa (Table 3). The most contrasting locus was Idh-1 for which alleles Idh-1° and Idh-1° prevailed in A. angustum at frequencies ranging from 0.545 to 0.676 and 0.296 to 0.455 respectively, while allele Idh-1“ was the most frequent in A. asplenioides with frequencies ranging from 0.828 to 0.950. At Pgi-2 both taxa shared Pgi-2" as their most frequent allele; however, Pgi-2° was repre- sented in all A. angustum populations at much higher frequencies, ranging from 0.154 to 0.450, than in populations of A. asplenioides for which the fre- quency of this allele ranged from 0.000 to 0.106. At Pgm-2, both taxa shared Pgm-2° and Pgm-2° as principal alleles, the former prevailing in A. angus- tum populations at frequencies ranging from 0.856 to 1.000, the latter pre- vailing in A. asplenioides populations at frequencies ranging from 0.580 to 0.857. At Tpi-2 both taxa shared three prevalent alleles, with Tpi-2* being of substantial frequency in all populations. Allele Tpi-2° occurred at high fre- quencies, ranging from 0.355 to 0. 550, in A. angustum populations whereas pi-2° was infrequent in this taxon, occurring at frequencies from 0.000 to 0.100. Conversely, Tpi-2° often was the most frequent allele in A. asple- niordes populations, occurring at frequencies from 0.316 to 0.606, while Tpi- 2” was rarer, occurring at frequencies from 0.000 to 0.255. The contrasting allele frequency trends for these four loci were highly consistent among pop- ulations within each taxon. Exceptions to these frequencies trends were in the northernmost sampled A. asplenioides population, Shirley, from south- ern New Jersey for the characteristically A. angustum alleles Idh-1" and Pgi- 2° and likewise for Tpi-2” in the highest elevation (1300 m) population of A. asplenioides sampled, Pond Drain, from the mountains of southwestern Virginia. Moreover, five of the six Shirley, NJ, individuals hypothesized to be first-generation hybrids on the basis of spore morphology were hetero- zygous for A. angustum and A. asplenioides 1 marker alleles for at least three of the four most divergent loci, i.e., Idh-1°", Pgi-2"°, Pgm-2°© (scored for only two of these individuals) and Tpi-2°°. No other individuals in the entire data set possessed this genotype combination. GENETIC VARIATION.—Genetic variation was quantified for each population and the species as a whole by computing three standardly used indices. Val- ues for percent loci polymorphic (P) ranged from 23.5% to 47.1%, with a mean of 36.48%; for mean number of alleles per locus (A), the range was 1.5 to 2.5, mean 1.97; and for mean expected heterozygosity (Hg), the range was 0.112 to 0.147, mean 0.129 (Table 4). The mean values for these indices in TABLE 3. Allele frequencies for 17 isozyme loci in ten populations of Athyrium filix-femina in eastern North America. angustum asplenioides Mt. Ste. Barnet North Schenevus’ Ralston Shirley Mountjoy Hopewell Sandy Run Pond Drain Locus Allele Hilaire QE VT Hudson NY NY PA NJ Store VA Vv Swamp NC V. Ald A. 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 0.969 B 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.021 G 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.010 (N) 9 i 1 48 Got A 0.000 0.033 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 B 1.000 0.951 1.000 1.000 1.000 0.955 1.000 0.917 1.000 0.980 C 0.000 0.016 0.000 0.000 0.000 0.045 0.000 0.083 0.000 0.010 D 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.010 (N) 5 6 Hk* A 0.000 0.000 0.000 0.000 0.150 0.000 0.017 0.000 0.000 0.016 B 0.960 0.965 1.000 0.955 0.800 0.868 0.967 0.967 0.893 0.952 C 0.040 0.023 0.000 0.045 0.050 0.105 0.017 0.017 0.089 0.032 E 0.000 0.012 0.000 0.000 0.000 0.025 0.000 0.017 0.000 0.000 F 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.018 0.000 (N) 25 43 19 11 10 19 30 30 28 ot Idh-1* A 0.019 0.007 0.000 0.000 0.000 0.828 0.950 0.833 0.859 0.920 B 0.596 0.676 0.596 0.545 0.650 0.172 0.017 0.133 0.094 0.060 C 0.385 0.296 0.385 0.455 0.350 0.000 0.000 0.000 0.000 0.020 D 0.000 0.021 0.019 0.000 0.000 0.000 0.000 0.000 0.000 0.000 F 0.000 0.000 0.000 0.000 0.000 0.000 0.033 0.033 0.047 0.000 (N) 26 11 20 32 30 30 32 50 Lap A 0.000 0.000 0.000 0.000 0.000 0.018 0.000 0.000 0.000 B 0.100 0.038 0.063 - 0.175 0.000 0.036 0.000 0.031 0.016 C 0.850 0.885 0.750 - 0.550 1.000 0.946 1.000 0.938 0.935 D 0.050 0.058 0,188 - 0.275 0.000 0.000 0.000 0.016 0.032 E 0.000 0.019 0.000 - 0.000 0.000 0.000 0.000 0.016 0.000 F 0.000 0.000 0.000 - 0.000 0.000 0.000 0.000 0.000 0.016 (N) 0 86L (z00z) € MASINON 26 ANNTIOA “TYNUNO! NUdA NVOIMSNV TABLE 3. Continued. angustum asplenioides Mt. Ste. Barnet North Schenevus Ralston Shirley Mountjoy Hopewell Sandy Run Pond Drain Locus Allele Hilaire QE VT Hudson NY NY PA NJ Store VA VA Swamp NC VA Mdh-1 A 0.981 0,950 0.942 1.000 1.000 0.939 0.967 1.000 0.969 1.000 B 0.019 0.020 0,058 0.000 0.000 0.061 0.033 0,000 0.000 0.000 Cc 0.000 0.030 0,000 0.000 0.000 0.000 0.000 0,000 0.031 0.000 (N) 26 50 26 Mdh-2 A 1,000 0.990 1.000 1.000 1.000 1,000 0.933 0.950 1.000 1.000 B 0,000 0,010 0.000 0.000 0.000 0.000 0.067 0.050 0.000 0.000 (N) 11 20 33 15 20 Mdh-3 A 1.000 1.000 1.000 1.000 1.000 0.985 0.967 1.000 0.953 1.000 B 0,000 0.000 0.000 0.000 0.000 0.015 0.033 0.000 0.047 0.000 (N) 33 50 Mdh-4 A 0.960 0.920 0.962 1,000 1.000 1.000 0.967 0.950 1.000 1,000 B 0.040 0.070 0.038 0,000 0,000 0.000 0.033 0.050 0.000 0.000 C 0.000 0.010 0.000 0,000 0.000 0.000 0.000 0.000 0.000 0,000 (N) Pgi-1 A 1.000 1.000 1.000 1,000 1.000 1.000 1.000 1.000 1.000 1,000 (N) 11 10 13 25 24 27 Pgi-2 A 0.000 0.000 0,000 0,000 0.000 0.015 0.033 0.000 0.000 0.000 B 0.000 0.014 0.019 0,000 0.000 0.000 0.017 0.133 0.000 0.000 Cc 0.019 0.00 0,000 0,000 0,000 0.000 0.000 0.017 0.000 0.010 D 0.000 0.000 0.000 0.000 0.000 0.045 0.100 0.000 0.078 0.010 E 0.692 0.715 0.808 0.500 0.550 0.742 0.800 0.850 0.906 0.930 F 0.019 0.014 0.000 0.000 0.000 0.045 0.017 0.000 0.000 0.020 G 0.269 0.250 .154 0.409 0.450 0.106 0.033 0.000 0.016 0.030 H 0,000 0,000 0.000 0,000 0.000 0.045 0.000 0.000 0,000 0.000 I 0,000 0,000 .019 0.000 0.000 0.000 0.000 0.000 0.000 0.000 J 0,000 0,000 0.000 0.091 0.000 0.000 0.000 0.000 0.000 0.000 (N) 26 te 26 33 Pgm-2 A 0.024 0.000 0,000 0.000 0.000 0.000 0.000 0.083 0.000 0.020 B 0,857 0.856 0.906 1.000 0.975 0.143 0.300 0.100 0.222 0.390 C 0.119 122 0.094 0.000 0.025 0.857 0.683 0.800 0.704 0.580 WOIYAHLY NVOTYIWNV HLYON NYALSVS “TV LY AAOTIAN 661 TABLE 3. Continued. 0072 angustum asplenioides Mt. Ste. Barnet North Schenevus’ Ralston Shirley Mountjoy Hopewell Sandy Run Pond Drain Locus Allele Hilaire QE VT Hudson NY NY PA NJ Store VA VA Swamp NC VA D 0.000 0.022 0.000 0,000 0.000 0.000 0.017 0.017 0.074 0.000 E 0,000 0,000 0.000 0,000 0.000 0.000 0.000 0.000 0.000 0.010 (N) 21 45 16 i ba | 20 21 30 30 27 50 6Pgd-1 A 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 B 1.000 1.000 1.000 1,000 1.000 0.957 1.000 1.000 1.000 1.000 0,000 0.000 0,000 0.000 0.000 0.043 0.000 0.000 0.000 0.000 (N 11 38 11 6 23 6Pgd-2 A 0,000 0.000 0.000 0,000 0,000 0.015 0.033 0.017 0.017 0.021 B 1.000 1.000 1.000 1.000 1.000 0.985 0.950 0.967 0.983 0.947 0.000 0,000 0.000 0.000 0,000 0.000 0.017 0.017 0.000 0.021 D 0.000 0.000 0.000 0.000 0.000 0,000 0.000 0.000 0.000 0.011 (N 26 69 21 11 Skdh A 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.067 0.000 0.000 B 0.981 0.955 0.942 0.955 1,000 1,000 1.000 0.933 1.000 0.990 Cc 0.019 0.036 0.058 0.045 0.000 0,000 0.000 0.000 0.000 0.010 D 0.000 0.009 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 (N) 26 56 26 11 20 Tpi-1 A 0.947 1.000 0.981 0.900 1.000 0.970 1.000 1.000 0.984 1.000 B 0.053 0.000 0.000 0.100 0.000 0.030 0.000 0.000 0.016 0.000 Cc 0.000 0.000 0.019 0.000 0.000 0,000 0.000 0.000 0.000 0.000 (N) 19 47 26 5 33 Tpi-2* A 0.579 0.601 0.404 0.400 0.450 0.318 0.440 0.420 0.406 0.429 B 0.368 0.355 0.538 0.500 0.550 0.076 0.100 0,000 0.063 0.255 C 0.053 0.036 0.058 0.100 0.000 0.606 0.460 0.580 0.531 0.316 E 0.000 0.007 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 (N) 19 69 26 5 10 a0 (z00Z) € WAAINAN 26 SWN'IOA *TVNYNOL NYA NVORANV KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 201 these eastern North American Athyrium populations were somewhat greater than the means for ferns (P=36.0%, A=1.65, H=0.109) obtained by averaging across 32 taxa (Li and Haufler, 1999), which are in turn similar to the means for angiosperms (P=34.2%, A=1.53, Hp=0.113; Hamrick and Godt, 1989). Mean values across A. angustum populations for P (34.12%) and A (1.84) were slightly lower than those for A. asplenioides (mean P=38.84%, mean A=2.1), while the mean value in A. angustum for Hy=0.138 was greater than that of 0.119 in A. asplenioides. This indicates that while A. asplenioides possesses greater allelic diversity, frequencies of alleles within loci are more evenly distributed in A. angustum populations. Values for mean observed heterozygosity (Ho) were also computed for comparison to Hg. Values of Ho were very similar to (although tending to be slightly greater than) those for Hg, suggesting a tendency toward random mating. COMPARISON TO HARDY-WEINBERG.—Genotype proportions for each polymorphic locus in each population were compared to Hardy-Weinberg expected values by computing the fixation index F, and the statistical difference of F from 0 was evaluated using the chi-square test, with pooled genotype classes if the number of alleles exceeded 2 (Table 5). The test was considered valid if two of the three genotype classes were represented by expected values > 5. Of 42 validly tested loci, 40 conformed to Hardy-Weinberg expectations. Moreover, 45 out of 51 non-valid tests also indicated conformance to Hardy- Weinberg values, despite the tendency for non-valid tests to indicate false non-conformance due to low expected values. Thus, mating was inferred to approximate random mating via predominant outcrossing between gameto- phytes, as appears to be the general case in most ferns (Soltis et al., 1988) including Athyrium species (Schneller, 1979). GENETIC RELATEDNESS OF POPULATIONS AND TAxXA.—The degree of genetic diver- gence among populations (Table 6) was quantified by computing F-statistics (Wright, 1965, 1978), including hierarchical analysis (Wright, 1978). Values for Fsy computed among all populations varied among loci from a low of 0.019 for 6Pgd-2 to high values of 0.468 for Pgm-2 and 0.460 for Idh-1, the latter two loci having exhibited the greatest allele frequency difference between A. angustum and A. asplenioides. The value across loci of Fs;r= 0.255 indicated very substantial differentiation among populations. This val- ue is very high in comparison to other fern species examined, exceeding, for example, computed values for mean Fs of 0.024 among populations of Poly- stichum munitum ranging from Oregon to Idaho (Soltis et al., 1987), Fs;= 0.152 among Pteridium aquilinum populations ranging from Massachusetts to Florida (Speer et al., 1998), and Fs;=0.100 to 0.248 in various species of Dryopteris ranging widely across eastern North America (Werth, ms.). To evaluate the contribution of differences between A. angustum and A. asplenioides to overall population differentiation, hierarchical F-statistic analysis (Wright, 1978) was carried out (Table 6). For most individual loci, as well as for the combined values across loci, the variance between the two TaBLe 4. Estimates of genetic variation at 17 loci in ten populations of Athyrium filix-femina s. 1. (standard errors in parentheses). Mean sample size Mean no, of alleles t loci Mean heterozygosity ardy-Weinberg H Expected (Hp) aa per locus per locus (A) Re asia (P) Observed (H,) A. angus Mt. Ste. a QE 21.4 (1.4) 1.9 (0.2) 35.3 0.150 (0.056) 0.140 (0.047) Barnet, VT 52.5 (3.4) 2.5 (0.3) 41.2 0.134 (0.042) 0.141 (0.042) North Hudson, NY 21.7 (1.5) 1.8 (0.2) 41.2 0.133 (0.046) 0.136 (0.047) Schenevus, NY 9.1 (0.8) 1,5:(0,2) 2a.0 0.145 (0.071) 0.126 (0.055) Ralston, PA 15.5 (1.4) 1.5 (0.2) 29.4 0.150 (0.061) 0.147 (0.057) ean + 1.84 34.12 0.142 A. asplenioides Shirley, NJ 28.2 (2.0) 2.0 (0.3) 35.3 0.127 (0.041) 0.127 (0.040) Mountjoy Store, VA 22.8 (2.1) 2,2:(0:3) 41.2 0.133 (0.052) 0.123 (0.042) Hopewell, VA 25.7 (1,8) 1.9 (0.2) 47.1 0.125 (0.038) 0.118 (0.036) Sandy Run Swamp, NC 28.9 (1.5) 2.0°(0.2) 85.0 0.104 (0.039) 0.117 (0.041) Pond Drain, VA 44.5 (2.3) 2.4 (0.3) 35.2 0.118 (0.049) 0.112 (0.046) ean 29.62 ypil 38.84 0421 0.119 Mean across all 26.83 1.97 36.48 0.131 0,129 populations * A locus is considered polymorphic if the frequency of the most common allele does not exceed 0.95. > Unbiased estimate (Nei, 1978 c0Z (z00Z) € MAGWON 26 ANN TIOA “TVNUNOL NSA NVOMENV TABLE 5. ye 280 of se fixation index F as a measure of the conformance of population genotype ratios to those saa under Hardy-Weinberg ulation, and tested for statistical difference from 0 (i.e ci with more than two alleles and therefore more than three genotypic ts were se ang valid only if two of the three genotypic classes were represented by an expected values > 5. Results fro ackets. Values marked as ‘“‘ns’ be phen ne sane ‘a eae to Hasty eres expected proportions (p > equilibriu classes, Chi-square tes F was c omputed for each polymorphic locus in each popu Weinberg cone using the chi-square test. Pooling was carried out for loc non-valid tests are provided in and therefore are not statistically different from 0. Values marked w conformance to Hardy- < 0.05 (one asterisk), 0.05), p = 0.01 (two asterisks), or p < 0.001 (three asterisks). Mt. Ste. Mountjoy Sandy Run Hilaire Barnet N Hudson Schenevus _ Ralston Shirley e Hopewell Swamp Pond Drain Locus QE VT NY NY PA NJ VA VA Ald [—0.025 ns] Got 040 {[—0.048 ns] —0,091 ns {[—0.015 ns] Hk {[-0.042 — {—0.028 ns] —0.048 ns} [0.403 ns] —0.124ns [-—0.026 ns] [—0.026 ns] 0.266 ns {—0.039 ns] Idh —0.00 102 ns 0.148 ns {0.267 ns] —0.208 ns [0.653***] -—0.163 ns 0.002 —0.070 ns Lap [10, = on —0.087 ns 0.216 ns 0.154 ns {[—0.043 ns] [0.216***] [-—0.046 ns] Mdh-1 {—0.020 ns .376** [-0.061 ns] {[-0.065 ns] [—0.034 ns] [-0.032 n Mdh-2 {-0.010 n [-0.071 ns] [—0.053 ns] Mdh-3 [-0.015 ns] [—0.034 ns] [0.650***] Mdh {1.000***]} 0.192 ns [—0.040 ns {[—0.034 ns] [—0.053 ns] Pgi-2 —0.375 0.021 ns -—0.190 ns [—0.266 ns] —0.212 n —0.054 ns 0.136 156 ns —0.088 ns —0.048 ns Pgm-2 0.240 ns 0.120 ns [—0.103 ns] {—0.026 ise -0.167 ns —0.280 ns 0.028 ns 0.259 ns —0.057 ns 6Pgd-1 [1.000***] 6Pgd-2 {-0.015 ns] [—0.040 ns] [—0.026 ns] [—0.017 ns] -—0.038n Skdh {[—0.020 ns} —0.039 ns [—0.061 ns] [—0.048 ns| {—0.071 ns] {[-—0.010 aa No. tests 3[5] 8[2] 4[5] o[5] 3[3] 5[5] 27] 5(5] 5[3] 5[5] Total: 40[45] showing conformance to HW' No. of tests 1[(1] 1[0] o[0]} o[1] o[o] o[1] o[1] o[o] 0[2] o[0} Total: 2[6] showin 8 non-conformance to HW! * number of non-valid tests in brackets. WOIYAHLVY NVOTHYNV HLYON NYSLSVa YTV LY AAOTISN £02 TABLE 6, i statistic analysis (Wright, 1965, 1978), including enc wage ie 1978), for 16 polymorphic loci across Athyrium filix- femin pulations in eastern North America. Statistical differe evaluated using contingency chi-square analysis. For biseared tal analysis, Athyrium populations were assigned to their ccs a eens or asplenioides). Differences between values of Fs; and Fiocality/total are attributable to differences in computational method (discussed in Swofford and Selander, 1981). Hierarchical F-statistic Analysis (variance @ F Statistics mponents in parentheses) FOZ Total Limiting Variance Locus Fis Fer Fst Focality/Total FLocality/Taxon Fraxon/Total ria Ald —0,025 —0.002 0.022 ns 0.012 (0.00007) 0.014 (0.00009) —0.002 (—0.00002) 0.00623 Got —0.060 =O:017 0.041 ns 0.028 (0.00109) 0.030 (0.00116) —0.002 (—0.00007) 0.03895 Hk 0.116 0.160 0.049*** 0.021 (0.00273) 0.031 (0.00398) —0.010 (—0.00125) 0.12816 Idh-1 0.043 0.485 0460*** 0.452 (0.29124) 0.002 (0.00073) 0.451 (0.29050) 0.64377 Lap 0.077 0.204 O137*** 0.113 (0.02383) 0.092 (0.01886) 0.024 (0.00497) 0.21012 Mdh-1 0.036 0.060 0.025 ns 0.008 (0.00037) 0.014 (0.00068) —0.006 (—0.00031) 0.04939 Mdh-2 —0.059 =—0:013 0.044** 0.017 (0.00042) 0.016 (0.00040) 0.001 (0.00001) 0.02501 Mdh-3 0.302 0.321 0.028 ns 0.007 (0.00013) 0.002 (0.0003) 0.005 (0.00010) 0.01889 Mdh-4 0.208 0.230 0.029 ns 0.010 (0.00046) 0.014 (0.00067) —0.005 (—0.00021) 0.04724 Pgi-2 —0.143 0,012 OF 15" ** 0.096 (0.03914) 0.040 (0.01551) 0.058 (0.02363) 0.40775 Pgm-2 —0.001 0.469 0.468""* 0.459 (0.23433) 0.045 (0.01293) 0.434 (0.22140) 0.51042 6Pgd-1 1.000 -000 0.039 ns 0.018 (0.00016) 0.023 (0.00020) —0.004 (—0.00004) 0.00866 6Pgd-2 —0.032 —0;013 0.019 ns 0.011 (0.00037) 0.000 (0.00000) 0.011 (0.00037) 0.03325 Skdh —0,052 —0.019 0.031"** 0.013 (0.00061) 0.011 (0.00054) 0.001 (0.00006) 0.04782 Tpi-1 =0,070 —0.020 0.047 ns —0.003 (—0.00014) 0.000 (0.00001) —0.004 (—0.00' ) 0.04269 Tpi-2 —0.214 =0,029 6:159""* 0.137 (0.08853) 0.013 (0.00763) 0.125 (0.08091) 0.64843 Combined across loci —0.052 0.217 0.255""* 0.238 (0.68333) 0.028 (0.06343) 0.216 (0.601990) (Z00Z) € WAAWON 26 ANN'TIOA “TVNYNOlL NAA NVORMANV KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 205 taxa with respect to the total (Fxyy=0.216) was an order of magnitude greater than variance among populations within taxa (Fxy=0.028). Thus, differences between taxa explained most of the variance among populations with respect to the total (Fxy=0.238). This result is consistent with and explained by the large allele frequency differences at the four most polymorphic loci (Idh-1, Pgm-2, Pgi-2, and Tpi-2) between populations of different taxa as opposed to populations of the same taxon. Values for Nei’s Genetic Identity, I, (Nei, 1978) and Rogers’ Genetic Simi- larity, S, (Rogers, 1972) were computed for each pair of populations (Tables 7 and 8). Values for these indices were consistently higher between popula- tions of the same taxon, ranging from I=0.990 to 1.000 and S=0.930 to 0.975, than between populations of different taxa, ranging from I=0.875 to 0.938 and S=0.803 to 0.881 (Table 8). Populations were clustered using the Unweighted Pair-group Method with Averaging (UPGMA) based on both S and I. The two indices resulted in very similar dendrograms that differed only in the association among some of the A. asplenioides populations; only the dendrogram based on S is illustrated (Fig. 5). Two clusters, each comprising all the populations of one taxon, were joined at S=0.849. Within the A. angustum cluster, the two northern- most populations Mt. Ste. Hilaire, QUE and Barnet, VT were placed as most similar, joined at S=0.975, and to this cluster the other three A. angustum populations were joined successively in order from north to south; the southernmost A. angustum population Ralston-1, PA joined the A. angustum cluster at S=0.938. The topology of this A. angustum cluster was identical in the dendrogram based on I (not shown). In A. asplenioides, the most simi- lar populations based on S=0.966 were the southernmost population Sandy Run Swamp, NC and the next most southeastern occurring population Mountjoy Store, VA, and to these were joined the Pond Drain, VA popula- tion from the mountains of southwestern Virginia at S=0.958 to form a subcluster. A second subcluster, comprising the two northernmost A. asple- nioides populations Shirley, NJ and Hopewell, VA, joined the more southern subcluster at S=0.947. The topology of the A. asplenioides cluster differed somewhat in the dendrogram based on I (not shown) indicating that the geo- graphic ‘‘signal”’ in the A. asplenioides data is weaker than in the A. angus- tum data. DISCUSSION The A. filix-femina complex, distributed across four continents and com- prising as many as four North American taxa with overlapping ranges, pro- vides an especially suitable context for exploring patterns and processes of divergent evolution and its taxonomic consequences in ferns. The two east- ern North American taxa, A. angustum and A. asplenioides, long have been perceived as close relatives separable by distinctive characters that are con- sistent within the vast northern and southern areas they respectively occupy, 206 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) LE 7. x of pairwise values for Rogers (1972) genetic similarity (above diagonal) and Nei heer ae genetic identity (below diagonal). Population al 2 3 4 5 6 7 8 5 10 1. Mt. Ste. Hilaire, QE - 0.975 0.962 0.953 0.939 0.853 0.865 0.847 0.862 0.881 2. Barnet, VT 1.000 - 0.956 0.938 0.930 0.851 0.864 0.853 0.858 0.879 3. North Hudson, NY 0.999 0.997 - 0.944 0.943 0.844 0.856 0.840 0.852 0.873 4. Schenevus, NY 1.000 0.996 0.995 - 0.942 0.841 0.842 0.831 0.843 0.863 5. Ralston, PA 0.993 0.990 0.994 0.993 - 0.817 0.821 0.803 0.825 0.843 6. Shirley, NJ 0.915 0.913 0.906 0.900 0.883 - 0.943 0.951 0.959 0.932 7. Mountjoy Store, VA 0.924 0.921 0.916 0.908 0.892 0.996 - 0.950 0.966 0.959 8. Hopewell, VA 0.912 0.911 0.903 0.893 0.875 0.998 0.997. - 0.951 0.937 9. Sandy Run Swamp, VA 0.924 0.921 0.916 0.905 0.891 0.998 1.000 0.998 - 0.956 10. Pond Drain, VA 0.938 0.936 0.934 0.923 0.911 0.990 0.998 0.990 0.996 - but that intergrade and recombine to form a hybrid zone in their relatively narrow region of overlap. This perception is supported by the data from the present study. The spores of the two taxa show striking and consis- tent differences in the perispore sculpturing, a low papillate perispore in A. angustum versus a rugose perispore in A. asplenioides, and frequencies of allozymes exhibit strong differences between as compared to within the taxa. An abrupt shift in allele frequencies at the four most polymorphic loci (Idh-1, Pgm-2, Tpi-2, and Pgi-2) of Athyrium corresponds to the geographic boundary between A. angustum and A. asplenioides, i.e., between northern Pennsylvania and southern New Jersey in our sample. Although virtually all alleles were shared between the two taxa, some alleles that predominated or were frequent in one taxon were nearly absent in the other, e.g., Idh-1“ of A. asplenioides, Idh-1° of A. angustum, and Pgi-2° of A. angustum. In other cases, alleles were more extensively shared between the taxa but at very different frequencies, e.g. Idh-1", Pgm-2°, Pgm-2°, and Tpi-2" (Table 3). UPGMA analysis of Athyrium resulted in two distinct taxon clusters joined at a substantially lower similarity value (S=0.849) than that joining popula- tions within their respective taxon clusters (S=0.938 for A. angustum; TABLE 8. Means of pairwise values for iy va Similarity (S) and Nei’s Genetic Identity (I) for eae within and between Athyrium taxa. Ranges are given in parentheses. Each category was represented by ten pairwise comparisons. Taxon combination I S angustum—angustum 0.996 0.948 (0.990—1.000) (0.930—0.975) asplenioides—asplenioides ; 0.951 (0.990—1.000) (0.932-0.966) angustum-asplenioides 0.911 (0.875—0.938) (0.803-0.881) KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 207 Mt. Ste. Hilaire, QUE Barnet, VT North Hudson, NY angustum ei Schenevus, NY Ralston-1, PA Shirley, NJ — [ i Hopewell., VA i Mountjoy Store, VA asplenioides = Sandy Run Swamp NC Pond Drain, VA Rogers’ Similarity . 5. Dendrogram resulting from UPGMA analysis based on pairwise values of Rogers’ Simi- larity (Table 7). S=0.947, for A. asplenioides). These spore and allozyme differences for the most part are consistent among populations of each taxon, yet there is evi- dence of introgressive hybridization in their region of overlap as discussed below. The distinctness of these taxa and the consistency of the character differ- ences specified by Butters (1917) frequently have been questioned by state- ments indicating that the taxa intergrade extensively (e.g., Benedict, 1934; Weatherby, 1936; Fernald, 1946; Shaver, 1954; Wherry, 1961). However, such statements seem anecdotal in that they are not accompanied by docu- mentation of character combinations in specimens. The most extensive specimen-based data set is that of Liew (1971, 1972), who carried out a phe- netic analysis of North American Athyrium sensu lato based on 170 speci- mens scored for 99 characters. Liew indicated that cluster analysis based on paired similarity coefficients separated each Athyrium taxon, including A. angustum and A. asplenioides, into its own cluster, but the summary dendrograms presented leave it uncertain as to whether some of the speci- men placements were inappropriate or equivocal. In the present study, the degree of differentiation between A. angustum and A. asplenioides may appear exaggerated by the omission of localities where the two taxa co-occur. This omission was neither intentional nor an oversight, rather it resulted from a failure to discover such localities despite a substantial effort to do so. A search for co-occurring populations in south- ern Pennsylvania resulted in finding only a few isolated individuals of A. asplenioides in this highly agriculturalized and urbanized region, and no sizable populations from which allele frequency data could be obtained. All individuals in populations sampled for the present investigation from Quebec south to northern Pennsylvania were readily assignable to A. angus- tum on the basis of leaf and spore morphology. Similarly, all individuals in 208 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) populations from North Carolina north to northern Virginia were assignable to A. asplenioides. However, a genetic influence from A. angustum is sug- gested by the high frequency of characteristically A. angustum alleles Idh-1” and Pgi-2° in the northernmost A. asplenioides population Shirley, NJ, as well as Pgm-2” and Tpi-2” in the highest elevation A. asplenioides popula- tion Pond Drain, VA. Evidence of introgressive hybridization is augmented by the intermediate spore morphologies encountered at the Shirley, NJ, locality. No individuals assignable to A. angustum were found at this site, but it is possible that A. angustum spores could have migrated from populations fur- ther north and effected hybridization (Wagner, 1943). The prevalence of fully intermediate spores in six of the individuals, five of which are heterozygous for taxon marker allozymes, suggest that these are first-generation hybrids. The skewed array of spore morphologies of other individuals suggests that they are backcrosses and provides evidence that hybridization has gone beyond the first generation resulting in introgression. Thus, spore and iso- zyme data combine to support previous morphology-based suggestions that the region of overlap between the taxa, and possibly higher elevations in the southern Appalachians as well, could represent a hybrid zone resulting from secondary contact between recently diverged sister taxa (Benedict, 1934; Shaver, 1954; Sciarretta et al., ms). Narrow hybrid zones in which formation of fertile hybrids and backcrosses result in intergradation between divergent taxa occur between numerous spe- cies or subspecies pairs of animals and angiosperms (reviewed by Arnold, 1997), and occur in a few agamosporous ferns (Gastony and Windham, 1989). Hybridization between fern species usually results in spore abortion due to abnormal meiosis, and introgressive hybridization between homo- ploid taxa as divergent as A. angustum and A. asplenioides is decidedly rare in ferns, with only three cases having been documented previously: (1) swarms of fully fertile hybrids involving Pteris quadriaurita and P. multia- urita, first generation as well as backcrosses, occur in disturbed forests of Sri-Lanka (Walker, 1958); (2) extensive hybridization among three species of the tree fern genus Alsophila resulted in a complex swarm of fertile hybrids in Puerto Rico, a scenario hypothesized to give rise to new species through allogamous allohomoploidy (Conant and Cooper-Driver, 1980; Conant, 1990); (3) morphological and molecular data provide evidence of hybrid swarms between Polystichum munitum and P. imbricans of northwestern North America (Mayer and Mesler, 1993; Mulleniex et al., 1998). Of these, the situa- tion in Polystichum most closely parallels that of the Athyrium taxa angus- tum and asplenioides. The two Polystichum taxa share alleles at most enzyme loci, but are differentiated with respect to frequencies at three loci resulting in a lower magnitude of genetic identity between the taxa (I= 0.842) than within each taxon (means: I=0.974 for P. imbricans, 1=0.957 for P. munitum; Soltis et al., 1990, 1991). Additionally, the Polystichum taxa hy- bridize introgressively, although they do not form a hybrid zone as in Athy- rium; rather hybridization occurs at various localities across a broad area of sympatry between the two taxa (Mayer and Mesler, 1993; Mulleniex et al., KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 209 1998). Maintenance of distinction between these two Polystichum species most likely results from a combination of their partial intersterility and di- versifying selection imposed by the very different habitats—forests versus exposed cliffs—occupied by P. munitum and P. imbricans respectively. The nature, frequency, and geographic extent of hybridization between A. angustum and A. asplenioides remain uncertain and merit further research that combines field, herbarium, molecular, and breeding studies. It is critical to determine with greater precision and full documentation the degree to which these two taxa maintain versus blend their macromorpho- logical, micromorphological, and allozymic differences where they coexist. It is unknown whether there is preferential mating within taxa, whether taxon characters tend to remain associated in the face of hybridization or are com- pletely recombined, and whether there exists a cline for allozyme frequen- cies and morphological characters within the overlap region. There is a need for intensive exploration for coexisting A. angustum and A. asplenioides pop- ulations in the areas of their overlap from the eastern seaboard through the midwest, and in the southern Appalachians where the existence of cryptic taxa has been hypothesized (Wagner and Wagner, 1966). Isozyme studies of these populations should be coordinated with critical analyses of morphologi- cal character combinations obtained from numerous specimens and with experimental crosses that can quantify the propensity of the taxa to hybridize. Beyond the taxonomic significance of hybridization, the unprecedented formation of normal intermediates between spores of such divergent mor- phology provides an opportunity to gain insight into the genetics underlying spore morphology (Schneller, 1989). The variability of spore morphology within individuals of the putative primary hybrids from the Shirley, NJ site, which includes expression of parental types, indicates that inheritance is polygenic rather than a simple one-or-two-gene inheritance mechanism, and that the spore genotype determines or at least influences perispore pheno- type. Contrasting observations and inferences were obtained by Schneller (1989), who reported the formation of normal intermediate spores in experi- mentally produced hybrids between A. angustum and A. asplenioides, but found that all spores from a sporangium were of the same type, implicating sporophytic determination of the perispore morphology. Explanations that can account for these differing observations include the possibility that the Shirley, NJ, plants were not true F1 hybrids, or that there may be a maternal effect that varies as a polymorphism. Clearly, further studies of the inheri- tance of spore morphology are in order. TAXONOMIC CONCLUSION: AT WHAT RANK SHOULD A. ANGUSTUM AND A. ASPLE- NioipES BE RECOGNIZED?—Over the second half of this century, the nature of pteridophyte species has been clarified significantly by application of tech- nological advances in cytology and molecular systematics in combination with detailed morphological and field studies (Manton, 1950; Wagner, 1963; Haufler, 1987, 1989; Paris et al., 1989; Conant, 1990). Nonetheless, definitive ranking of closely related divergent taxa with overlapping ranges, such as 210 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) the two Athyrium taxa considered here, remains challenging due to the un- settled controversy as to the nature and definition of species (e.g., Mayr, 1992; Davis and Nixon, 1997; Baum, 1998; DeQueroz, 1998) as well as to the unpredictable nature of reproductive interactions between diverged taxa ex- periencing secondary contact (Arnold, 1997). The rank assigned to A. angus- tum and A. asplenioides has varied considerably in floristic treatments published in this century. While Small (1938) and Wherry (1948, 1961) fol- lowed Butters (1917) in treating these taxa as separate species, other authors have tended to treat them as infraspecific taxa, either subspecies (Lellinger, 1985) or varieties (Fernald, 1950; Mickel, 1979; Cody and Britton, 1989; Gleason and Cronquist, 1991; Kato, 1993). Spore and isozyme data combined indicate that populations of these two taxa are significantly divergent, exhibiting greater differences than ordinarily encountered within single species of ferns thus far studied. However, the fertility of hybrids (Schneller, 1989) provides a potential for introgression between the two taxa. The preliminary evidence that hybridization and introgression do in fact occur indicates that A. angustum and A. asple- nioides would be considered conspecific under species definitions as differ- ent as the Biological Species Concept (Mayr, 1942) and the Phylogenetic/ Diagnostic Species Concept (Cracraft, 1983; Davis and Nixon, 1992). None- theless, in practice numerous pairs of taxa that form hybrid swarms or zones are treated as distinct species (Grant, 1981; Arnold, 1997). The occurrence of northern and southern infraspecific taxa in eastern North American Athyrium is paralleled in Pteridium aquilinum L., which comprises northern and southern varieties Jatiusculum (Desv.) Underw. and pseudocaudatum (Clute) Heller, respectively, and which shows north to south clines in allele frequencies near the overlapping taxon boundary (Speer et al., 1998). However, the pattern of allozyme variation in Pteridium differs from that in Athyrium in that the most abrupt shift in allele frequen- cies occurs within the range of P. latiusculum, genetic identities between populations of the two Pteridium varieties ranged substantially higher (I= 0.916 to 0.999) than those of the two Athyrium taxa (I=0.875—0.938, mean =0.911), and UPGMA clustering failed to separate the varieties (Speer et al., 1998). On the basis of the lack of genetic differentiation between P. latiuscu- lum and P. pseudocaudatum, variety was indicated as the highest rank at which to recognize them (Speer et al., 1998). In contrast, the more substan- tial differentiation between the Athyrium taxa angustum and asplenioides and the consistent characters uniting a vast number of individuals north and south of their hybrid zone suggest that the taxa should be ranked at least at the level of subspecies. Ranking at the species level would not be inconsis- tent with the treatment of such taxa in the broader plant literature. AACCKNOWLEDGMENTS The authors thank John Kress and Vicki Funk for facilitating portions of this research in vari- ous ways; Mark Strong, Cynthia Caplen, Kim Sciarretta, Erin Potter and Heather Hartless for KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 211 help with sng Se Aig Nowicke for advice on preparation and interpretation of electron sae sane: Walter R. Brown and Susann Braden for help with the use of the SEM; Stanley Yank for bec on light microscopy; Cheryl Roesel, Lori Croisatiere, Thao Le, and Leigh Ann aida for assistance with electrophoresis; Greg McKee for carrying out germination tests; and David Lellinger for comments on the manuscript. This research was supported by a eorge Mason University Research Fellowship awarded to CLK, by NSF Equipment Grant the Structure and Evolution of Terrestrial Ecosystems awarded by the U. S. National Herbarium and NSF Grant DEB-9220755 awarded to CRW. A portion of this study is based on the MS thesis of CLK completed at George Mason University. LITERATURE CITED ARNOLD, M. L. 1997. Natural hybridization and evolution. Oxford University Press, NY. Baum, D. A. 1998. Individuality and the existence of species through time. Syst. Biol. 47: 64 BENEDICT, R. C. 1934. Can dens readily distinguish the northern and southern lady fern spe- cies. eI Fern J. 24:117-11 Butters, F. K. 1917. Taxonomic and geographic studies in North American ferns. I. The genus Athyrium pee the North American ferns allied to Athyrium filix-femina. Rhodora 19: 169-207 Copy, W. J. AND D. M. Brirron. 1989. Ferns aha fern allies of Canada. Publication 1829/E. Resea eck picages pearance Canada, Ott Conant, D. S. 0. Observations on the reproductive biology of Alsophila species and hybrids ee a aak Ann. Missouri ae sas :290-296 AND G. COOPER-DRIVER. 1980. togamous silokiomoplosdy in Alsophila and Nephelea ieijattienceand a new hypothesis is speciation in homoploid homosporous ferns. Amer. J. Bot. 67:1269-1288. CracrarT, J. 1983. Species concepts and speciation analysis. Curr. Ornith. 1:159-187. Davis, J. I. AND K. C. Nixon. 1992. Ph eating genetic variation, and the delimitation of phylo- genetic come Syst. Biol. 41:421-4 Dawes, C. J. 1971. Biological techniques in mee microscopy. Barnes and Noble, New York. Say sa Pp erlocher, eds. Endless forms: Species and Secaceeiarie Oxford Ss) aa thy versity Press, Oxford, Englan ERDTMAN, G. 1969. Handbook of palynology. Hafner, New York. dicing M. L. 1946. Some trivial American forms of lady fern. Rhodora 48:389—391. 950. Gray’s manual of botany: a handbook of the flowering plants and ferns of the cen- aa ne northeastern United States and adjacent Canada. American Book Company, Gastony, G. J. AND D. C. Darkow. 1983. Chloroplastic and cytosolic isozymes of the homosporous e . Species sin in eo the treatment and defini- tion of amnesonous epoctoe. Amer. Fern J. 7 Gueason, H. A. AND A. Cronquist. 1991. Manual of be cane plants of northeastern United States and adjacent Canada. 2d ed. The New York Botanical Garden, Bronx, NY. Grant, V. 1981. Plant speciation. Columbia University Press, N Hamrick, J. L. anp M. J. W. Gopt. 1989. Allozyme diverary } in plant species. Pp. 43-63 in A. H. D. Brown, M. T. Clegg, A. L. Kahler, and B. S. Weir, eds. Plant population genetics, breedi ing and ete resources. Sinauer Associates, Sunderland, MA. Haur.er, C. H. 1987. Electrophoresis is modifying our concepts of evolution in homosporous pteridophytes. Amer. J. Bot. 74:95—-966. ———. 1989. Species concepts in the pteridophytes: summary and synthesis. Amer. Fern J. 212 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) Kato, M. 1993. Athyrium. Pp. 255-258 in FNA Editorial Committee, - Flora of North America, me : sige wnbal and gymnosperms. Oxford Univ. Press, iste C L. 0. A cones iss of the eastern North American va fern, Athyrium filix- femina Hes panies: M. S. Thesis, George Mason University, Fairfax LELLINGER, ch a 1985. A field manual of the Lone in fern-allies of the ate States and Canada. sonian Institution Press, Washington Li, + AND cee H. HAvuFLer. 1999. Genetic variutio n, cearte systems, and oe of diversifica- mn in io lyorcca Polypodiam (Polypodiaceae). Syst. Bot. 24:339-35 Liew, Ky Ss: . Studies on North American Athyrium (L.) Roth. Ph.D. Tau: Columbia ene . 1972. Numerical taxonomic studies on North American Lady Ferns and their allies. Tai- nia 17:190-221. ions B. 1971. Contribution a la connaissance de la morphogenése et de la structure me parois ages chez les filicinées isosporées. Thesis of Université Paul Sabatier de Toulou 976. Sur la structure fine de l’exospore dans les devers groups de Sat udicie ero (microspores et isospores). Linn. Soc. Series 1:231-2 — Sur la formation du sporoderme chez Psilotum triquetram Sw. (Psilotaceae). Grana 165; Manton, ‘4 pn Problems of cytology and evolution in the pteridophytes. Cambridge University Press, London Mayr, E. he Stomakies and the origin of species. Columbia University Press, NY. . 1992. A local flora and the biological species concept. Amer. J. Bot. 79:222-238. orp ag . 1998. Molecular ones at hybrid swarms in the ie enue Polystichum aie Syst. Bot. 23:421-4 Murpny, R. W., J. W. Sires, Jr., D. G. BUTH, AND . HAUFLER. 1990. sn Isozyme electro- phoresis. Pp. pees in D. Hillis and C. Maite, eds. Molecular Systematics. Sinauer, Sunderland, MA. Nel, M. 1978. Estimation of Pig oe and genetic distance from a small number of individuals. Genetics 89:583— Paris, C. A., F. S. WAGNER, AND in “ Wacner, JR. 1989. Cryptic species, species delimitation, and taxonomic practice in the homosporous ferns. Amer. Fern J. 79:46—54. Pete J. S. 1972. Measures of genetic seep and genetic Rieti Studies in Genetics, University of Texas Publication 7213:145-15 SCHNELLER, J. J. 1979. a a investigations | on the lady fern (Athyrium filix-femina). P. ma Evol, 132:255- 989. Remarks on heer regulation of — wall pattern in intra- and interspecific crosses of 2 . Bot. J. Linn. Soc. 99:115— ND B. W. Scumip. 1982. lnvetigation on the inspec variability in Athyrium filix- ‘eats L) Roth Bull. Mus. Natl. H oe B. 3—4:215-— SELANDER, R. K., M. H. Situ, S. Y. YANG, W E. JOHNSON, wien 'B. GENTRY. 1971. Biochemical ii ta and systematics in the genus Peromyscus. I. variations in the old-field 90. 7 Tennessee with the fern allies exchndad. George Peabody College for Teachers, Nashville, IN. SMaLL, J. K. 1938. Ferns of the southeastern United States. The Science Press Printin Lancaster, PA. Soitis, D. E., C. H. Haurter, D. C. Darrow, anv G. J. GASToNny. 1983. Starch gel electrophoresis of ferns: A compilation ‘of grinding buffers, gel and electrode buffers, and staining schedules. Amer. Fern J. 73:9—27. Soxtis, P. S., ALVERSON. 1987. Electrophoretic and morphological confirma- tion of interspecific hybridization between Polystichum kruckebergii and P. munitum. Amer. Fern J. 77: g Company, KELLOFF ET AL.: EASTERN NORTH AMERICAN ATHYRIUM 213 Soxtis, D. E., P. S. Soitis, AND P. G. Wor. 1990. Allozymic divergence in North American Poly- stichum (Dryopteracoan) Syst. Bot. 15:205-215. Soxtis, P. S., D. E. Soitis, AND K. E. Housincer. 1988. Estimates of shy ares aa selfing and interpopulational te me in ng aneen ferns. Amer DP. 1. Allozymic and chloroplast DNA ia Hil of Polyploiy i in Polstican pence irraneret) I. The origins of P. californicum and P. scopulinum. Sys SPEER, W. I., C. R. WertuH, AND K. W. Hiu. 1998. Relationships between two infraspecific taxa of ache poh ate tDennstaedicene II. Isozyme evidence. Syst. Bot. 23:313-326. SworrorpD, D. AND R. SELANDER. 1981. BIOSYS-1. A FORTRAN program for the wae olrelig analysis of a onttiey ne data in population genetics and systematics. J. Hered. 72:281—283. Tryon, A. F. 1986. Stasis, diversity and function in spores based on an electron annem sur- vey of the Pteridophyta. Pp. 233-249. in S. Blackmore and I. K. Ferguson, eds. Pollen and spores: form and function. No 12. Academic Press, London AND B. LuGaRDON. 1991. Spores of the pteridophyta. ea wall structure, and diversity based on electron microscope studies. Springer-Verlag, N Wacner, W. H., Jr. 1943. Hybridization by remote control. Amer. Fern J. 33:71-73. bier A biosystematic survey of United States ferns - coat Se ti eh fi Amer. Fern J. 53:1- D . WaGNER. 1966. pla ale of sig ety Lake area, Giles Co., Virginia: bio- ae lige: 1964-6 anea 31:1 Wa tker, T. G. 1 Dect in some conn peas L. Evol. 12:82-92. WEATHERBY, C. ‘ a A list of varieties and forms of the ferns of ana North America. Amer. WenpeL, J. F. anp N. F. WEEDEN. 1989. Visualization and interpretation of plant isozymes. Pp. 5-45 in D. E. Soltis and P. S. Soltis, eds. Isozymes in plant biology. Dioscorides Press, Portla oa OR. WERTH, . as 1985. Implementing an isozyme laboratory at a field station. Va. Je Sci. 36:53-76. 0. rani Pre-prepared frozen isozyme assays. Isozyme Bull. 09. WHERRY, a T. 1948. Remarks on the American Lady Fern. Amer. Fern J. 38:155-158. 61. ies fern guide. Northeastern and midland central states and adjacent Canada. ‘dilate and Co., Garden City Wricut, S. 1965. The Beiasryp ere of ee ee structure by F-statistics with special regard to systems of mating. Evol. 19 . 1978. Evolution and tha onan of populations, as 4. Variability within and among netarel populations. University of Chicago Press, Chic American Fern Journal 92(3):214—228 (2002) A New Hybrid Polypodium Provides Insights Concerning the Systematics of Polypodium scouleri and its Sympatric Congeners T. J. HILDEBRAND, CHRISTOPHER H. HAUFLER, JAMES P. THERRIEN, and CatHy WALTERS Department of Ecology and Evolutionary Biology, University of Kansas, Lawrence, Kansas 66045-2106 PuiLtip HAMMOND University Herbarium, University of California, Berkeley, CA 94720 AsstRACT.—With its thick, leathery leaves, reticulate venation, and large sori, Polypodium scou- leri, located in a narrow band along the Pacific coast of North America, is the most distinctive member of the cosmopolitan P. vulgare species complex. Although early studies based on mor- and trnL DNA sequences with isozymic analyses suggested that P. scouleri originated relatively recently and is closely allied to and sympatric with P. californicum and P. glycyrrhiza. Consis- The Polypodium vulgare L. complex (Polypodiaceae) first drew attention with the publication of cytological counts for several European and North American members (Manton, 1950). Controlled crosses among members of the complex followed (Shivas, 1961), and resulted in further taxonomic stud- ies founded on interbreeding boundaries and subtle morphological distinc- tions. A major organizational leap in the circumscription of North American Polypodium L. species was the definition of eastern and western complexes (Lloyd and Lang, 1964). In their search to identify parental lineages for allo- polyploid members, Lloyd and Lang grouped the complexes on the presence (eastern) or absence (western) of sporangiasters. Although sporangiasters were later shown to constitute synapomorphies in several western species (P. amorphum Suksdorf, P. saximontanum Windham, and P. sibiricum Siplivinsky [Haufler and Windham, 1991]), recognizing the importance of ese unique soral features within American polypodiums was insightful. Additionally, Lloyd and Lang (1964) hypothesized that the progenitors of the allotetraploid P. californicum Kaulfuss were P. glycyrrhiza D. C. Eaton (2n) HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM 215 and diploid P. californicum Kaulfuss. It was not until much later, however, that Whitmore and Smith (1991) described the tetraploid cytotype as a dis- tinct and reproductively competent species, P. calirhiza S. A. Whitmore & A. R. Smith. Research explorations of morphology (Haufler and Windham, 1991; Whitmore and Smith, 1991; Haufler et al., 1993), cytology (Haufler and Wang, 1991), isozyme variation (Haufler et al., 1995b), chloroplast DNA restriction site analysis (Haufler et al., 1995a), and DNA sequence data (Haufler and Ranker, 1995) have further defined relationships within the Polypodium vulgare complex. Yet a review of the above reveals the exclu- sion of P. scouleri Hooker & Greville from all but two studies (Haufler et al., 1993; Haufler and Ranker, 1995). Confined to a narrow distribution along the Pacific Coast of North America, and tentatively recognized as a member of the western complex, the dis- tinctive morphology of P. scouleri clearly separating it from its congeners precluded the formulation of accurate hypotheses about phylogenetic rela- tionships. The overview by Haufler et al. (1993, pp. 315-323) in their treat- ment of Polypodium in Flora of North America (Fig. 1) allied P. scouleri with other western members, and provided a detailed morphological de- scription. In addition, the investigation of rbcL sequence data suggested a sister taxon relationship with P. glycyrrhiza that was supported by isozyme profiles (Haufler and Ranker, 1995). Haufler and Ranker (1995) hypothesized that the particularly distinctive morphological features of P. scouleri may have evolved through adaptive response to environmental stress. That study did not consider another close relative, P. californicum, and, because only diploids were included, did not incorporate allopolyploid P. calirhiza. Recently, leaves resembling P. scouleri but having some atypical features were collected at a site in California. At the same time, leaves of P. calirhiza and more typical P. scouleri were obtained. The present study was designed to examine the atypical leaves morphologically and use molecular ap- proaches to determine whether these plants originated through hybridization between P. scouleri and other members of the western complex. Several mor- phological features of suspected hybrid leaves were investigated and com- pared with features of typical P. scouleri and P. calirhiza leaves. Starch gel electrophoresis was used to reveal isozyme marker alleles and characterize the suspected hybrid and parental lineages. MATERIAL AND METHODS Stupy ArEA.—Located within the confines of highly populated San Francisco County are the natural vegetation preserves of Tank Hill and Mt. Sutro (Fig. 2). The Open Space Program of the San Francisco Recreation and Parks Department retains ownership of these sanctuaries. Land stewards manage vegetation of the preserves, emphasizing enhancement and restoration of native species as well as the removal of invasive exotic plant populations. 216 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) P. scouleri P. californicum P. glycytthiza P. hesperium P. amorphum P. saximontanum P. sibiricum Fic. 1. Kinship of diploid and tetraploid members of the Western Polypodium vulgare complex of P. sibiricum, which is circumboreally distributed in North America and Asia. Shaded ovals represent sterile backcrosses (3n) in the complex. The dashed line between P. scouleri and P. calirhiza indicates the subject of the present study. A dense forest of mature, non-native cypress (Cupressus macrocarpa Hartw. Ex Gord.) and eucalyptus (Eucalyptus globulus Labill.) was planted on Mt. Sutro as early as 1870 and now covers much of the hill. Polypodium species on Mt. Sutro flourish beneath the forest canopy, nestled among rocks and at the bases of trees as hemiepiphytes. Leaves of Polypodium calirhiza and P. scouleri were collected from the east face of Mt. Sutro (Site 1, Fig. 2; Table 1). Eucalyptus and cypress were also planted on Tank Hill, but, in contrast to Mt. Sutro, they are sporadic on Tank Hill, and primarily at lower elevations. The summit of Tank Hill is dominated by large tracts of rocky fields and ledges of Franciscan radiolarian chert. In the fields and on the ledges are Polypodium populations exposed to the harsh sun and buffeted by gusting winds. Native species associated with ferns in this predominantly open habitat include: Nootka reed grass (Calamagrostis nutkaensis (Presl) Steud.), yarrow (Achillea millefolium L.), coast barberry (Berberis pinnata Lag.), soap plant (Chlorogalum pomeridianum (DC.) Kunth. var. divaricatum (Lindl.) Hoov.), and seaside daisy (Erigeron glaucus Ker.). Among the boulders and crevices near the crest of Tank Hill, leaves of Polypodium calirhiza and HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM 217 San Francisco County Fic. 2. Collection localities on Mt. Sutro and Tank Hill, San Francisco County, California. P. scouleri, as well as the suspected hybrid, were collected (Site 2, Fig. 2; Table 1). To characterize the amount of electrophoretically detectable genetic variation across the range of P. scouleri, population samples were obtained from four additional California populations. Representative specimens of all 218 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) TaBLE 1. California sites from which Polypodium plants were collected. All specimens are deposited in the McGregor Herbarium, University of Kansas (KANU). N = number of individuals for each population. Collection Site (county) Species (ploidy level: x = 37) N Voucher ~ Mt. Sutro (San Francisco) calirhiza (4x) Mt. Sutro (San Francisco) P. scouleri (2x) woo as a Sf Dm OD Sse Q 8 55 a8 Go NR hd&S —~ me On Tank Hill (San Francisco) P. calirhiza (4x) 12 Hildebrand 3219 Tank Hill (San Francisco P. scouleri (2x) 7 Hildebran Tank Hill (San Francisco) P. calirhiza X scouleri (?) 5 Hildebrand 3220 & 3221 Fern Canyon (Trinity) P. scouleri (2x) 20 ~=Therrien s.n. Point Reyes National P. scouleri (2x) 15 Therrien s.n. Seashore (Marin Trinidad (Humboldt) P. scouleri (2x) 6 Therrien s.n. Fort Ross (Sonoma) P. scouleri (2x) 3. ‘Therrien s.n. collections (Table 1) were pressed and deposited at McGregor Herbarium of the University of Kansas (KANU). MeETHODS.—Leaves were removed from live material in the field, placed in plastic bags, and shipped on ice. Upon arrival, bags were transferred to 4°C storage where they were kept until preparation for electrophoresis. At least one sample of plant material for each leaf was prepared immediately upon receipt in the laboratory. Most leaves, particularly of P. scouleri and the sus- pected hybrid, retained their fresh appearance in storage for considerable time (up to one month). Preparations from fresh plant material stored for longer periods yielded banding patterns comparable to that prepared imme- diately, indicating extended retention of enzymatic activity. Plant material was prepared by crushing in phosphate-PVP buffer (Soltis et al., 1983) followed by absorption of the homogenate into filter paper wicks and storage of the wicks at —80° C. Freezing prepared fresh material allowed storage of samples for several months with no loss of enzymatic activity. Selection of enzymes and the systems best suited for their survey was based on previous studies of Polypodium (Haufler et al., 1995a, 1995b) and other fern genera (Haufler, 1985b; Haufler et al., 1990: Soltis et al., 1990; Pryer and Haufler, 1993). Banding patterns were obtained by electrophoresis on 12.4% starch gels for the following enzymes: aldolase (ALD), fructose 1,6- bisphosphatase (FBP), glyceraldehyde 3-phosphate dehydrogenase (G-3PDH), hexokinase (HK), isocitrate dehydrogenase (IDH), phosphogluconate dehy- drogenase (PGDH), triosephosphate isomerase (TPI), and phosphoglucoiso- merase (PGI). Bands were resolved best for ALD and IDH on system 11 (Haufler, 1985b) whereas only the 7.5 pH version of the morpholine/citrate system (MC) (Clayton and Tretiak, 1972) revealed clear bands for G-3PDH. Both system 11 and MC resolved bands for FBP, MC and system 8 (Haufler, 1985a) for MDH, and MC and system 6 (Soltis et al., 1983) for PGDH. System 8 also revealed bands for HK, LAP, and, in addition to system 6, for PGI. TPI bands were revealed only with system 6. Digital images were obtained of all HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM TABLE 2. and P. calrhiz 219 or aan of character states identified for Polypodium scouleri, calirhiza X scouleri, Character Polypodium scouleri Polypodium calirhiza X scouleri Polypodium calirhiza Frond texture Rhacis scales Pinna venation Guard cells - shape - mean length: uum (range) peponinno cell argins Paces ate stomata density Spore length: um (s.d.) - distribution - size stiff, leathery broadly ovate, tapering to ca 3 cells in width; not occurring in pairs regularly anastomosing, forming one row of areoles round 35 (33-38) smooth 150/mm? 61.3 (+ 8.4) round or oblong; closest to midrib generally > 3 mm stiff, leathery narrowly ovate to sear lanceolate; forming areoles round 38 (35-43) distinctly lobed 70/mm* malformed round or oblong; both shapes on individual plants variable herbaceous lanceolate to lanceolate-ovate; 3-6 cells wide, tapering to 1-3 cells; not occurring in pairs free, no areoles med elliptic 56 (48-63) distinctly lobed 40/mm* Bib (= 71) oblong midway between costa and margin generally < 3 mm variable, 14 mm, to sent gels using a Nikon CoolPix 950 camera and visualized with Adobe Photo- shop 5.5 software for Macintosh. Morphological characters documented as useful for delimiting Polypodium species (Haufler et al., 1993; Whitmore and Smith, 1991) were examined on leaves from both parental species and the hybrid. These included leaf tex- ture, sorus diameter, pinna venation, spore length, and rachis scale width. In addition, lower surface (abaxial) epidermal peels were produced to measure sizes and characterize shapes of guard and subsidiary cells, as well as differ- ences in stomatal density (Table 2). RESULTS MorpHoLocy.—The leaves of Polypodium scouleri differ from those of other members of the genus (although leaves of P. calirhiza may become somewhat thickened in exposed coastal environments) by their leathery texture, and 220 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) mean = 61.5 mean = 61.3 70 | s.e.= 7.1 se. = 8.4 mean = 55.3 s.e. = 6.2 Spore length (micrometers) 45 = P. calirhiza P. scouleri FP. scouleri (CA) (CA) (OR) n= 2 = 2 n= x = 205 x = 189 “2 55 Fic. 3. Comparisons of spore length from plants gathered at the Mt. Sutro and Tank Hill Poly- podium populations, and from a P. scouleri population from Oregon. Vertical bars represent +/— one standard error. Expected mean spore lengths (Haufler et al., 1993) for P. calirhiza and P. scou- leri are > 58 wm and < 53 um, respectively. n = number of plants sampled; x = number of spores measured. their well-defined aeroles produced by anastomosing venation. Other dis- tinctive morphological features include individual leaf segments that are greater than 12 mm in width, soral diameters of more than 3 mm, and rachis scales that are large, pale reddish brown and broadly triangular, tapering to a point less than three cells wide (Table 2). The pair of guard cells surrounding the stomata on the lower epidermal surface is circular in outline; individual guard cells average 35 jm in length, and are adjoined by smooth-margined subsidiary cells. Stomata density is approximately 150 stomata/pm’. Tank Hill and Mt. Sutro P. scouleri populations had a mean spore length of 61.3 um (Fig. 3). Polypodium calirhiza plants have herbaceous leaves lacking the leathery texture of P. scouleri and are, in general, of smaller stature. Pinna venation is free or weakly anastomosing with some to many segments lacking aeroles. Leaf segments seldom exceed 12 mm wide, and soral diameters are less than 3 mm. Rachis scales, as in P. scouleri, are a translucent, pale reddish brown, HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM 221 but, in contrast to P. scouleri, are lanceolate, only a few cells wide proxi- mally, and narrow to only one to three cells distally. Epidermal peels showed elliptic paired guard cells (average length: 56 wm) and subsidiary cells with distinctly lobed margins. In striking contrast to that observed for P. scouleri stomata, stomata density is approximately 40 stomata/mm‘’. Spores of the California P. calirhiza plants averaged a mean length of 61.5 um. The putative hybrid individuals showed both hybrid vigor (Charlesworth and Charlesworth, 1987) and dwarfing of leaves. On nearby Mt. Davidson, a more sheltered locale, P. scouleri produced leaves greater than 70 cm in length and lush in appearance. Polypodium calirhiza morphotypes vary with exposure, and increasingly open areas produce plants of more diminutive stature. Morphological character states observed for hybrid leaves were either i) a combination of discrete features inherited without change from each parent, or ii) an additive blending of parental traits resulting in intermediate mor- phological states (Table 2). An exception was the consistently leathery, stiff texture of hybrid leaves, a phenotype that resembles only the P. scouleri parent. Leaves of the hybrid never had the herbaceous texture of the P. cali- rhiza lineage. Pinna venation of hybrid leaves incorporates features of both parents: most veins are free, but occasional anastomoses and areoles occur. Soral size and development are extremely variable on hybrid leaves and of three general categories. Sori were 1) as large or slightly larger than those typical of P. scouleri (> 3 mm), 2) smaller and resembling P. calirhiza sori (< 3 mm), or 3) entirely undeveloped. Translucent, reddish brown scales occur along the rachis of hybrid leaves, but, in contrast to the deltate and lanceolate parental scales (P. scouleri and P. calirhiza, respectively), rachis scales on hybrid leaves are narrowly triangular. In addition, adjacent rachis scales are often fused (from the base) for approximately one third their length. Paired guard cells were orbicular in outline, as observed for P. scouleri, whereas subsid- iary cells showed the distinctive, lobed margins of P. calirhiza. The average length of guard cells (38 ym), and stomatal densities of approximately 70 stomata/mm are intermediate between parental lineages. All spores on hybrid plants were shrunken and malformed, suggesting inviability. MoLEcULAR.—Previous isozymic work by Haufler et al. (1995b) investigated P. californicum and P. glycyrrhiza (the progenitors of P. calirhiza) and iden- tified electrophoretic markers for each diploid member, based on sampling the infraspecific variation from eleven populations. In the present study, iso- zyme profiles of the Tank Hill and Mt. Sutro individuals of P. scouleri were verified as representative for the species by sampling other populations (Table 1). In contrast to the genetic variability detected for other Polypodium diploids, P. scouleri isozymes were monomorphic across all populations and for all enzymes sampled. Of the ten enzyme systems considered, four (HK, PGI, PGDH, and MDH) yielded reproducible, well resolved banding patterns 222 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) that discriminated individuals representing parental species and hybrids. Allozymes (allelic variants within loci) are distinguished from isozymes (products from different loci) by assuming that models of gene product com- partmentalization (Gastony and Darrow, 1983) are applicable to Polypodium enzymes. The following results were obtained for each of the applicable enzyme sys- tems. Gels stained for the monomeric enzyme hexokinase (HK) showed a slow-migrating allele in P. calirhiza samples in contrast to the faster allele present in P. scouleri. The hybrid expressed both alleles, one from each parental species (Fig. 4a). Two isozyme loci were resolved for phosphogluco- isomerase (PGI) and phosphogluconate dehydrogenase (PGDH). The faster migrating isozymes (Pgi 1, Pgdh 1) for both enzymes were monomorphic for all samples investigated. In contrast, banding patterns for the slower migrat- ing isozymes (Pgi 2, Pgdh 2) were more variable for each enzyme (Fig. 4b & 4d). Polypodium calirhiza possessed a slower migrating allozyme for Pgi 2, whereas P. scouleri revealed a faster allozyme. The hybrid exhibited an addi- tive banding pattern for dimeric phosphoglucoisomerase, possessing both allelic variants (Fig. 4b). A similar pattern was observed from gels stained for phosphogluconate dehydrogenase (PGDH). The polymorphic locus (Pgdh 2) expressed only the faster migrating allele in P. calirhiza, both allelic variants in P. scouleri, and an additive banding pattern in the hybrid (Fig. 4d). Three isozymes were revealed for malate dehydrogenase of which two (Mdh 2, Mdh 3) were monomorphic across all samples. The polymorphic locus (Mdh 1) produced bands that represent a fast allele for P. calirhiza and a slower migrating allele present in P. scouleri samples. In addition, both intra- and inter-locus heterodimers were formed during electrophoresis and were subse- quently revealed by staining for malate dehydrogenase. Bands produced by the formation of both intra- and inter-locus heterodimers further supported the hybrid pattern of additive banding from parental species (Fig. 4c). DISCUSSION Hybridization between species is a frequent phenomenon in plants and is especially common in pteridophytes (Wagner, 1968). The clarity and preci- sion of species recognition and the accuracy of phylogenetic hypotheses can be enhanced by identifying and characterizing naturally occurring hybrids. Especially in groups such as the Polypodium vulgare complex, where spe- cies differences are particularly subtle, unrecognized interspecific hybrids that usually blend features of the parental individuals can appear to bridge gaps between otherwise distinct species. Zymograms were exceptionally informative for delimiting P. scouleri and P. calirhiza, and for unequivocal verification of hybrid leaves collected on Tank Hill and Mt. Sutro. Typically, additive banding patterns indicate the presence of both parental alleles and are observed in hybrids (Crawford, 1990; Murphy et al., 1996). The additive banding patterns visualized on gels HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM 223 Pgi2 Mdh1 Mdh2 Mdh3 Pgdh2 Fic. 4. Representative gels stained for enzymes from bis seca —— included in present study. P. calirhiza = ca; P. calirhiza x scouleri = casc; P. scouleri = sc. eee HK); B. ee (PG I); C. malate phen (MDH); wad D. phosphogluco- nate dehydrogenase (PGDH). Sampling in B. & C. corresponds to species as labeled in D. See text for interpretation of banding patterns. 224 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) stained for enzymes HK, PGI, PGDH, and MDH combine and confirm paren- tal contributions from the P. scouleri and P. calirhiza genomes to the hybrid. The difficulty encountered in developing equivalently distinctive morpho- logical characterizations of the hybrid individuals requires further discus- sion. Hybrids are often recognized initially because they have morphological peculiarities that can signal the amalgamation of two distinct genomes. Wagner (1962) reviewed deviations from morphological symmetry and their role as indicators of hybrid origins in ferns. For genera studied (Asplenium, Cystopteris, Cheilanthes, Osmunda, Polystichum, Pteris, Woodsia), he devel- oped three broad conclusions regarding hybrid morphology: First, hybrid structures form symmetrically, but blend traits from parental lineages. If large differences occur between parental lineages, hybrids tend toward irreg- ular or asymmetric development. Second, when asymmetric development does occur, it is retained in hybrids, and may be useful in identifying the hybrid individual. Third, the discovery of morphological irregularities should key investigators to the possibility of hybrid origins and further study of possible parental lineages. Thirty years of further investigation of these fern genera (e.g., Moran, 1982; Murakami et al., 1999; Yatskievych et al., 1988; Mickel, 1979; Haufler et al., 1990), in addition to other pteridophytes (e.g., Montgomery, 1982; Palmer, 1998; Pryer and Haufler, 1993; Tyron, 1968) support the conclusions on hybrid morphology summarized by Wagner. Likewise, our study reported intermediate, blended traits in P. cali- rhiza X scouleri that are consistently expressed in all leaves, and that help to resolve the definition of the hybrid and the identification of parental spe- cies. In comparison, other character states were not intermediate between pa- rental lineages, or were so highly variable that they could not contribute to the definition of P. calirhiza < scouleri (Table 2). The loss of reticulate venation found in Asplenium and Polystichum (Wagner, 1962), and other Polypodium hybrids (Whitmore and Smith, 1991) is also observed in P. calirhiza < scouleri. Whitmore and Smith (1991) in- vestigated other members of the western P. vulgare complex, and revealed a loss of vein anastomosis following hybridization. They observed only 0-33% of the veins per pinna in Polypodium calirhiza anastomose, whereas paren- tal species P. californicum has weakly to fully anastomosing venation (5— 100%) and P. glycyrrhiza produces pinna with entirely free venation. The primarily free venation occurring in P. calirhiza x scouleri provides further evidence for a propensity toward less reticulated venation whenever ge- nomes differing in this character are present. Position of fertile pinnae on the leaf (terminal, mid-, lower), often a combi- nation of parental features in fern hybrids (e.g., Osmunda hybrids, Wagner, 1962), was not a useful diagnostic character for P. calirhiza X scouleri. Poly- podium scouleri and P. calirhiza fertile segments tend to be positioned ter- minally in the former, and may comprise all but the lower 1-3 segment pairs in the latter. Nonetheless, large variation occurs on parental leaves, particu- larly in P. calirhiza. Likewise, hybrid leaves show great variation ranging from few, often terminal segments with sori, to all segments producing sori. HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM 225 However, the large variation in soral maturation on segments aptly indicates hybridization, with sori found in all developmental stages, albeit with mal- formed spores. Fertile vein development and soral position frequently aid in delimiting fern hybrids. For example, Polystichum lonchitis produces fertile veins prog- ressing from the midrib to the margin with sori midway and “dorsal” upon the vein whereas P. acrostichoides has fertile veins terminating at sori half- way between the margin and the midrib. A combination of parental traits occurs in fertile vein development of the hybrid P. acrostichoides x lonchitis with some sori terminal on fertile veins, other sori dorsal, and some sori lacking development (Wagner, 1962). In contrast, no distinction in soral position between P. scouleri and P. calirhiza, as well as the hybrid (when it occurs), is observed. Sori are terminal on fertile veins of parental species and P. calirhiza < scouleri. Close observation of parental species failed to determine differences in the manner by which fertile veins terminated, and, although P. calirhiza veins appear to end in a more reduced and less club- like form than those of P. scouleri, this difference may merely result from differences in leaf texture. Leathery hybrid leaves produce fertile veins more closely resembling P. scouleri, but, again, this may be an artifact of similar leaf textures. Sori do occur closest to the costa in P. scouleri whereas they are located midway between the costa and margin on P. calirhiza pinnae. The hybrid is highly variable producing sori both near the costa or midway between it and the pinna margin. Leaf outlines of hybrids often blend those of parental individuals (Wagner, 1962), but this morphological feature is not transitional in P. calirhiza X scouleri leaves. Texture, leaf outline, pinna width and apex, as well as sinus angle and depth are all commonly useful for hybrid identification, but are not useful in delimiting P. calirhiza X scouleri. The resemblance of hybrid leaves to P. scouleri may be the most significant factor contributing to past difficulties in recognizing hybrid populations. An unexpected discovery of the present study regards the average spore length for sampled P. scouleri plants from California. Published average spore length (Haufler et al., 1993) for P. scouleri is less than 53 um, whereas P. calirhiza spores exceed 58 1m. Indeed, P. calirhiza spores from Tank Hill and Mt. Sutro measured well within expected values whereas P. scouleri spores exceeded the expected average (Fig. 3). In an effort to explain the increased spore size, P. scouleri spores were harvested and measured from plants in Oregon (Hildebrand #3214, KANU), and, with an average of 55.3 uum, fell within the expected range. Three explanations may account for the larger than average spores obtained from the California P. scouleri plants. Certainly, contamination of spore samples may have occurred. Although spores were removed directly from sporangia, P. calirhiza spores from dried material may have contami- nated the P. scouleri plants measured. Two alternative possibilities are more difficult to assess. Environmental factors could account for the increase in average spore length found in California populations of P. scouleri. Spore 226 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) sizes of Isoetes species have been found to be strongly affected by environ- mental parameters including temperature, solar radiation, and elevation (Cox and Hickey, 1984). Finally, the increased average spore length observed in plants of P. scouleri from California could be explained by an increase in chromosome number. Although not always correlated to ploidy level, in- creases in spore size may indicate the formation of both auto- and allopoly- ploids. For example, average spore length was found to increase with ploidy level by a multiplier of 1.26 in species of neotropical Polystichum (Barring- ton et al., 1986). The increase in average spore length for sampled California P. scouleri plants (ca. 15%) may correspond to an increased chromosome number via autopolyploidy. To further explore this possibility, guard cells, often positively correlated with ploidy level (Barrington et al., 1986), were compared between Oregon and California P. scouleri plants. Guard cell size was previously determined to predict allopolyploidy in the western P. vulgare complex. Barrington et al. (1986) found tetraploid P. calirhiza guard cells average 1.12 and 1.2 times larger than progenitors P. californicum and P. glycyrrhiza, respectively. Nevertheless, differences in guard cell length between California and Oregon P. scouleri populations were not observed, and decreased the possibility of an autopolyploid event in P. scouleri from sampled California populations. Early cytological investigations by Manton (1951) of three plants identified as P. scouleri from Point Reyes, California revealed n pairs and n univalents at meiosis (vs. typically unpaired chromosomes in triploids) and suggested a closer genetic relationship between Polypodium tetraploids and diploids in western North America. It now seems most likely that the California plants Manton investigated were not P. scouleri, but P. calirhiza x scouleri. No cytological study has been completed for P. scouleri plants from Mt. Sutro and Tank Hill populations, and efforts to re-locate the Point Reyes hybrid populations have been unsuccessful. The present study provides morphological and molecular evidence for the hybridization of Polypodium scouleri and P. calirhiza in California, and helps to secure the placement of the former in the western P. vulgare com- plex of North America. Questions remain regarding the California P. scouleri populations that merit further investigation. Future cytological studies of these populations, in conjunction with spore length measurements from fresh material, may aid in clarifying any remaining ploidy level conun- drums. Polypodium calirhiza x scouleri Stem stout (5-15 mm diameter), occasionally whitish, acrid to slightly sweet-tasting. Rhizome scales uniformly brown to weakly bicolored with pale margins, lanceolate to lanceolate-ovate, symmetric, with occasional teeth on erose margins. Blades to 39 cm in length with a stout petiole to 3 mm in diameter. Lamina stiff and leathery, ovate-lanceolate, pinnatifid, usually widest at or just above the base, to 20 cm wide: sparsely scaly to HILDEBRAND ET AL.: A NEW HYBRID POLYPODIUM 227 abaxially glabrescent; rachis sparsely puberulent adaxially. Rachis scales, concolorous, pale reddish brown, lanceolate-ovate; often adjacent, fused (from the base) to one third their length. Segments oblong to linear, usually more than 12 mm wide, with rounded apices, sparsely crenulate margins, midribs adaxially sell Venation primarily free but with some anastomo- ses and irregularly formed aeroles. Sori oval to circular, but with widely varying development, usually closer to midrib than margins, 1-4 mm in di- ameter, producing malformed spores. Sporangiasters absent. ACKNOWLEDGMENTS We thank R. Cranfill for initial field recognition of atypical leaves, and A. R. Smith for his pre- liminary cytological examination of collected material. LITERATURE CITED BARRINGTON, D. S., C. A. Paris AND T. A. RANKER. 1986. Systematic inferences from spore and sto- mate size in ferns. Amer. Fern J. CHARLESWORTH, D. AND B, CHARLESWORTH. 1987. Inbreeding depression and its evolutionary conse- quences. Ann. Rev. Ecol. Syst. 18: Cayton, J. W. AND D. N. TRETIAK. 1972. ke ae buffers for pH control in starch gel electro- gan J. Fish. Res. Board of Canada 29:1169-1172. Cox, P. A. aNp R. J. Hickey. 1984. Convergent megaspore evolution and Isoetes. Amer. Naturalist i nei Crawrorp, D. J. 1990. Plant molecular systematics: macromolecular approaches. John Wiley & Sons. New fee New Yor Gastony, G. J. AND D. C. Darrow. 1983. arg ere and cytosolic isozymes of the homosporous Athy ae -femina L. Amer. J. Bot. 70:1409-1415. preeny C. H. 1985a. Enzyme variability and sas of evolution in Bommeria (Pteridaceae). Syst. Bot. 10: aC 104 . 1985b. Pteridophyte evolutionary biology: the electrophoretic approach. Proc. Roy. Soc. Edinbur gh 86:3 eres po RbcL sequences provide apa insights among sister spe- cies ste iis forn genus lige um. Amer. Fern J. 85:361—-374. D. E. Sottis aNp P. S. Sotis. 1995a. Phylogeny af tha Polypodium vulgare complex: in- hts oe sae nan restriction site data. Syst. Bot. 20:110-11 R. Wanc. 1991. Chromosomal analyses and the contest of allopolyploid Polpodium cate (Polypodiaceae). Amer. J. Bot. 78:624-629. WINDHAM. . New species of North American gi and Polypodium, ae enonidents on their reticulate relationship. Amer. Fern J. 81 , M. D. WinpHaM, F. A. LANG AND S. A. WHITMORE. 1993. ee. Pe 315-323 in Flora of North America Editorial Committee. Eds. Flora of North America north of Mexico. Oxford Ache Press, New Yor M. INDHAM AND E. W. ee “1995b. Reticulate evolution in the Polypodium vulgare compl = dd Bot. 20:89-10 WINDHAM AND T. A. te 1990. an analysis of the Cytopteris tennes- seensis = complex Ann. Missouri Bot. Gard. 77:314—329. Lioyp, R. M. anp F. A. Lanc. 1964. The Priel tg vulgare complex in North America. Brit. Fern Gaz. es 168-17 Manton, I. 1950. Polpoodiud vulgare. Chapter 8 in Problems of trad and evolution in the Pteridophyta. Cambridge University Press, Cambridge, Great Bri ——. 1951. The cytology of Polypodium in America. Nature 167: ns 228 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) eer P 1979. The fern genus Cheilanthes in the continental United States. Phytologia —437. eee J. D. 1982. Dryopteris in North America. Part II. The hybrids. Fiddlehead Forum anada. Eee Fern J. 72:5— MurakaMI, N., S. Nocami, M. WATANABE AND K. IWATSUKI. esi i ioe of Aspleniaceae inferred from rbcL nucleotide sequences. Amer. Fern J. 89:232— Murpuy, R. W., J. W. Sires Jr., D. G. Bur AND C. H. HAuFLer. ou sel isozyme electropho- resis. Pp. 51-120 in D. M. Hillis, C. Moritz, and B. K. Mable. Eds. Molecular Systematics (2nd Edition). Sinauer Associates, Inc., Sunderland, MA Pater, D. D. 1998. Cibotium x heleniae hyb. nov. (C. chamissoi < C. menziesii; Cyatheaceae): a eres occurring hybrid from Oahu, Hawaii. Amer. Fern J. 88:150-154. Pryer, K. M. AND C. H. HAur.er. 1993. Peete and chromosomal evidence for the allote- ape ere of the common oak fern Gymnocarpium dryopteris (Dryopteridaceae). Syst. Bot. 18 72 Suivas, M. G. 1961 disesag to the cytology oo apa of Polypodium in Europe and America. II. Taxonomy. J. n. Soc., Bot. 58:2 So.tis, D. E., C. H. HAvuF.er, D. C. as ow AND G. J. i ASTONY. 1983. Starch gel electrophoresis in erns; a counjeilatien: of grinding buffers, gel and electrode buffers, and staining schedules. Amer. Fern J. 7 27 SOLTIS, S., D. E. SO $ AND P. G. Wo r. 1990. oe divergence in North American Polysti- m(Dryoperdacoae Syst. Bot. 15:205-215 Tyron, “ F, Comparisons of sexual and inant: races in the fern genus Pellaea. Rhodora o “a Wacner, W. H. 1962. Tragulae morphological development in hybrid ferns. Phytomorphology 12:87-100. . 1968. Hybridization, taxonomy, and evolution. Chapter 9 in V. H. Heywood. Ed. Modern nuethods in _ taxonomy. Academic Press, London. Wuitmokg, S. A. AND A. R. SmiTH. 1991. Recognition of the noe Polypodium calirhiza (Poly- podiaceae), in western North America. Madrofio 38:233-24 YATSKIEVYCH, G., D. G. STEIN AND G. J. GASTONY. 1988. pet lata DNA evolution and Plasto . rhea (Dryopteridaceae) and related fern genera. Proc. Natl. Acad. Sci. U.S.A. 89-25 American Fern Journal 92(3):229-238 (2002) Comparative Research of Gametophytes of Olfersia alata and Olfersia cervina (Dryopteridaceae) ANICETO MENDOzA, BLANCA PEREZ-GarciA, and RAMON Ripa! Departamento de Biologia-Botdnica Estructural y Sistematica Vegetal, Universidad Autonoma Metropolitana—Iztapalapa, Apdo. Postal 55-535, 09340, México, D. F. Fax: (55) 58 04 46 88 AsstTract.—The prothallial development of gametophytes of Olfersia alata and Olfersia cervina s (Dryopteridaceae) is described and compared. Spores are monolete, ellipsoid, and with broadly winged perispore. Germination is Vittaria-type and the prothallial development is Aspidium- type. Adult gametophytes are cordiform-spatulate to cordiform-reniform, with marginal and These two species share some features with some species of Arachniodes, Cyrtomium, Dryo ris, Phanerophlebia, and Polystichum, such as type of germination and prothallial development and trichomes. They differ from Didymochlaena truncatula, which has prothallial development of the Adiantum-type and lacks trichomes on the sexual phase. The genus Olfersia Raddi (Dryopteridaceae), has two species: Olfersia alata C. Sanchez & Garcia Caluff and Olfersia cervina (L.) Kunze. Olfersia alata is endemic to Cuba; its main characteristics are all sterile pinnae have decurrent bases, and fertile leaves which are smaller and have fewer pinna pairs than the vegetative leaves. It grows in mountainous mesophytic forests, between 350—400 m (Sanchez et al., 1991). Olfersia cervina is widely distrib- uted in the tropics, from Southern Mexico (Chiapas, Oaxaca, Veracruz), to Southeastern Brazil and the West Indies. In this species the bases of the ster- ile pinnae are not decurrent onto the rachis, and pinnae are short-petiolu- late. It grows between 450-1000 m in damp tropical forests, on rocky and very shady banks (Moran, 1986, 1995; Riba and Pérez-Garcia, 1999). Both taxa are usually terrestrial, rarely hemiepiphytic, with a short trailing rhi- zome. Leaves are markedly dimorphic, sori are exindusiate and linear to ob- long, and spores are monolete, echinulate with a broad perispore. This paper complements existing information about the morphogenesis of the gametophytic phase of dryopteriod ferns and, particulary, focuses on gametophytes of Olfersia. We hope to contribute in this way to the knowledge of the sexual phase of Mexican ferns. MATERIALS AND METHODS Spores of O. cervina were collected from living plants from the following site: Lote 69 in “Los Tuxtlas” Biological Station, between Laguna Azul and Laguna Seca, Mun. of San Andrés Tuxtla, in the state of Veracruz, Mexico; vouchers are in UAMIZ (AMR 202 and AMR 251). Spores of O. alata were 230 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) collected by Carlos Sénchez and L. del Risco (77825) near Farallones de Moa, Farallon Redondo and La Escondida, Mun. Moa, Holguin province, Cuba; the voucher is in HAJM. Fertile pinnae were kept in paper bags until spores were shed. Subse- quently, the spores and shed material were passed through a sieve (with pores 0.074 mm in diameter) in order to eliminate traces of sporangia and indusia. Spores of each species were sown at an average density of 150-200 spores per cm’ in two small pots with a mixture of black soil and organic matter and in 30 Petri dishes, 5 cm in diameter, containing Thompson’s solution of mineral salts and agar on a sterile nutrient medium (Klekowski, 1969). Petri dishes and pots were kept inside transparent plastic bags in order to avoid contamination and desiccation, with a photoperiod of 12 h light/dark- ness, with artificial light (75 Watt lamps, daylight) and a temperature of 23- 25° C (Mendoza et al., 1999a, 1999b). Two dishes were kept in darkness in order to determine photoblastism. After 100 days, none of the spores grown in darkness had germinated. All pictures of microscopic material were taken from living material grown in the laboratory. RESULTS Spores of both species are monolete, nearly spherical, with a light brown perispore. Spores of Olfersia alata measure (64) 73 (83) x (49) 53 (55) um, including the winged perine around the spore; the perispore measures (15) 16 (20) jm wide (Fig. 1). Spores of O. cervina measure (44) 48 (51) < (37) 39 (40) um, also including the winged perine which measures (5) 6 (8) um wide (Fig. 2). Spores of O. alata are larger than those of O. cervina primarily due to the size of the perispore. These measurements were obtained from an average sample of fifty spores per species. Germination is Vittaria-type (Nayar & Kaur, 1971) in both species. In Olfersia alata germination began 20-23 days after spores were sown, whereas in O. cervina it began 8-12 days after sowing. Gametophytes of both species first develop a rhizoid, which is short, hyaline, and without chloroplasts. The first prothallial cell is short and oval; division begins in this cell with a transverse wall and ultimately forms a short germ-filament, 2—4 cells long. This filament eventually ends in an apical trichome. During this stage of development, the spores retain their coat (Figs. 3-5). In O. alata, the prothallial plate begins to develop approximately 25 days after spore germination from intercalary cells of the filaments which undergo longitudinal divisions (Figs. 6-7). In some cases, the terminal cell of the fila- ment, after producing a trichome, will divide longitudinally in such a man- ner that the trichome is placed over one of the daughter cells, which will remain inactive until other cells develops into a gametophytic plate. This plate, from which a meristematic cell will emerge, is usually asymmetric MENDOZA ET AL.: GAMETOPHYTES OF OLFERSIA phases of Olfersia. 1. Spor cervina (8 days). 4. O. hallial and filamentous and laminar s of germination. 3. prot 1-9. Spores, germination, Spore, O. cervina. 3-4. I Initial stages (22 Sie 5 7. Germ filament. 5-6. O. alata (22 days). 7. O. cervina ye days). 8-9. Young lami- 9. O. cervina (35 day e = cover of spores, cp = trichome, zm cietitiaiathe zone. nar phases 8. O. alata (38 days cell, cr = rhizoid cell, p = perispore, t = 232 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) Fics. 10-16. Laminar arp tae and secretory, unicellular, capitate trichomes of Olfersia. 10-11. O. cervina (41 and 100 days). 12-13. O. alata (95 and 138 days). 14-16. Trichomes. 14. O. cervina (175 days). = O. alata (149 days). 16. O. cervina (175 days). a a = archegonia, cse = extracellular secretion cover, t = trichome. with an apex that continues to change and form a notch. Finally, after 90— 100 days, the prothallial plate becomes cordate, the so-called Aspidium-type prothallial plate development (Nayar & Kaur, 1969; Figs. 8-10). Afterwards, a cushion bearing the gametangia forms and the adult gametophyte is cordi- form-spatulate with many marginal and superficial trichomes. The formation MENDOZA ET AL.: GAMETOPHYTES OF OLFERSIA 233 Fics. 17-20. Adult gametophytes and gametangia of Olfersia. 17-18. Adult phases of O. alata (109-158 days). 19-20. Gametangia. 19. Antheridium, O. cervina (246 days). 20. Mouths of archegonium, O. alata (138 days). ba = mouth of archegonium, cb = basal cell, cm = ring cell, = opercular cell. of the prothallial plate in O. cervina takes less time, beginning on day 15. The pattern of prothallial development is the same as in O. alata. The first adult cordiform gametophytes are completely differentiated 60-80 days after the spores were sown (Figs. 12-13). Trichomes are unicellular, capitate, and secretory (Fig. 14). In O. alata they measure approximately 36 um long by 23 um wide at the base. The api- cal third of the trichome is globose, 17 1m high by 24 um wide, with a thin cover of extracellular secretion ca. 3 pm thick (Fig. 15). Olfersia cervina tri- chomes are 34 pm long by 20 ym wide at the base and the apical third of the trichome is globose, 21 pm high by 26 pm wide, with an extracellular secre- tion 8 ym thick (Fig. 16). These measurements are from mature trichomes, found at the middle basal region of the gametophytes. The gametangia are typical of leptosporangiate homosporous ferns. They begin differentiating between days 120-244 in O. cervina while in O. alata they develop between days 100-150. Antheridia are distributed on the lower surface of the plate on the basal half of the cushion (Fig. 19). Antheridia are globose and consist of a basal cell, a ring cell, and an opercular cell. These three cells surround the androgenous cell. In both species, the necks of the archegonia, have four tiers of neck cells. Archegonia are found on the central region of the plate, on the cushion, and TasLeE 1, Comparison of different stages of the prothallial development of Olfersia alata, and O. cervina with other genera and species of Dryopteridaceae. Type of prothallial Type of Filamentous evelopment and Spores germination phase adult form Trichomes Antheridium Archegonium '*Arachniodes —_ Monolete, Vittaria Long filaments Aspidium, Unicellular, 3 cells 4 rows of cells, with perispore, -5 cells), cordiform capitate, with each row with measuring 3 with an apical gametophytes with a thin coat of 4-6 cells 42 «um trichome lacerate margins extracellular secretion 2 :m *Cyrtomium Monolete with Vittaria Long filaments Aspidium, Unicellular, 3cells 4 rows of cells perispore, (2-5 cells), cordiform capitate measuring 32 with an apical = ga metophytes, with 45 um trichome lacerate margin *Didymochlaena_ Monolete with Vittaria Short filaments Adiantum, Absent 3-4 cells 4 rows of cells, perispore, (2-3 cells), cordiform- throughout each row with measuring 30 apical tri- reniform development 4-6 cells 37 um chome absent gametophytes with entire argins *Dryopteris Monolete with Vittaria_ Long filaments Aspidium, Unicellular, 3cells 4 rows of cells, perispore, (2-5 cells), po capitate each row with measuring 36 with an heii ifor 4-5 cells 51 pm trichom gametophytes with lacerate margins Olfersia alata Monolete with a Vittaria Short filaments Aspidium, Unicellular, 3cells 4 rows of cells, broad1 (2—4 cells), spatulate- capitate with each row with winged with an apical _—_cordiform an extra- 4-5 cells perispore, trichome gametophytes cellular measuring 53 with entire secretion coat 73 pm margin i vEC (200z) € WAGON 26 SNNTIOA “TVNYNOl NYaA NVOMANV TABLE 1. Continued. Type of prothallial Filamentous development and Spores germination phase orm Trichomes Antheridium = Archegonium O. cervina Monolete with Vittaria Short filaments Aspidium, Unicellular, 3 cells 4 rows of cells, i 2—4 cells), spatulate- apitate wi each row with perispore, with an apical _cordifor extra- 4-5 cells measuring 39 trichome iia with cellular 48 pm ee margins secretion coat 8 :m thick 'Polystichum Monolete with Vittaria Long filaments | Aspidium, Unicellular, 3-4 cells 4 rows of cells perispore, cordiform appilate measuring 34 without gametophytes with — capitate 45 pm trichomes lacerate margins secretors and non secretors *Phanerophlebia Monolete with Vittaria Long filaments Aspidium Unicellular, 3-4 cells 4 rows of cells, perispore, —6 cells), PRE HN -cordiform capitate, wit each row with measuring 25 with an apical gametophyes with an extra- 4-6 cells 33 pm trichome lacerate margins cellular secretion coat 5 :m thick ' Chandra & Nayar 1970, * Mendoza et al. 1999a, 1999b,” Pérez-Garcfa et al. 1999. * Mendoza, unpublished. VISHAITO AO SHLAHdOLAWVD “TV LY VZO0NEW Sez 236 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) near the meristematic zone. The necks are oriented toward the basal region of the gametophytes (Figs. 13, 20). Two hundred days after sowing the spores, the young sporophytes had not yet formed. DISCUSSION AND CONCLUSIONS There is literature dealing with the morphology of the gametophytic phase of ferns closely related to Olfersia, of the Dryopteridaceae (both Old and New World), e.g., Arachniodes, Cyrtomium, Didymochlaena, Dryopteris, and Polystichum (Atkinson, 1973; Chandra & Nayar, 1970; Cousens, 1975; Kaur, 1977; Mendoza et al., 1999a, 1999b; Pérez-Garcia et al., 1999: Stokey & Atkinson, 1954). Spores of O. alata average 73 X 53 «um, including the winged perispore; spores of O. cervina average of 48 X 39 um. Spores of O. alata seem much larger, but in reality, if the perispore is not considered, the spores are 38 X 23 um, and the winged perine is 16 um wide or more in its widest part. Spores of O. cervina are 39 X 31 wm and the perispore is approximately 6 jum wide at its widest point and tends to be more spherical, which is an in- dication that the spores of O. alata are a little smaller than those of O. cervi- na. (Figs. 1, 2). Both species share the same germination pattern, Vittaria-type, which is the most common type in ferns. In this type, the rhizoid develops first after the formation of a wall perpendicular to the polar axis of the spores. Eventu- ally, the first prothallial cell divides by means of the formation of a perpen- dicular wall thus giving rise to two cells. The apical cell then divides again, giving rise to a short filament 2-4 cells long. The time for germination differs between these two species; spores of O. cervina germinate faster (8-12 days) compared to spores of O. alata (20-22 days). Prothallial development in both species is of the Aspidium-type in which the germ filament commonly ends in a trichome, and the prothallial plate is formed by the activity of the intercalary cells of the filament. The adult gametophyte develops faster in O. cervina (60-80 days) than in O. alata (90-100 days). Trichomes differ in size and in the thickness of the extracellular secretion; the longest ones, belongins to O. alata (36 X 23 um), have a thinner extra- cellular secretion (3 um), whereas trichomes of O. cervina are slightly shorter (34 X 20 um) and have a thicker (8 um) extracellular secretion. Olfersia alata and O. cervina share features with the following dryopterid genera: Arachniodes, Cyrtomium, Dryopteris, Phanerophlebia, and Polysti- chum (Atkinson, 1973; Chandra and Nayar, 1970; Cousens, 1975; Mendoza et al., 1999b; Pérez-Garcia et al., 1999). These genera all have monolete spores with perispore, a Vittaria-type germination pattern and an Aspidium- type prothallial development. However the two Olfersia species differ from the rest in the shape of their trichomes, which are short and wider at the base, capitate, with a globose apex, and with a dense extracellular secretion. MENDOZA ET AL.: GAMETOPHYTES OF OLFERSIA 237 The gametophyte margins are entire in Olfersia, compared with the lacerate margins of species of the other genera. These other genera also have longer, capitate trichomes, with very thin extracellular secretions distributed on the lacerate margins and on the surfaces of the plate (Table 1). Olfersia alata and O. cervina, together with the above mentioned taxa, share some features with Didymochlaena truncatula, such as the monolete spores and Vittaria-type germination. This last species differs from the rest in that is has a prothallial development of the Adiantum-type, characterized by a differentiation of an apical meristematic cell during the early stages of the plate’s formation. The gametophytes of Didymochlaena, are completely glabrous throughout their development, in contrast to the other species of Dryopteridaceae mentioned. d on our results, we conclude that Olfersia alata and O. cervina share characteristics such as the Vittaria-type germination pattern, Aspidium-type prothallial development, and unicellular capitate trichomes with a uniform extracellular secretion. These same characteristics are characteristic of spe- cies of Arachniodes, Cyrtomium, Dryopteris, Phanerophlebia, and Polysti- chum (Table 1). The most common feature of all of these genera is the development of an apical trichome during the filamentous stages of prothal- lial development. Gametophytes of Didymochlaena truncatula differ from these genera in having prothallial development of the Adiantum-type and lacking trichomes. Finally, with the exception of Didymochlaena truncatula we did not find important differences among the different taxa of the Dryo- pteridaceae. ACKNOWLEDGMENTS This paper is part of an MSc (Plant Biology) thesis, Morfogénesis de la fase sexual de pterido- fitas mexicanas, familia et aan eb Adee Sgn of the sexual phase of Mexican Pteri- dophytes, family Dryopteridaceae”), written by the first author, developed under the supervision his photographic assistance, and the anonymous reviewers for their suggestions and criticisms. LITERATURE CITED ATKINSON, L. M. 1973. The Gametophyte and Family Relationships. In: A. C. Jermy, J. A. Crabbe & B. A. Thomas (Eds.). The Phylogeny and Classification of The Ferns. J. Linn. Soc. Bot. a No.1, 67:73-90. CHANDRA, P. & B. K. NAyar. 1970. Morphology of Some hg ge shane i. ny Gametophytes Arachniodes, Cyrtomium and Polystichum. J. Linn. Soc. B 3:265— Cousens, M. I. 1975. Gametophyte Sex Expression in Some ones ot ciehoni Amer. Fern. J. 0:13-—27. Kaur, S. 1977. Morphology of a yee and Juvenile Sporophytes of Some Species of Dry- opteris. Proc. Indian Acad. S$ —171. — E. J., JR. 1969. eae pe of the Pteridophyta. II. A study of the Blechna- ceae. J. Linn. Soc. Bot. 62:361-377 238 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) MEnpoza, A., B. PEREZ-Garcia & R. Ripa. 1999a. Morfologia y anatomia del gametofito de Didymo- chlaena oe (Dryopteridaceae). Rev. Biol. Trop. 47:87—93 . 1999b. Morfogénesis = la fase sexual del lisleaho Arachniodes denticu- ta Dryopteridacen) Rev. Biol. Trop. 47:791—797. mers 2 G. 6. The neotropical fern ie ski in Fern J. 76:161-17 1995. achat meri Pages 210-226. In: R. C. Moran & R. Riba (eds.). Ae Mesoameri- cana, Vol. 1: Psilotaceae a Salviniaceae. Instituto re Biologia, Universidad Nacional o; DF Nayar, B. K. & S. Kau ee is of prothallial development in homosporous ferns. Phyto- ee, 19: ote 971. Cametopytes of homosporous ferns. Bot. Rev. 37:295-396. PE eR B., (l MENpoza, I. REYES & R. Ripa. 1999. Morfogénesis de la fase sexual de seis es- pecies geet de helechos del género Dryopteris Fries aie araigeh Rev. Biol. Trop. 47:63-— Ripa, R. & B. Vee oe pts Dryopteridaceae. Flora de México, Consejo Nacional de la Flora e México oe SANCHEZ-VILLAVERD oe ciA-CALUFF & C. ZAVARO-PEREZ. 1991. Nueva onciigee cubana del género Olfersia ect eesalie esate ey Fontqueria 31:229-23 Story, A. G. & L. R. ATKINSON. 1954. The gametophyte of Didymochlaena hee Desv. Phyto- morphology 4:310-315 American Fern Journal 92(3):239-—246 (2002) SHORTER NOTES Botrychium hesperium in the Wallowa Mountains of Oregon.—The Wallowa Mountains of northeastern Oregon boast the greatest fern diversity in the state. We reported 47 taxa in the range (Zika & Alverson, Amer. Fern J. 86: 61-64. 1996), which included 14 taxa of Botrychium. A number of ele- ments from the Rocky Mountains are found in Wallowa County, to which we can now include Botrychium hesperium (Maxon & R. T. Clausen) W.H. Wagner & Lellinger, an addition to the Oregon flora (Wagner & Wagner, Ophio- glossaceae in Flora of North America, Vol. 2, Oxford Univ. Press, 1993). Botrychium hesperium is restricted in the Wallowa Mountains to a narrow elevational band in the Lostine River drainage, between 1535-1660 meters, where steep canyon walls shade the valley floor from direct sunlight early and late in the day. It is found in mesic meadows or forest edges, in full sun or partial shade, at all aspects, but only on gentle slopes or flats on the valley floor. It has yet to be located on steep slopes at higher elevations. The forests are primarily Pinus contorta Dougl. ex Loud., with low wet areas dominated by Picea engelmannii Parry ex Engelm. Associated herbs include: Achillea millefolium L., Agoseris aurantiaca (Hook.) Greene, Antennaria rosea Greene, Calamagrostis rubescens Buckl., Carex concinnoides Mack., C. geyeri Boott, C. hoodii Boott, Elymus glaucus Buckl., Erigeron compositus Pursh, Festuca occidentalis Hook., Fragaria virginiana Duchesne, Gentiana amarella _L., Hieracium albiflorum Hook., Linnaea borealis L., Sedum stenopetalum Pursh subsp. stenopetalum, Senecio pseudaureus Rydb., and Viola adunca Sm. It grows with Botrychium lanceolatum (Gmel.) Angstr. subsp. Janceolatum., B. minganense Victorin, B. paradoxum W. H. Wagner, B. pedunculosum W. H. Wagner, and B. pinnatum St. John. The sites are valley bottom Quater- nary surficial deposits, locally reworked by the Lostine River or small tributar- ies. Adjacent slopes are sedimentary bedrock in the Triassic/Jurassic Hurwal Formation. In places the upper west wall of Lostine Canyon is granite, and the east wall is pure limestone of the Martin Bridge Formation. It is possible that all or most of the Botrychium populations are influenced by basic or cir- cumneutral groundwater percolating through calcareous glacial till or morai- nal debris. It may be no coincidence that the richest diversity and greatest abundance of Botrychium species are found in the calcareous canyons of the Wallowa Mountains, rather than in the granitic or volcanic basins. We are aware of four extant populations of Botrychium hesperium in the Wallowas. The Oregon range of the species is included in ca. 5.5 km of river valley. The total known number of plants at this time is less than 100, and they face threats from fire suppression, pack animal grazing, wood-cutting, and recreation-associated activities, despite the fact that most or all plants are within the Lostine River Wild and Scenic River corridor, a part of the Eagle Cap Ranger District of the Wallowa-Whitman National Forest. 240 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) Collections of Botrychium hesperium were first made in 1981 (W. H. Wagner 81130 MICH), with later collections in 1991 (Zika & Alverson 11295 WTU), 1992 (Zika & Alverson 11794 WTU), 1993 (Wagner et al. 93047 MICH) and 1996 (Zika & Alverson 12908 OSC). We were puzzled by these plants for many years, and thought they might represent an undescribed taxon, related to B. hesperium, but with slightly angular upper pinnae and shorter basal _pinnae. This was a false impression, based in part on the large Wallowas plants growing in sheltered or partly shaded sites, and based on a limited sample of B. hesperium from Oregon and elsewhere. To get a better idea of variation in B. hesperium, we studied large living populations in Montana, Arizona and Colorado. Finally, as we saw more Oregon plants, we con- cluded they were part of the natural variation of B. hesperium, united by their grayish-green color, exaggerated and asymmetrical basal pinnae, broad rounded upper pinnae, and ample sporophores. We are pleased to acknowledge our funding sources for fieldwork: the Native Plant Society of Oregon, the Oregon Natural Heritage Program, and the Wallowa-Whitman National Forest. We are grateful for specimens and discussions of B. hesperium, provided by Peter Root, Peter Lesica, Kathy Ahlenslager, and Don Farrar.—PrTrer F. ZIKA and Epwarp R. ALVERSON, Herbarium, Dept. of Botany and Plant Pathology, Oregon State University, Corvallis, OR 97331, and Warren H. Wacner (deceased) and FLORENCE S. Wacnrr, Department of Ecology and Evolutionary Biology, University of Michigan, Ann Arbor, MI 48109. A Binomial for the Hybrid Polypodium of Eastern North America.—Two species of Polypodium (Polypodiaceae) occur in eastern North America, the diploid P. appalachianum Haufler & Windham and the tetraploid P. virginia- num L. These species hybridize, producing a sterile triploid recognized by its abortive spores and intermediate morphology. The differences between these three taxa are well described by Haufler and Wang (Amer. J. Bot. 78:624—629. 1991) and Haufler and Windham (Amer. Fern J. 81:7—23. 1991). The triploid hybrid so far has been found only on the Appalachian Plateau where P. appalachianum and P. virginianum are sympatric. The hybrid has been documented so far in Ontario, Canada and eight states: New Hampshire, New Jersey, North Carolina, Ohio, Pennsylvania, Tennessee, Vermont, and Virginia (Evans, Research Div. Monograph 2. Virginia Poly- technic Inst. and State Univ., Blacksburg, VA. pp. 117-146. 1970; Haufler & Wang, op. cit.; Montgomery, Bartonia 59:113-117. 1996). Kentucky and West Virginia can be added to this distribution, based upon specimens at OS and WVU, respectively. The hybrid likely will be documented in other states and provinces as well. Indeed, the triploid may prove rather frequent, as shown for New Jersey and Pennsylvania by the work of Montgomery cited above. SHORTER NOTES 241 It seems appropriate and practical that this widespread hybrid have a binomial. Perhaps providing this taxon with an epithet may raise botanists’ awareness of this taxon and spur future discoveries and understanding of this hybrid. Polypodium Xincognitum Cusick, fAybr. nov.—Holotype: Ohio, Meigs County, sandstone exposures on mesic slope above Leading Creek, Co Rt 10, 0.25 mi (0.02 km) SW of Twp Rt 27, N of Dexter, Sect 6, Salem Twp, 6 Aug 1985, Cusick 24620, OS; Isotypes, MICH, MU, NY. Hybrida e Polypodium appalachianum et P. virginianum exorta, aliis char- acteribus inter parentes media, sporis abortivus. My research was supported in part by the Division of Natural Areas and Preserves, Ohio Department of Natural Resources.—ALLISON W. Cusick, Divi- sion of Natural Areas and Preserves, Ohio Department of Natural Resources, 1889 Fountain Sq. Ct., F-1, Columbus OH 43224. Lycopodium lagopus New in West Virginia.—West Virginia is a southern outpost for many boreal species (e.g. Larix Jaricina in Preston County) that were stranded in the state’s highlands and arctic-like bogs following the last glacial retreat (P.D. Strausbaugh and E.L. Core, Flora of West Virginia, Mor- gantown WV, Seneca Books, 1997). Along the Allegheny Front, elevations reach 1482 m (Spruce Knob) and there are ten peaks over 1430 m. Lycopodium lagopus (Laestradius ex C. Hartman) G. Zinserling ex Kuzeneva-Prochorova, (Fl. Murmansk Obl. 1:80, 1953), generally more northern in its distribution, was recently located here as well. A small, but thriving population grows on the site of a coal strip mine, now used as a cross country ski trail in Blackwater Falls State Park, Tucker County, at an elevation of about 1070 m. Its sister spe- cies, L. clavatum, is also here in abundance, but the two lycopods remain dis- tinct; L. lagopus features single strobili on slender peduncles, a more compact growth habit, more appressed and shorter leaves, and sporophylls that taper gradually to a hair tip. Lycopodium lagopus (formerly L. clavatum var. monostachyon Hooker and Greville) goes by the apt common name “‘one-cone club-moss”’ (Flora of North America, New York, Oxford Univ. Press, 1993). It shares many characters with the common club moss, L. clavatum, e.g., general growth and branching patterns, stalked strobili, and hair-tipped leaves, but L. clavatum has multi- ple strobili (typically two) on most of its peduncles, spreading and longer leaves, and sporophylls that end abruptly in hair tips. No hybrids are docu- mented between these closely related species, nor, for that matter, between any species in the genus Lycopodium s.s. This is in sharp contrast to the many hybrids described since 1956 within the related genera Lycopodiella, Huperzia, and Diphasiastrum (J. Eiger, Biol. Rev. City Coll. 18:17-—22, 1956; Flora of North America). As a boreal plant, L. lagopus occurs from Alaska to Newfoundland, Green- 242 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) land, Scandinavia, and northern Eurasia. In the contiguous 48 states it has been reported from Maine, Michigan, Minnesota, Wisconsin, New York, New Hampshire, Vermont, and now West Virginia. Michigan would be the closest known neighbor of the West Virginia population, about 650 km dis- tant (Flora of North America, cited above). The West Virginia site is the flat top of an abandoned coal strip mine within Blackwater Falls State Park, near the town of Davis. The park was established in 1937, and mining operations within its borders ceased prior to that. Since that time a remarkable assem- blage of plants has reclaimed the coal spoils piled along the mine highwall. In cool ravines there is Tsuga canadensis, while on the open, sunny, coal- strewn areas Picea rubens, Acer rubrum, Rhododendron maximum, Kalmia latifolia, and Vaccinium species dominate the woody flora. Three orchids are prominent among the herbaceous plants- Cypripedium acaule, Platanthera clavellata, and Spiranthes cernua. Many grass, sedge, and Sphagnum species grow in the boggy soils near two ponds at the intersection of the Dobbin House and Woodcock ski trails in the area of L. Jagopus. An impressive group of fern allies has also reclaimed this disturbed, acidic habitat, includ- ing Lycopodiella inundata on moist soil near the ponds. Diphasiastrum digi- tatum, D. tristachyum, and their hybrid D. Xhabereri are abundant on exposed tailings. Lycopodium obscurum, L. dendroideum, and L. hickeyi al- so occur along the wooded edges. And, the aforementioned L. clavatum is found in nearly all surrounding habitats. Several Dryopteris species and Pteridium aquilinum (with some rare, fertile colonies) are also common in the immediate area. Asplenium montanum grows on granite rocks along the Pase trail nearby, and Vittaria appalachiana (Appalachian gametophyte) is spreading under a sandstone ledge near the prominent waterfall for which the 683 hectare (1688 acre) park is named. The Lycopodium lagopus colony consists of about a dozen long rhizomes, three occurring on exposed soil adjacent to the Woodcock Trail and the rest in a protected area about 6 m into low spruce woods beyond the trail. The colony is probably clonal and is quite fertile, nearly all upright shoots bear- ing characteristic single strobili when the site was surveyed in mid-July, 2001. A voucher specimen of one fertile, upright shoot was collected for de- posit with the herbarium of the Carnegie Museum in Pittsburgh, PA (sheet No. 494379, CM). The origin of the ‘‘one-cone club-moss” here is uncertain, but it is hardly the only disjunct, rare pteridophyte known in West Virginia. The western species Asplenium septentrionale and Cheilanthes eatonii (C. castanea) occur on shale in Hardy and Monroe Counties (W. H. Wagner, Jr., Ann. Missouri Bot. Gard. 59:203-217, 1972).—JoAN EIGER GOTTLIEB, 2310 Marbury Road, Pittsburgh, PA 15221. SHORTER NOTES 243 Marsilea mutica in Virginia.—Of the six species of water-clover, Marsilea, described in Flora of North America, five are native and M. quadrifolia L. is introduced in much of the northeastern United States. Johnson (pp 332-333, in FNA ed. comm., Flora of North America, vol. 2, 1993) suggested that M. quadrifolia has been deliberately planted as a curiosity because many local- ities are artificial bodies of water. A seventh species, the paleotropical Marsi- lea minuta L., has been collected in Florida (Burkhalter, Sida 16:544—-549, 1995). We document here the apparent establishment of a previously un- reported water-clover in southeastern Virginia, M. mutica Mettenius, which, unlike most species in the genus, can be readily identified in sterile condi- tion by its two-toned leaves (Fig. 1). This station represents the first docu- mented occurrence of this Australasian species in North America. A small but vigorous population was discovered on 20 October 2001 grow- ing in shallow water (<1.0 m depth) and adjacent muddy shores for 10 m along Cooper’s Ditch, a canal dug through non-tidal forested wetlands in the early 1980’s as a stormwater in the city of Chesapeake. Plants were re- stricted to shallow water and associated with Utricularia sp., Hydrilla verti- cillata (L.f.) Royle, Hydrocotyle verticillata Thunb., Myriophyllum pinnatum (Walter) BSP, and filamentous algae. Like other species in the genus, floating leaves were much larger than leaves on terrestrial plants (Fig. 1). Rhizomes were rooted. No sporocarps were present. Collection data follows: Virginia: City of Chesapeake, NE side of Hillwell bridge across Cooper’s Ditch, Co- ordinates: 36°42'01”N, 76°13'32”W, 20 October 2001, L. J. Musselman and D. A. Knepper, 2001-36 (ODU). Winter hardiness in this species is not known. Nothing is known about the invasiveness of this species and the population should be monitored to measure its persistence and its spread.—Davip A. Knepper, U S Army Corps of Engineers, Fort Norfolk, Norfolk, VA 23510-1096; Davin M. JOHNsoN, Department of Botany-Microbiology, Ohio Wesleyan University, Delaware, Fic. 1. Habit view of Marsilea mutica. 244 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) OH 43015; and Lyrron JoHN MussELMAN, Department of Biological Sciences, Old Dominion University, Norfolk, VA 23529-0266. 3,8-Di-C-arabinosylluteolin, a new flavonoid from Pteris vittata.—In spite of the fact that fern flavonoids are of chemotaxonomic interest, little is known of the distribution of these compounds in some fern families (e.g. Pteridaceae). Previous work on the flavonoids of Pteris vittata L. (Pterida- ceae) has led to the identification of an anthocyanin (luteolinidin 5-O-gluco- side) by Harborne (Phytochemistry 5:589-600, 1966). In addition, acid hydrolysis of extracts of this fern has led to the identification of kaempferol, quercetin, leucocyanidin and leucodelphinidin by Voirin (Ph. D. thesis, Uni- versity of Lyon, p. 151, 1970). More recently 3-C-(6'’’-acetyl-cellobiosyl)- apigenin (Amer. Fern J. 89:217-220, 1999) and 6-C-cellobiosylisoscutellarein 8-methyl ether together with quercetin 3-O-glucuronide and rutin (Amer. Fern J. 90:42—47, 2000) have been identified by Imperato and Telesca. Three kaemperol glycosides (3-O-glucoside, 3-O-glucuronide and 3-O-(X",X’-di- protocatechuoyl)-glucuronide) together with quercetin 3-O-(X",X"-di-protoca- techuoyl)-glucuronide have been found in this fern by Imperato (Amer. Fern J. 90: 141-144, 2000). In the present paper a new C-glycosylflavone (identified as 3,8-di-C-arabi- nosyl-luteolin (I)) and 6-C-arabinosyl-8-C-glucosylluteolin (II) have been iso- lated from Pteris vittata L. growing in the Botanic Garden of the University of Naples. This fern has been identified by Dr. R. Nazzaro (University of Naples); a voucher specimen (149.001.001.01) has been deposited in the Her- barium Neapolitanum (NAP) of the University of Naples. Flavonoids (I and II) were isolated from an ethanolic extract of aerial parts of Pteris vittata L. by preparative paper chromatography in BAW (n-butanol- acetic acid-water, 4:1:5, upper phase), 15% HOAc (acetic acid) and BEW (n-butanol-ethanol water, 4:1:2.2). Further purification was carried out by Sephadex LH-20 column chromatography eluting with methanol. Color reactions (brown to yellow in UV+NHs), ultraviolet spectral analysis in the presence of usual shift reagents (Amax (nm) (MeOH) 258, 272, 348; +NaOAc 283, 322 (sh), 402; +NaOMe 266, 281, 340 (sh), 407 (increase in in- tensity); +AICl; 277, 299 (sh), 331 (sh), 426; +AICl,/HCl 280, 300 (sh), 358, 388) and chromatographic behaviour (Ry values on Whatman No 1 paper: 0.15 in BAW; 0.31 in 15% HOAc; 0.12 in HO) suggested that flavonoid (I) may be a flavonoid glycoside with free hydroxyl groups at positions 5, 7, 3’ and 4’. Since treatment with 2N HCl (2 hr at 100°C) failed to produce an aglycone, flavonoid (I) may be a C-glycosylflavonoid. Electrospray mass spec- trum (ESMS) showed a pseudomolecular ion at m/z 573 [((M+H)+Na]*, an ion at m/z 595 [(M+H)+2 Na]* and an ion at m/z 1123 [(M x 2)+ Na+H]* which corresponds to a dimer. These data suggest that flavonoid (1) is a di-C- pentosylluteolin. 'H NMR spectrum (DMSO-d,) showed signals at 6 3.11— 3.91 (ten sugar protons, m), 5 4.61 (1H, d, J=8 Hz, anomeric proton), 6 4.68 (1H, d, J=8 Hz, anomeric proton), 5 6.27 (1H, s, H-6), 6 6.93 (1H, d, J=8.8 SHORTER NOTES 245 Fic; 1. Hz, H-5’), 6 7.39 (1H, dd, J=2.0 and 8.8 Hz, H-6’) and 6 7.40 (1H, d, J=2 Hz, H-2'). These data suggest that flavonoid (I) is a luteolin 3,8-di-C-pentoside. Wessely-Moser isomerization (3N HCl; 3 hr at 100°C) gave a mixture in which flavonoid (I) and four isomers were detected by paper chromatography in BAW; these isomers were not present in sufficient amount to allow charac- terization. 'H NMR spectrum (DMSO-dg) of the above mixture showed a sin- glet at 6 6.56 (H-8) and a singlet at 6 6.26 (H-6) confirming that flavonoid (I) is a 3,8-di-C-glycosylflavone. Since flavonoid (I) gave at least four isomers, ara- binose may be attached at C-3 and/or C-8 because C-arabinosylflavones on acid treatment undergo pyranose-furanose isomerization and a-linkage-f-link- age isomerizatiion of C-glycosidic link as described in a review of Chopin et al. (pp. 449-503 in J.B. Harborne and T. J. Mabry eds., The Flavonoids: Ad- vances in Research, Chapman and Hall, London, New York, 1982). Treatment of flavonoid (I) with 2,2-dimethoxypropane and 6N HCl-dioxan in dry di- methylformamide according to Jarman and Ross (J. Chem. Soc. (C):199-203, 1969) gave a diisopropilidene derivative ([M]* at m/z 630 in EI-mass spec- trum); hence arabinose is attached at C-3 and C-8 of flavonoid (I) since this isopropilidenation is specific for C-galactosyl and C-arabinosyl residues in mono- and di-C-glycosylflavones as described in the above review by Chopin et al. FeCl, oxidation of flavonoid (I) gave L-arabinose. The above results show that flavonoid (I) is 3,8-di-C-arabinosylluteolin (Fig. 1), a new natural product; this is the first report of a 3,8-di-C-glycosylflavone from ferns. 3,8-Di-C-glycosylflavones were found for the first time in plants in 1985 by Matsubara et al. (Nippon Nogeikayaku Kaishi 59: 405-410, 1985) who iso- lated apigenin 3,8-di-C-glucoside and diosmetin 3,8-di-C-glucoside from Cit- rus sudachi peelings; subsequently these two flavonoids have been found also in Citrus sinensis peelings (Agric. Biol. Chem. 50: 781-783, 1986) and the former flavonoid has been found also in Citrus junos peelings (Nippon Nogeikayaku Kaishi 59: 683-687, 1985) by Kumamoto et al. Flavonoid (II) was identified as luteolin 6-C-arabinoside-8-C-glucoside by ultraviolet spectral analysis with usual shift reagents, treatment with 2N HCl (which failed to give an aglycone), ESMS (which gave a pseudomolecular 246 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 3 (2002) ion at m/z 603 ([(M+H)+Na]*), FeCl, oxidation (which gave D-glucose and L-arabinose) and "H NMR spectrum (DMSO-d,). Assignment of D-glucose to C-8 was based on doublings of signals in 'H NMR spectrum (two signals were observed for H-3 and H-6’) since this feature is due to the presence of a C-linked hexose at C-8 as described in a review by Jay (pp. 57-93, in J.B. Harborne ed., The Flavonoids, Advances in Research since 1986, Chapman and Hall, London, 1994). Flavonoid (II) is a new fern constituent; it has pre- viously been found in bryophytes (Blepharostoma tricophyllum and Pleuro- zia conchifolia) and in angiosperms (Lespedeza capitata, Glycine max and Astrantia major) as described in the review by Jay. The author thanks MURST (Rome) for financial support. Mass spectral data were provided by SESMA (CNR, Naples).—Fiuirro Imperato, Dipartimento di Chimica, Universita della Basilicata, I-85100 Potenza, Italy. American Fern Journal 92(3):247 (2002) REVIEW Pteridophytes of Upper Katanga (Democratic Republic of Congo), by Jan Kornas, Anna Medwecka-Kornas, Francois Malaisse and Malgorzata Matjasz- kiewicz. 2000. Botanical Papers 35, Institute of Botany Jagiellonian Univer- sity, Kracow, Poland. This fairly standard pteridophyte flora covers the Upper Katanga, which is the southeastern part of the Democratic Republic of Congo (formerly Zaire). The area is bordered by Tanzania, Zambia and Angola in central Africa. The work is the outgrowth of collections by two Belgian botanists in Katanga as identified by the Polish pteridologist Jan Kornas, author of many articles on African ferns. The paper was completed by Anna Medwecka-Kornas after the death of Dr. Kornas in 1994. The 180-page book includes a description of the area (climate, geology, soils and vegetation), a list of species by family and genus (alphabetically), distribution maps, and remarks on the pteridophyte flora. The description of vegetation types is brief but complete, with correlation to soils and climate. There are, unfortunately, no keys or descriptions for the species; however, for each species, distribution, relative abundance, herbarium citations and habit are given. Taxonomic notes are useful for some species where there are questions or problems. The final section of the work is a brief discussion of the taxonomic compo- sition of the Pteridophytes to adjacent regions of Africa. The flora includes 183 species in 60 genera. Asplenium is the largest (27 species), as expected in this part of the world, followed by Thelypteris (12), Selaginella (11), and Pteris and Trichomanes (9 each). Although some species are relatively com- mon, more than 50 species have been found at only one or two stations. Katanga has a richer pteridophyte flora than some other adjacent regions, but not as rich as tropical East Africa. The publication of this flora is important because the destruction of forests and political unrest in the region may prevent gathering of additional information in the near future.—JAmes D. Montcomery, Ecology III, Inc., 804 Salem Blvd., Berwick, PA 18603. American Fern Journal 92(3):248 (2002) REVIEW The Illustrated Flora of Illinois. Ferns. 2nd ed., by Robert Mohlenbrock. Southern Illinois University Press, Carbondale. 240 pp. $39.95. Local, illustrated fern floras provide a wonderful service by providing a handy, complete, and easy to use reference without the ‘“‘chatter” of species not found in the area. These floras are designed to provide a streamlined, comprehensive view of the pteridoflora. Without a doubt, Mohlenbrock’s first edition accomplished this goal and more. The Introduction provided a brief account of the history of Illinois fern collecting, a brief account of pteri- dophyte morphology, and a very nice discussion of habitats. Add to this the combination of functional keys, state maps, detailed species accounts, and excellent illustrations and you have a recipe for success. Having said this, the second edition is a considerable disappointment. All of the new material is packed into a 45 page Appendix. The new taxa are keyed out separately from all the rest in a section entitled “Key to the Additional Ferns of IIli- nois’’. For a plant not covered in the first edition, therefore, one must first attempt to identify the species in the main part of the book and then go to the Appendix and start over. Much of the Appendix consists of a “do-it-yourself” editing experience, in which you look at information provided and then go into the main part of the text and fix it up. As an example, on page 180 you are told that the name for Lycopodium porophilum Lloyd & Underwood (see p. 24) is Huperzia po- rophila (Lloyd & Underwood) Holub and that you should add dots to the map on page 26 to include Brown, Carroll, Cumberland, JoDavies, Johnson, Lake, Lee, Randolph, Rock Island, Schuyler, and Will counties. To make things worse, for any genus in which new taxa have been added there are new keys that should be utilized, so please make note not to use the ones in the first 173 pages. The new illustrations while adequate for identification, in most cases lack the detail and precision of those found in the previous edition. While the second edition does provide up to date information, it seems unreasonable in this day of electronic wizardry that the changes found in the Appendix could not have been incorporated into a cohesive w- hole before publication—R. James Hickey, Botany Department, Miami Uni- versity, Oxford, Ohio 45056. tanical Garden Libr, WA WIM If iy Authors are encouraged to submit manuscripts pertinent to pteridology for pub- lication in the American Fern Journal. Manuscripts should be sent to the Editor. Acceptance of papers for publication depends on merit as judged by two or more referees. Authors are encouraged to contribute toward publishing costs; however, the payment or non-payment of page charges will affect neither the acceptability of manuscripts nor the date of publication. Authors should adhere to the following guidelines; manuscripts not so prepared may be returned for revision prior to review. Submit manuscripts in triplicate (xerocopies acceptable), including review copies of illustrations and originals of illustrations. After review, submission of final versions of manuscripts on diskette (in PC- or Mac-compatible formats) is strongly encouraged. Use standard 8% by 11 inch paper of good quality, not “erasable” paper. Double space manuscripts throughout, including title, authors’ names and addresses, short, informative ab- stract, text (including heads and keys), literature cited, tables (separate from text), and figure captions (grouped as consecutive paragraphs separate from figures). Arrange parts of manuscript in order just given. Include author’s name and page number in upper right corner of every sheet. Provide margins of at least 25 mm all around on typed pages. Do not submit right-justified copy, avoid footnotes, and do not break words at ends of lines. Make table headings and figure captions self-explanatory. Use S.I. (metric) units for all measures (e.g., distance, elevation, weight) unless quoted or cited from another source (e.g., specimen citations). For nomenclatural matter (i.c., synonymy and typification), use one paragraph per basionym (see Regnum Veg. 58:39—40. 1968). Abbreviate titles of serial publi- cations according to Botanico-Periodicum-Huntianum (Lawrence et al., 1968, Hunt Botanical Library, Pittsburgh) and its supplement (1991). References cited only as part of nomenclatural matter are not included in literature cited. For shorter notes and reviews, omit the abstract and put all references parenthetically in text. Use Index Herbariorum (Regnum Veg. 120:1—693. 1990) for designations of her- bairia. Illustrations should be proportioned to fit page width with caption on the same yage. Provide margins of at least 25 mm on all illustrations. For continuous-tone illustrations, design originals for reproduction without reduction or by uniform amount. In composite blocks, abut edges of adjacent photographs. Avoid com- bining continuous-tone and line-copy in single illustrations or blocks. Coordinate sequence and numbering of figures (and of tables) with order of citation in text. Explain scales and symbols in figures themselves, not in captions. Include a scale and reference to latitude and longitude in each map. Proofs and reprint order forms are sent to authors by the printer. Authors should send corrected proofs to the editor and reprint orders to the printer. Authors will be assessed charges for extensive alterations made after type has been set. For other matter of form or style, consult recent issues of American Fern Jour- nal and The Chicago Manual of Style, 14th ed. (1993, Univ. Chicago Press, Chicago). Occasionally, departure from these guidelines may be justified. Authors are Satie to consult the editor for assistance with any aspect of manuscript prepara’ 5 aaa louse than 32 printed pages may be sent to the Editor of Pteridologia (Memoir Editor, see cover 2). PTERIDOLOGIA ISSUES IN PRINT The following issues of Preridologia, the memoir series of the American Fern Society, are available for purchase: 1. Wagner, David H. 1979. Systematics of Polystichum in Western North America North of Mexico. 64 pp. $10.00 postpaid. 2A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Chocé (Part 1: Psilotaceae through Dicksoniaceae). 364 pp. $32.00 postpaid. 3. Lellinger, David B. 2002. A Modern Multilingual Glossary for Taxonomic Pteridology. 263 pp. $28.00 postpaid. Send your order with a check or money order to: American Fern Society, Inc., c/o U.S. National Herbarium MRC-166, Smithsonian Institution, Washing- ton, DC 20560. AMERICAN FERN JOURNAL ON MICROFICHE Volumes 1—61 of the American Fern Journal are available as archival quality, silver positive microfiches. Single volumes or the entire run may be purchased. The fiches are easily read with 10X or greater magnification (using a dissecting microscope and transmitted illumination or a fiche reader). Silver negative micro- fiches of vols. 1-50 are also available. The price is $4.00 per volume or $244.00 per set of 61 volumes, postpaid. Send your inquiry or order with a check or money order to: American Fern Society, Inc., c/o Dr. James D. Montgomery, Ecology III, Inc., R.D. 1, Box 1795, Berwick, PA 18603. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://www.amerfernsoc.org/ OK | A395 AMERICAN os FERN number JOURNAL pe QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY The Morphological and Genetic Distinctness of Botrychium minganense and B. crenula- tum as Assessed by Morphometric Analysis and RAPD Markers Linda M. Swartz and Steven J. Brunsfled 249 Reproductive Behavior of Cloned Gametophytes of Pteridium aquilinum (L.) Kuhn Forbes W. Robertson 270 A New Population of Aleutian Shield Fern scp gare aleuticum C. Christens.) on Adak Island, Alaska andra Looman Talbot and Stephen S. Talbot 288 Shorter Note Trichomanes ribae (Hymenophyllaceae), a New Filmy Fern from Costa Rica and Panama Leticia Pacheco 294 Referees for 2002 296 Index to Volume 92 (2002) 297 The American Fern Society Council for 2002 CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS 66045-2016. President TOM RANKER, University Museum, Campus Box 265, University of Colorado, Boulder, CO 80309-0265. Vice President W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI 53233- : id retary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916-1 : 10. Treasurer GEORGE YATSKIEVYCH, Missouri Botanical Garden, P. O. Box 299, St. Louis, MO 63166-0299. Membership Secretary JAMES D. MONTGOMERY, Ecology III, 804 Salem Blvd, Berwick, PA 18603-9801. Back Issues Curator R. JAMES HICKEY, Botany Dept., Miami University, Oxford, OH 45056. Journal Editor DAVID B. LELLINGER, U.S. National Herbarium MRC-166, mithsonian rene ee DC 20560-0166. Memoir Editor gto JOHNSON-GROH, Dept. of Biology, Gustavus Adolphus College, OO W. College Ave., St. Peter, MN 56082-1498. Bulletin Editor American Fern Journal EDITOR R. JAMES HICKEY Botany Department, ami University, Oxford. OH 45056 ph. (513) Sy 6000, e-mail: hickeyrj@muohio.edu ASSOCIATE EDITORS GERALD J GASTONY (2... — of Biology, Indiana University, Bloomington, IN 47405-6801 CHRISTOPHER H. “spariakacnie Dept. of Botany, University of Kansas, Lawrence, KS 66045-2106 ROBBIN C. MORA New York Botanical Garden, Bronx, NY 10458-5126 JAMES H. PECK Dept. of tag University of Arkansas—Little Rock, 1 S. University Ave., Little Rock, AR 72204 The “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the een stanky of ferns. It is owned by the American Fern Society, and published at The American Fern Society, % Missouri Botanical Garden, P. O. Box 299, St. Louis, MO 63166-0299. Periodicals postage aid at St. is, MO, and additional entry. Claims for missing issues, made 6 months (domestic) to 12 months (foreign) after the date of issue, and orders for back issues should be addressed to Dr. James D. Montgomery, Ecology II, 804 Salem Blvd. oe PA 18603-9801. Changes of address, dues, and applications for membership should be sent to the Membership General inquiries concerning ferns should be addressed to the Secretary. Subscriptions $35.00 to U.S.A., Canada, and Mexico; $45.00 i oo in the world (—$2.00 agency fee, if applicable); sent free to members of the American Society (annual dues, $25.00 + .0O0 mailing surcharge beyond U.S.A., Canada, and Mexico; Pa eens. $300.00 + $140.00 mailing surcharge beyond U.S.A — ae and Mexico Back vol yon olumes are available for most years as printed issues or on microfiche. Please contact the Back Issues Curator for prices as; availability. POSTMASTER: Send address changes to AMERICAN FERN JOURNAL, Missouri Botanical Garden, P. O. Box 299, St. boas MO 63166-0299 FIDDLEHEAD FORUM Bloc: editor of the Bulletin of the American Fern Society oo contributions from members a members, including miscellaneous notes, offers to exchan one x purchase materials, aE feat ceeak notes, and reviews of non-technical books on fern SPORE EXCHANGE ia Mandt, 12616 Ibbetson Ave., Downey, CA 90242-5050, is ae. Spores exchanged and ae a available spores sent on request. _http://amerfernsoc.org/spo mil GIFTS AND BEQUESTS Gifts and bequests to the Society enable it to expand it ices to members and to others interested in ferns. Back is Reems ik die 1 4 + ot. -o ‘. 1 4 - rs deductible o" ig ees * oe 2 1. 2% ae ets g ie MISSOURI BOTA: -AL JAN 1 4 2003 American Fern Journal 92(4):249—269 (2002) GARDEN LIBRARY The Morphological and Genetic Distinctness of Botrychium minganense and B. crenulatum as Assessed by Morphometric Analysis and RAPD Markers Linpa M. Swartz! and STEVEN J. BRUNSFELD” Department of Forest Resources, University of Idaho, Moscow, ID 83844-1133 AsstTract.—Two species of Botrychium subgenus Botrychium (moonworts, Ophioglossaceae), Botrychium minganense Victorin and B. crenulatum W. H. Wagner, can sometimes be confused in the field, even by experts, because of their reduced morphology. Botrychium minganense can imitate B. crenulatum, which is more rare. They are afforded different degrees of protection on Federal lands, making the distinctness of these species a question of management, conservation, and systematic interest. The purpose of this study was to compare a morphometric analysis of these two species with an analysis of DNA markers from the same individuals, and to assess their distinctness under each method. Collections were made in Washington, Oregon, Idaho, and Montana from seven auger i - B. ehidpessdess and 18 populations of B. minganense. Each plant was measured, emph g cited by authors in the original species descriptions. Canonical variate analysis ser on SAS separated the samples into two oe groups with 32% overlap. RAPD genetic markers revealed more genetic variation than has previously been UPG be confirmed or ruled out with markers from one or two RAPD primers. Both B. crenulatum and B. Junaria have been suggested as possible diploid parents of tetraploid B. minganense. All RAPD markers absent in B. crenulatum but present in B. minganense were also present or polymorphic in B. lunaria, supporting B. /unaria as a possible parent. One very small population of B. minganense showed a monomorphic RAPD profile, consistent with inbreeding, but all other populations had multiple genotypes. Some plants of B. minganense clustered most closely with plants from populations up to 400 km away, suggesting that variation may be introduced into populations by occasional colonization by spores from distant sources. Species of Botrychium subgenus Botrychium (moonworts, Ophioglossaceae) are an enigmatic part of the temperate flora, notable for their small size, reduced morphology and difficult identification. Several moonwort species in North America are listed as sensitive or rare because of small and/or few known populations. Small populations are more vulnerable to extirpation, whether from natural stochastic events or human activities. Understanding the threat to species requires accurate information on numbers of individuals and " Current address: USDA Forest a Hiawatha National Forest, St. Ignace Ranger District, 1798 West Highway 2, St. Ignace, MI 4 * Corresponding author. 250 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) populations. If species intergrade morphologically, questions can arise not only about actual numbers of individuals and populations but also about species boundaries and the genetic distinctness of each species. All moonworts are relatively small and bear a single leaf with a fertile segment (sporophore) and sterile segment (trophophore) each season from an underground bud. They are notoriously hard to find, especially in thick vegetation. As increasing emphasis has been focused on rare plants in recent decades, more concentrated searches have extended the known ranges of common Botrychium species and provided material from which 13 new species have been described since 1980. Several of these are endemic to western North America: B. crenulatum W. H. Wagner, B. echo W. H. Wagner, B. lineare W. H. Wagner, B. montanum W. H. Wagner, B. paradoxum W. H. Wagner, B. pedunculosum W. H. Wagner, B. pumicola Coville, and B. pinnatum H. St. John (the latter two described in 1900 and 1929, respectively). Distinguishing moonwort species in the field often depends on subtle differences in phenology, color, texture, proportions of the parts of the single leaf, and dissection of the pinnae. Such species, poorly morphologically differentiated but evolutionarily distinct, have been called cryptic species (Stebbins, 1950; Paris, Wagner, and Wagner, 1989; Hauk and Haufler, 1999). Although species differences may be subtle, some species are also quite variable among regions, among sites, and even within the same site (e.g., Wagner and Lord, 1956). One of those species is B. minganense Victorin. This study was initiated in response to the practical need to distinguish between B. crenulatum and B. minganense. These two species have been confused in the western United States by many botanists (Zika, 1992). Although both have been listed as “sensitive” in the past by National Forests in the Pacific Northwest Region (Region 6) and the Northern Region (Region 1), B. minganense has been delisted in Region 6 in response to the discovery of many more populations, while B. crenulatum retains its official status as rare. Species designation affects management options where B. crenulatum occurs. The documented distribution of B. crenulatum is the mountain states of the American west (Arizona, California, Idaho, Montana, Oregon, Nevada, Utah, Washington, and Wyoming), whereas B. minganense is widespread in the western mountains and across northern North America (Wagner and Wagner, 1993). Botanists have employed both lumping and splitting approaches to the confusing variability of B. minganense. Botrychium minganense has been interpreted by many authors as a variety of B. Junaria (L.) Sw. (see Wagner and Lord, 1956 for discussion). In Flora of the Pacific Northwest (Hitchcock and Cronquist, 1973), only five moonworts are recognized, and the taxon to which B. minganense keys is called B. Junaria var. onongadense (Underw.) House. Cronquist said that B. minganense is “...morphologically scarcely separable from diploid var. onongadense...” and considered it conspecific with B. Junaria (Gleason and Cronquist, 1991). Botrychium minganense is a currently accepted taxon (International Taxonomic Information System database http:// www. itis.usda.gov/plantproj/itis, April 15, 2000; Kartesz, 1994). In the most SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 251 recent treatment of North American moonworts (Wagner and Wagner, 1993), B. minganense is reported to be sometimes misidentified as B. dusenii of South America. It is also easily confused with B. Junaria (Wagner and Lord, 1956; Farrar, 1998), B. ascendens (Zika, 1992; Farrar, 1998), B. pallidum (Zika, 1992), B. spathulatum (Zika, 1992), and B. crenulatum (Wagner and Lord, 1956; Lellinger, 1985; Wagner and Devine, 1989; Zika, 1992; Farrar, 1998). Zika (1992) described B. minganense as “‘treacherously variable’’. Just as B. minganense was recognized as an independent species from the more widespread and common B. lunaria, so too were B. pallidum and B. spathulatum formerly confused with B. minganense. Both Wagner and Wagner (1988), and Wagner (1994) have suggested that B. minganense may represent a species complex. Unlike B. minganense, B. crenulatum is more constant in form when well developed, but as with any moonwort, the identity of small plants can be ambiguous. Botrychium minganense can approach the form of B. crenulatum closely. Wagner and Wagner (1981) state that some of the collections on which the original description of B. crenulatum was based were originally identified as B. lunaria var. minganense. Botrychium crenulatum is diploid (2n = 90, F. S. Wagner, 1993), whereas B. minganense is tetraploid (2n = 180 Wagner and Lord, 1956; but see Hauk and Haufler, 1999). Many fern species, however, have races with different ploidy levels (e.g. Asplenium trichomanes, Wagner et al., 1993), and ideally, additional evidence of genetic differences would be employed to separate species (for discussion, see Gastony and Windham, 1989). Molecular techniques are well suited to clarify problems of cryptic species. Hauk (1995) used rbcL sequences in a phylogenetic analysis of 20 species of Botrychium subgenus Botrychium. Hauk found that four samples of B. minganense (from Michigan, Colorado, and Ontario) shared identical sequences, along with B. paradoxum and B. Xwatertonense, and lacked the single synapomorphies that distinguished the simplex and campestre subclades of the “‘simplex-campestre” clade. Botrychium crenulatum formed a separate clade with B. /unaria, identical in sequence to the United States B. lJunaria sample, and well separated from the “simplex-campestre” clade by a total of nine substitutions. In contrast to the rbcL data, which did not distinguish B. crenulatum from B. Junaria, isozymes differentiated B. crenulatum from all others (Farrar, 1998: Hauk and Haufler, 1999). Among the sampled diploids B. crenulatum was most similar to B. /unaria, but their genetic identity (Nei, 1978) was only 0.53 (Hauk and Haufler, 1999). Botrychium minganense possessed the highest variability of the western moonworts (Hauk and Haufler 1999, Farrar 1998), but neither study inferred the variation to be indicative of species-level differentiation within B. minganense. Random Amplified Polymorphic DNA (RAPD) markers, a type of genetic fingerprint, have revealed a level of genetic variation useful for distinguishing populations and sometimes species, and typically possess more variation than isozymes (for reviews, see Bachmann, 1997; Crawford, 1997). RAPD has been particularly useful in assessing variation in rare plants because, as a PCR-based 252 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) technique, it requires only small tissue samples, fresh or dried. DNA markers such as RAPD may provide important information for critically assessing morphometric analyses in taxonomically confusing groups. Morphometric analysis can suffer from a circular logic in which a taxon exhibits a certain range of morphologic variation because of assumptions made in the assignment of specimens to that taxon. Assigning specimens based on genetic markers can provide more robust morphometric insights, as has been shown in a variety of studies (e.g., Hardig et al., 2000). Thus, the goals of this study are 1) to determine the genetic distinctness of B. minganense and B. crenulatum, on the basis of RAPD markers 2) to document patterns of genetic variation within B. minganense and B. crenulatum, on the basis of RAPD markers, and 3) to assess quantitatively the morphological differences between plants of B. minganense and B. crenulatum classified on the basis of genetic markers. METHODS Collections.—Samples were collected from seven populations of Botrychium crenulatum and 18 populations of B. minganense in the states of Washington, Oregon, Idaho, and Montana (Table 1). Within this region populations were chosen to include a full range of habitats and geography. Plants with morphology intermediate between the two species were collected when found, and small plants were collected as well as large, well-developed ones to represent a full spectrum of the morphology found in each population. Plants were collected throughout the spatial extent of each site. Sample sizes are given in Table 1. In addition, two populations of B. Junaria and one of B. simplex E. Hitchcock (both subgenus Botrychium) were collected to provide a larger sample of species level molecular comparisons. The ecological associations of B. minganense and B. crenulatum were quite different in different parts of their ranges. Botrychium crenulatum in Washington is sometimes found in somewhat wetter and more open habitats than B. minganense, but the large populations sampled for this study were all growing under a Thuja plicata/mixed conifer canopy on subirrigated ground. By contrast, the Goofy Springs, Oregon, population was growing in heavy graminoid cover in an opening on seepy ground; the Stewart Creek, Montana, site was a wet mowed roadside and ditch; and at Lapover Ranch, Oregon (the one site on private land), B. crenulatum and B. minganense were growing together in grass cover under Pinus contorta with no spring evident. In Washington, B. minganense was found almost exclusively under riparian Thuja plicata stands with depauperate understory, but one sampled population (Mill Gate) came from an herbaceous mountain meadow. The association with Thuja plicata stands may be an artifact of the circumstance that moonworts have mainly been searched for in association with proposed timber sales. In Oregon, the sampled B. minganense populations were all under forest canopy open enough to support a luxuriant shrub and/or herb layer, except Dusty, which was from a wet meadow. SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES TABLE 1. segregated under more tha occurred at a single site, each species is listed as a minganense, c = B. crenulatum, L = B. lu 253 Collection locations of Botrychium used in this study. ania eng from some sites were n one collection number. Where more than one analyzed — population “ae an “aries letter naria). Vouchers are deposited in I Site Site Species (no. of plants) abbreviation Location Voucher B. minganense Watson Point (2) Watson OR, Wheeler Co. Swartz 387 wery Trail (6) Flowery WA, Stevens Co. Swartz 393 Kelsey Creek (6) Kelse MT, Lincoln Co Swartz 394 Rock Bottom (7) Rock ID, Boundary Co Swartz 398 Deer (7) Deer ID, Boundary Co Swartz 399 Wenatchee Ford WenT WA, Chelan Co. Swartz 401A TSHE (7 Wenatchee Ford WenR WA, Chelan Co. Swartz 401B PA (5 Devil’s Club Devil WA Chelan Co. Swartz 402 Creek (7) Mill Gate (17) Mill, MillIB WA, Chelan Co. Swartz 403, Swartz 453 Aladdin 1 (6 Aladdin1 WA, Stevens Co. Swartz 414 Bulldog Cabin (5) Bulldog WA, Stevens Co Swartz 420 Poison Springs (7) Poison OR, Grant Co Swartz 425 m Hodgson Creek (7) mHodgson WA, Ferry Co. Swartz 466 m Rd. #9576 (6) mRd9576 WA, Ferry Co. Swartz 468 ManleyX Swartz 486 m Manley Creek (15) mManley WA, Ferry Co. Swartz 470 La Grande 32 (6) aG32 OR, Union Co Swartz 504 Shady Camp (7) Shady OR, Wallowa Co Swartz 506 Dusty (6) Dusty OR, Union Co. wartz/ Riley 508 Swartz/ Yanskey 509 B. crenulatum — Goofy Spring (7) Goofy OR, Crook Co. Swartz 38 Stewart Creek (5) Stewart MT, Flathead Co. Swartz 396 Okanogan Cabin (7) OKCabin WA, Okanogan Co, Swartz 404 Aladdin Aladdin WA, Stevens Co. Swartz 427 Blowdown (7) Deadman Creek (5) Deadman WA, Ferry Co. Swartz 445 c Hodgson Creek (7) cHodgson WA, Ferry Co. Swartz 467 Lapover Ranch (10) apove OR, Wallowa Co. Swartz 507 B. lunaria L Rd. #9576 (5) LRd9576 WA, Ferry Co Swartz 469 L Manley Creek (7) LManley WA, Ferry Co Swartz 471 B. simplex La Grande LGMead OR, Union Co. Swartz 505 Meadow (5) Plants were collected by snipping them off at ground level to avoid disturbing the roots and the next year’s below-ground bud. This procedure is not believed to have a significant negative impact on survival (Johnson-Groh and Farrar, 1996; Montgomery, 1990). Where possible, plants were collected after they had shed spores. Plants were pressed, individually numbered, color photocopied, and digitally imaged before grinding for DNA extraction. 254 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) The color photocopies are deposited in the University of Idaho Herbarium (ID) as facsimile vouchers, along with additional collections from the same populations. Morphometric analysis.—As a quantitative approach to capturing morpho- logical subtlety, a morphometric analysis of characters that can be scored from herbarium specimens was made. Characters cited by authors in the original species descriptions were used whenever possible. Some characters that are valuable to botanists in the field, such as color, texture, or folding of pinnae, could not be scored because they are distorted or destroyed by pressing and drying. Forty-one different measurements or ratios were recorded for each plant. Measurements were made using a Panasonic WV-CD20 video camera and Mocha image analysis software (SigmaScan Pro version 3.0 Jandel Scientific). Analysis.—Canonical discriminant analysis identifies one or more canonical variables that are linear combinations of multiple measured characters. These canonical variables can show the greatest morphological differences between groups. Statistical calculations were performed using the SAS CANDISC procedure, SAS Release 6.11 (SAS Institute Inc.), on the same samples of B. minganense and B. crenulatum used for genetic analysis, excluding three plants that were browsed. Two very unusual plants of Botrychium minganense also were excluded from the morphological analysis. One was extremely large, and the other had only rudimentary peg-like pinnae. One plant (cLapover.09) was excluded because it displayed an additive RAPD profile, and thus was pos- sibly a hybrid. Preliminary one-way ANOVA showed that the means of many characters were significantly different between the species at the alpha = 0.05 level, including the ratio of trophophore width to the width of its axis (reflecting the tendency of pinna margins to be decurrent on the rachis); average angle of the margins of the four basal-most pinnae (degree of fanning); ratio of the length of the space between the first two pinnae pairs to greatest pinna width (a measure of overlapping of pinnae); ratio of greatest pinna width to least pinna width; ratio of length to width of trophophore; ratio of length to width of sporophore; ratio of pinna width to length; length of the sporophore; average angle made by the four basal-most pinnae with the rachis; total height (ground level to tip of sporophore); length of trophophore; length of trophophore stalk; and length of gap between first two pinna pairs. Measure- ments of these characters are illustrated in Figure 1. All pinna measurements were made from the same pinna for each plant, one of the largest pair. In the largest plants the lowest pinnae are sometimes partly transformed into sporangial branches. In that case one of the largest untransformed pair was chosen. The ratios of 1) trophophore width: trophophore axis width and 2) maximum:minimum pinna width were log-transformed, and the length of sporophore was square root-transformed to bring them closer to a normal distribution. The distribution of all variables used was judged to be within the limits of robustness of the procedures (K. Steinhorst, pers. comm.). Variances were compared between species groups for each character to see that they were equal, or if not, the variance of the larger group did not exceed that of the SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 255 Fic. 1. Measurements made for characters that were significantly different between Botrychium minganense and B. crenulatum in morphometric analysis. a. Length of common stalk. b. Length of sporophore stalk. c. Length of sporangia-bearing part of sporophore (b+c = length of sporophore, a+b+c = total height. These and any other curved lengths were traced directly on the image of the plant). d. Length of longest sporophore branch (2d = sporophore width). e. Length of trophophore stalk. f. Distance between centers “ fast are pena pairs. g. Balance of length of trophophore (e+f+g = length of trophophore). h inna (for average of four basal-most pinnae). i. Angle at which pinna meets axis of rachis i average of four basal-most pinnae). j. Greatest width of rachis. k. Trophophore width. |. Least width of largest pinna m. Greatest width of largest pinna. n. Length of largest pinna. smaller group by more than a factor of 2.5, a conservative level chosen for unequal sample sizes. In general, variances were greater for B. minganense. RAPD analysis.—DNA was isolated from 10 mg samples of each pressed plant. For those plants that were less than 10 mg, the whole plant was used. Plants were ground on ceramic well plates with liquid nitrogen, ground further 256 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) with 600 ul 70°C CTAB buffer, and transferred to 1.5 ml tubes. The grinding buffer and subsequent isolation procedures followed Stewart and Via (1993), with the following modifications: the homogenate was incubated at 70°C for 30 minutes before the chloroform extraction, the precipitated DNA pellet was washed with 1 ml cold 76% ethanol with 10 mM NH,AC, and the dry pellet resuspended in 50 ul TE. Of several tested, this protocol was the least likely to yield gummy residues coprecipitating with the DNA, which was a problem with some samples. The residue, when it occurred, was removed by centrifugation before quantifying the DNA with a fluorometer. DNA was amplified (Williams et al., 1990) in 25 ul reactions containing 1x buffer (Promega M190A), 0.1 mM of each deoxynucleotide, 2 mM MgCl, 0.00005 % bovine serum albumin, 5 pmols 10-mer primer (Operon), 10 ng genomic DNA, and 0.5 units Taq DNA polymerase (Promega), overlaid with 25 ul mineral oil. Samples were amplified in an MJ Research PT 100 thermocycler (44 cycles of 1 min at 94°C, 1 min at 36°C, 2 min at 72°C, with a final 5 min at 72°C). Products were electrophoresed in 1.5% agarose gels, visualized by UV illumination after staining with ethidium bromide, and imaged with Alphalmager v. 3.2 software. Populations were divided among multiple PCR runs, and a sub- sample was run multiple times to confirm repeatability of each band chosen. Bands were scored manually by comparison to standard size markers. Bands are designated by the name of the primer with the approximate size in base pairs as a subscript, e.g. B-11575. Primer screening.—Primers were screened against two samples each of B. minganense and B. crenulatum. Twelve primers (A-11, B-11, B-12, C-6, C-8, C- 9, C-10, C-11, D-11, D-16, D-20, X-1) showing the best well-spaced bands polymorphic in one or both species were selected for the final data set. One hundred ninety-four plants were scored manually for presence or absence of 74 RAPD bands each. As more species and populations were added, fewer primers and bands within primers could be used because some new bands were Close to the position of old bands or amplified with different intensity, making them difficult to score. Therefore, the scoring is conservative and reflects only minimum differences among all populations, whereas man additional differences that are not included in the data set are readily apparent among individuals and populations in the same gel. Cluster analysis of RAPD data.—UPGMA cluster analysis was performed on RAPD data with NTSYSpc version 2.02 (Rohlf, 1997) using simple matching and Jaccard metrics. RESULTS Morphometric analysis.—Optimal separation of the two species on a mor- phological basis requires consideration of multiple characters at once. The six variables in Table 2, when analyzed together, provided the greatest separation of the groups in this data set. The canonical variate analysis tested the null hypothesis that there are no differences between the two species based on the chosen variables. This hypothesis was rejected at the p = 0.0001 level, with SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 257 E 2. Correlations and coefficients of the six variables that provided the greatest discrimination between Botrychium minganense and B. crenulatum in Canonical Discriminant Analysis. Pooled Within Pooled Within-Class Canonical Structure Canonical Coefficients Variable CAN1 AVANGMAR 0.50 0.65 AVANGPIN 0.16 0.40 WIDLEN 0.20 -0.18 MLSPORO —0.18 —0.94 NTWIDAX 0.51 0.51 NPINMAMI 0.41 0.52 AVANGMAR ge angle of pinnae margins; AVANGPIN=average angle of pinnae with rachis; PWIDLEN=ratio of pins width: pinna mip ESO square root transformed length of sporophore; NTWIDAX=log-transformed t width of trophophore axis; NPINMAMI=log-transformed ratio of greatest pinna width: least pinna width. F = 35.52 and degrees of freedom of numerator 6 and denominator 164. Canonical scores may be computed by taking the original value for each plant on each measurement, multiplying it by the respective canonical coefficient from CAN1 (Table 2), and adding all these products plus a constant adjustment for the means. Scores graphed by species form two overlapping groups, with the mean for B. crenulatum at 1.81 and the mean for B. minganense at —0.71 (Fig. 2). CAN1 scores of 55 of the total of 171 plants, or 32%, fell in the 0 to 2 range where species identity is ambiguous. Scores of 92 plants of B. minganense out of 123 (75%) fell in the 0 to —4 range, and 23 plants of B. crenulatum out of 48 (49%) scored from 2 to 4, where each had a high probability of correct species identity. Only one B. minganense had a CAN1 score above 2 Canonical variates can be interpreted in terms of those variables that contribute the most to the separation of the groups. Although canonical variates are artificial and must be interpreted with caution, they can be identified in terms of their correlations with the original individual variables (Johnson and Wichern, 1992). These ‘‘within” structure coefficients indicate how closely a variable and the canonical variate are related, or the extent to which they carry the same information (Klecka, 1980). The ratio of trophophore width:trophophore axis width had the highest correlation, 0.51, followed by average angle of pinna margins, 0.50, and pinna maximum width:minimum width, 0.41. The within-class correlation for pinna width: length was 0.20, average angle of pinnae to rachis 0.16, and length of sporophore —0.18. Another way of looking at the contributions of each individual variable within classes is by comparing coefficients that have been transformed so their standard deviations are equal to 1. These standardized coefficients then measure the relative contribution of each variable to the canonical variate score. As a relative measure, the standardized coefficient of each variable will change depending on the contribution of other variables. If two variables share ses sini pes poeta fa ium minganense E Botrychium crenulatum Spp. 0 Botrych Ei Both eee S38 Bas penoanmons se sais i etree a8 a pach a iables (Table 2) of 123 Botrychium 1c var ically. ity was confirmed genet ies ident the standardized co- if one variable was not y could also be t the other. The > correlated) la but ’ The ise. ially cancels ou Tl part 10nS iate correlat lvar % a res ‘ fons i ee 2 ae ss sooiects ee 20 ee : poet ie tart i as asia se pee coat He see anima rise enrncncinese ice inunnnens ae pee prcnnny te nee SRR inna sees uusuanaannanenamnanmnmmninsaa me omen EERIE REE ENE TET Boe enaier 3 spaonons PEERED EDIE SEE See Stone ighly etween them AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) 258 see PON MN N08 se oe iene aisaasasanssensainty incon sia pee se pi ee eh nod . i eee phone z Raaeanart Sues apananie oe saat pies at sia a Be ice Sie —— eee ee ae si Bs Aouanbe1J sa ti ee sy es Se Saal ix morphometr ing s d 48 B. crenulatum plants whose spec Plot of CAN1 scores us minganense an Fig. 2. ion (are h some of the same informat efficient value will be partly d ided b 1V t of the other would so that one ts, by contrast, are simple b 1en used, the standardized coeffic larger but have oppos > ite sign within structure coefficien SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 259 (Klecka, 1980). The standardized coefficient for length of sporophore, —0.94, had the highest absolute value among the variables. This was the measurement reflecting total size of the plant that contributed most to separation of the species. When it was included, other direct measures of size, such as total height, had small absolute values. Average angle of pinna margins, 0.65, and ratio of pinna maximum:minimum width, 0.52, are each related to the fanning of the pinnae in different ways, and they do not cancel each other out. The pooled within-class standardized canonical coefficient for ratio of ttophophore width:trophophore axis width was also high, 0.51, and average angle of pinnae, 0.40, and ratio of pinna width:length, —0.18, were lower. RAPD diagnostic bands.—Botrychium crenulatum was most differentiated (Table 3), set apart by seven bands (B-111 400, C-6g25, C-8g50, C-81700, C-91180, C- 104275, and D-11,975) that did not occur in the other three species. One band (B- 114975) was present in B. crenulatum, absent in B. minganense and B. simplex, and polymorphic in B. Junaria. One band (C-9;909) was present in B. crenulatum, absent in B. minganense, and polymorphic in B. Junaria and B. simplex. Bands not present in B. crenulatum included six that were present in all individuals of B. minganense. Of these, three (C-11¢75, D-16775;, and D-20g99) were polymorphic in B. Junaria and present in all B. simplex, two (B-11575, C- 9690) were polymorphic in B. Junaria and absent in B. simplex, and one (D- 114225) was present in B. Junaria and polymorphic in B. simplex. Four bands not present in B. crenulatum were present at high frequency (0.97—0.99) in B. minganense. Three of these (D-11375, D-16400, D-165109) were polymorphic in B. lunaria and absent in B. simplex, and one (D-11;399) was polymorphic in both B. lunaria and B. simplex. No bands in the sampled plants were unique to B. minganense or B. lunaria, and one band (C-6459) was seen only in B. simplex. Bands common to all four species were not scored. Clustering —UPGMaA clustering of the RAPD data using a simple matching metric resulted in four well-defined species groups (Fig. 3). The B. crenulatum cluster was most distinct. The B. simplex cluster and the B. Junaria cluster grouped together, and the “simplex lunaria’”’ cluster associated most closely with the B. minganense cluster. Use of a Jaccard metric, discounting 0/0 matches, produced relationships conforming to those discussed below, except that the B. Junaria cluster associated most closely with the B. minganense cluster, and the B. simplex cluster grouped with the ‘‘minganense lunaria” cluster (dendrogram not shown). The B. Junaria and B. simplex clusters are displayed in Fig. 4. Botrychium crenulatum.—Within the B. crenulatum group (Fig. 5), the largest cluster contained all the plants from Washington except two. This Washington cluster contained four subgroups within which plants had identical profiles. One subgroup included five OKCabin plants, and three contained plants from Deadman, cHodgson, and/or Aladdin. All B. crenulatum populations were polymorphic. Associated with the Washington cluster was a cluster that contained samples from Montana (Stewart) and Oregon (Lapover), plus two genetically distinct plants from the Hodgson population from northeastern Washington, which includes both B. minganense and 260 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) TaBLE 3. Diagnostic RAPD bands. Band name B. crenulatum B. minganense B. lunaria B. simplex B-Ligyst# 0 1 P fy) B-111075** p 0 Pp 0 B-114400*** 1 0 0 0 C-6450 0 0 0 1 C-6395*** 1 0 0 0 C-8g50* ** 1 0 0 0 C-84700* ** 1 0 0 0 C-Secot# 0 1 P 0 C-94000* 1 0 P | C-9, 180* i 0 0 0 C-104275*** 1 0 0 0 C-11675tV 0° 1 F 1 -11g00* 1 0 P 1 D-11g75t# 0 P(1) P 0 D-114975*** 1 0 0 0 D-114225fO 0 1 1 P D-11) 3007 0 P(1) is P D-16490t# te P(1) ly 0 D-165;0t# 0 P(1) P 0 D-16775tV 0° 1 P(1) 1 D-20g90fV 0 1 P 1 1 = present, 0 = absent, P = polymorphic, P(1) = polymorphic, present very high frequency. *** Present in B. crenulatum, absent in B. minganense, B. lunaria, B. simplex. ** Present in B. crenulatum, absent in B. minganense, polymorphic in B. Junaria, absent in B. simplex. * Present in B. crenulatum, absent in B. minganense, polymorphic in B. Junaria, present or polymorphic in B. simplex. ¢ Present in B. minganense, absent in B. crenulatum. + Polymorphic at high frequency in B. minganense, absent in B. crenulatum. # Polymorphic in B. Junaria, absent in B. simplex. V Polymorphic in B. lunaria, present in B. simplex. O Present in B. lunaria, polymorphic in B. simplex. “ Also present in Lapover9. B. crenulatum. One of these individuals, ‘‘m’Hodgson.05, was classified in the field as B. minganense, but clearly groups genetically with B. crenulatum. The most distinct group was formed by Oregon plants. Goofy was the only population to form an exclusive cluster. Goofy and Lapover grouped together, but some members of Lapover also clustered with Stewart, the Montana population, in the mixed cluster. cLapover.09 was the most dissimilar member of the B. crenulatum cluster, and in fact displayed three bands otherwise found only in B. minganense, B. lunaria, or B. simplex (C-11¢75, D-16499, and D-16,75) as well as all the bands displayed only in B. crenulatum. Botrychium minganense.—In the B. minganense group (Fig. 6), three populations formed exclusive clusters: Watson (Oregon), Poison (Oregon), and Devil (Washington). All others formed mixed groups. Thirteen plants from the Idaho Panhandle and neighboring Montana populations (Rock, Deer, and Kelsey) had identical profiles. Some members of each of those populations SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 261 | B. crenulatum | B. minganense | B. lunaria = B. simplex eo Se T ma aes T is T | ae, T T T a oo | a amas | 0.72 Similarity coefficient Fic. 3. Species clusters from UPGMA dendrogram of RAPD data from a sg of 194 plants of four Be B, moonwort species: Botrychium crenulatum, 128 B. minganense, 12 B. lunaria, and six simplex. The scale represents the similarity coefficient between iar a s. clustered with other groups. One population, Aladdin1, was monomorphic. Members of WenR and WenT, which grew adjacent to each other, each grouped in a separate larger cluster. Mill, from the Washington Cascades, and Manley, from northeastern Washington, were particularly diverse: each had members in four different larger clusters, and other members that were highly divergent. DISCUSSION Morphometrics.—The cryptic moonwort species Botrychium minganense and B. crenulatum can be separated by canonical variate analysis into two partially overlapping groups. Plotting the CAN1 scores of 171 plants whose identity had been genetically confirmed showed that 32% fell in the zone of overlap where the two species could not be separated using the characters scored. The characters that contributed most to the separation were measures related to pinna shape, proportions of the trophophore, and size. Pinna fanning, as reflected in the pinna shape characters, is emphasized in descriptions of B. crenulatum (Wagner and Wagner, 1981; W. H. Wagner, 1993). Average size of B. minganense is larger than that of B. crenulatum (mean height of sampled plants 84 mm and 73 mm respectively), but each can be less than a centimeter tall (pers. obs.). The ratio, width of trophophore: maximum width of trophophore axis, is a complex character that combines elements of several differences between the species, including length of pinnae, angle at which the pinnae meet the axis, and tendency of pinnae to be decurrent on the trophophore axis or of the axis to be flattened. These characteristics have not been emphasized in taxonomic descriptions of B. minganense and 262 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) LRd9576.01 s ead. sLGMead.05 aaa 0.44 0.58 0.72 Similarity coefficient Fic. 4. Detail of Botrychium lunaria and B. simplex clusters from UPGMA dendrogram of RAPD data (Fig. 2). Plants are labeled by population abbreviation (Table 1) and individual number within population. Leading letter signifies species (L = B. Junaria, s = B. simplex). The scale represents the similarity coefficient between clusters. B. crenulatum, although the trophophore of B. minganense has been described as narrow (Wagner and Wagner, 1993). The statistical analysis of morphology was limited compared to field identification. Some useful morphological characters of live plants could not be used in an analysis of herbarium specimens. The characters that could not be captured include phenology, color, texture, and many aspects of plant habit, such as cupping of the pinnae. Other useful characters, such as the number of pinna pairs, or number of crenulations on pinna margins, are not nor- mally distributed and therefore violate the requirements for canonical variate analysis. The 32% ambiguity rate in the morphological analysis contrasts with the correct field identification of all but seven of the 171 analyzed plants. However, in the field, individual plants are not independently identified, as is the case with the statistical analysis. Field botanists generally examine the range of variation at a site and make an identification on the basis of a group of typical plants. Some well-developed plants will show characters that smaller ones lack. This is also true of herbarium identification. In fact, consulting botanists request collections of about a dozen plants for identification of moonwort species (Wagner, 1992; Wagner and Wagner, 1993; Zika et al., 1995). This study generally supports the assumption that similar plants associated at one site belong to the same species. In the genetic analysis, none of the populations identified in the field as containing B. minganense and not B. crenulatum, or B. crenulatum and not B. minganense, contained sampled individuals of the other species. However, even experienced botanists can occasionally be misled (example given in Farrar, 1998). Although we sought mixed populations for this study, only three of 23 (Hodgson, Manley, Lapover) had mixed B. minganense and B. crenulatum populations. SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 263 cGoofy.01 | 2 RARE uosdIO cDeadman.06 cHodgson.03 c : cHodgson.07 | & > Ser ht oss 556 g apeooes vA ee CERES REE RO BG S566s55e5S PP oor Ss! ee: ge 2 Seal eae he UO}FUTYSE AY Aladdin. cLapover.09 aA ons 8 ae ee T T T T T T T 1 aL T T 1 0.44 0.58 0.72 0.86 1,00 Similarity coefficient Fic. 5. Detail of Botrychium crenulatum cluster from UPGMA dendrogram of RAPD data (Fig. 3). Plants are labeled by population abbreviation (Table 1) and individual number within population. Leading letter signifies species as identified in the field (c = B. crenulatum, “‘m” = B. minganense as identified in the field). The scale represents the similarity coefficient between clusters. RAPD analysis: Interspecific variation.—In contrast to the morphometric results, all sampled plants, with one exception (cLapover.09), grouped clearly by species based on RAPD markers. Primers C-10 and/or D-11, run with a known B. crenulatum sample, would be sufficient to confirm or rule out the identification of B. crenulatum. Separation of B. minganense from B. lunaria by means of RAPD markers requires more primers; because neither has unique bands and they are separated in the similarity analysis by differences in 264 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) mWatson.01 mF lowery.02 ie b anes 5 a mM G3 mbower “ é ock. mS Ab) mWenR.28 mLaG32.02 mManley.03 mShady.04 mManley.05 mShady.02 mFlowery.06 mShady.03 mFlowery.08 mShady.06 mF lowery.09 mMiullB.01 mFlowery.07 Sena net i ulldo mese33 10 mBulldog 07 mr ae33 05 Wen 39 aG32. e mAladdin1.01 mataddint 03 mBulldog.o4 inl. ulldo mAladdin1.04 Idog matadgm!.t mRd95 1 inl.06 mRd9576.02 mHodgson.04 mRd9576.03 mHodgson.02 ey mManley 01 mRd9576.05 mi jodgson. 7 mKelsey.02 mHodgson.08 mKelsey.03 mHodgson.09 mKelsey.05 mHodgson. | 1 mRock:01 mManley. mRock.02 mManley. mRock.03 mManley. | mRock. mManl mRock.05 mManleyX.05 mDeer.03 mPoison.01 mDeer.04 mipowson.O¢ mieer.5 ‘olson. mDeer. Al mPoison.03 mDeer 08 mPoison.06 mKelsey.06 mPoison.09 mKelsey.08 mPoison. 10 mManleéy. 1 mMuill. 04 or. 06 mM mes 07 all. elsey. mMill.07 msrady 01 mDustyR.03 i mMill 13 mDustyR.01 mRock,06 mDustyR.02 mRd9576.04 mDuse 04 mManiey. 02 mDustyY O1 mManley 07 LS ELA eee ee ee mMillB.03 0.72 0.86 1.00 mMfaniey, 16 Similarity coefficient mWenT 02 mWenT .04 mWenT.06 mMill.06 B mWenT.05 mWenT.07 mWenT .03 mMill.09 mMill. 11 . mMill. 05 mMill.08 mMill. 10 A mMill. 14 mMill.01 : mDevil.03 B. minganense mDevil 04 peal t evil. mDevil.07 B mDevil.08 mDevil. 11 0.72 0.86 1.00 0.72 0.86 Similarity coefficient Similarity coefficient Fic. 6. Detail of Botrychium minganense cluster (c) and subclusters (a, b) from UPGMA dendrogram of RAPD data (Fig. 3). Plants are labeled by population abbreviation (Table 1) and individual number within population. Leading letter signifies species as identified in the field (m = B. minganense). The scale represents the similarity coefficient between clusters. SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 265 frequencies of bands. All B. simplex sampled had one band that appeared to be unique to the species, but more populations must be sampled to verify this as a species marker. The assignment of plants to species based on RAPD markers agreed with their classification in the field, with one exception. That plant, ‘“‘m’’Hodgson.05, had a CAN1 score of 0.88 in the morphometric analysis (Fig. 2) and fell in the range of minganense/crenulatum overlap. It was field-identified as B. minganense, but had the RAPD pattern of B. crenulatum. It was a small plant with non-crenulated pinnae, from a mixed population of B. minganense and B. crenulatum. RAPD analysis: Distribution of variation within species.—RAPDs provide greater evidence of variability in species of Botrychium than isozymes have. Hauk and Haufler (1999) reported 14 isozyme genotypes among 252 plants of B. minganense, and one genotype among nine plants of B. crenulatum. In this study, there were 100 RAPD genotypes among 128 individuals of B. minganense, and 28 genotypes among 48 plants of B. crenulatum. Within five populations of B. minganense (Watson, Bulldog, mHodgson, LaGrande32, and Dusty) and one of B. crenulatum (Deadman), no two sampled individuals had the same RAPD profile. The population showing the highest genetic similarity among individuals was B. minganense—Aladdin1. This small population of about ten plants growing in approximately 9 m*, was sampled heavily because it was morphologically ambiguous, and could not be identified to species in the field (W. H. Wagner, Jr., pers. comm.). The plants were small, light green, and displayed rather broadly fanned pinnae. Three of the sampled plants had CAN1 scores in the 0 to 2 range where the scores of species groups overlapped, and three had slightly negative scores in the ‘‘minganense” range. This was the only population that lacked within-population variation (scorable or unscor- able) on the gels, and might be described as clone-like. All other sampled Botrychium populations were larger, and had more than one RAPD profile. Most individuals from the Deer (Idaho), Rock (Idaho), and Kelsey (Montana) populations of B. minganense had the same RAPD profile. These populations were located within approximately 50 km of each other. However, proximity was not a good predictor of genetic similarity across all populations. The twin B. minganense populations from Wenatchee Ford, in Washington, grew about 30 m apart and were reported as one population on the Wenatchee Forest sensitive plant sighting form. They were kept separate in the analysis because one group was growing in deep shade in a riparian zone under Thuja plicata and Tsuga heterophylla, and the other was under Rubus parviflorus on a roadside. The Tsuga group clustered with the Mill population from a mountain meadow on the Wenatchee Forest, but the Rubus group clustered with Flowery and Bulldog, shaded forest sites from northeastern Washington, approximately 225 km away. No association between ecological sites and RAPD genotype is evident in this or other clusters. Although some populations, such as Goofy (B. crenulatum) and Devil (B. minganense), showed low similarity to any other, most populations had members in more than one cluster. For example, some B. minganense plants 266 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) from Mill clustered with plants from Manley in northeastern Washington, whereas others clustered with Shady and LaGrande32 in northeastern Oregon, about 400 km from the Mill site. Genetic variation within B. minganense did not suggest any coherent genetic groups that might be associated with its morphological variability. Hauk and Haufler (1999) reported more isozyme variability within B. minganense than any other polyploid sampled. Given that RAPDs are revealing more variability than isozymes, additional sampling from across the species range may reveal genetic patterns within the species. Genetic structure of populations.—The contrasting patterns of genetic similarity may result from processes that hinder or promote genetic isolation. Two important factors influencing the structure of genetic variation in plants are breeding system and dispersal of propagules. The breeding system of moonworts is not known from experimental investigations, because they have not been cultivated successfully. The most direct evidence comes from the allozyme work of Hauk and Haufler (1999) on other species of subgenus Botrychium. Low variability within populations hampered their inferences of breeding systems, but they attributed the low frequency of heterozygotes found in four populations of diploid moonworts (B. simplex and B. lanceolatum) to inbreeding. Electrophoretic studies on Botrychium species in subgenus Sceptridium (McCauley et al., 1985; Watano and Sahashi, 1992) and subgenus Osmundopteris (Soltis and Soltis, 1986) reported extremely high levels of inbreeding. Outcrossing may be hindered by the underground gametophytes of this genus (Tryon and Tryon, 1982), although moonwort hybrids have been reported (Ahlenslager and Lesica, 1996; Wagner, 1980, 1991; Wagner et al., 1984; Wagner, Wagner, and Beitel, 1985; Wagner and Wagner, 1988), demonstrating at least occasional outcrossing. The number of allopolyploids also documents that outcrossing is an important evolutionary process in subgenus Botrychium (Hauk and Haufler 1999). In our study, the lack of genetic diversity in the RAPD profiles of the small Aladdin1 population is consistent with inbreeding. Genetic variability within populations of an inbreeding species could be increased by immigration of propagules from distant sources, and occasional outcrossing. Fern spores are light and can travel long distances, as ferns colonize remote islands (Tryon, 1970; Tryon, 1986; Ranker et al., 1994). Tryon (1970) presented evidence that 800 km is not a significant barrier to the migration of a fern flora. Tryon and Tryon (1982) characterized Ophioglossa- ceae in particular as a colonizing group. Because of the dominant inheritance of RAPD markers (Bachmann, 1997; Crawford, 1997), these data do not provide unequivocal insights into the breeding system of moonworts. Although the variability detected in this study may not have been predicted based on isozyme studies, it is not inconsistent with a high dispersal rate and a largely inbreeding mating system. Ancestry of B. minganense.—The rbcL sequence of tetraploid B. minganense did not match that of any known diploid (Hauk, 1995). On the basis of the match between the hypothetical isozyme profile of the non-chloroplast parent SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 267 of B. minganense and the isozyme profile of B. crenulatum, Hauk and Haufler (1999), proposed B. crenulatum as the most likely candidate for that parent. The RAPD data, however, did not support this relationship, because B. crenulatum showed seven unique bands absent in B. minganense. An earlier hypothesis (F. S. Wagner, 1993) based on morphological data, proposed B. lunaria and B. pallidum as the parental diploids. The RAPD evidence is consistent with a close relationship between B. minganense and B. Junaria, because neither had bands that did not occur in the other. More genetic evidence is needed to clarify the origins of B. minganense. cLapover.09.—The identity of one plant from the Lapover site in the Lostine River Valley, Oregon, was uncertain when it was collected. It combined the color and luster of B. crenulatum with rounded, broad-based pinnae otherwise seen only in B. minganense. The RAPD profile of this plant included all seven diagnostic B. crenulatum bands, plus three characteristic B. minganense bands including C-1157; and D-16,75 (also polymorphic in B. Junaria and present in B. simplex), and D-16499 (polymorphic in B. Junaria and not present in B. simplex). Eight bands documented in all sampled B. minganense were not present in the plant. cLapover.09 appears to be a hybrid, both because of intermediate morphology and mixed markers. The three “‘minganense”’ markers could also have come from B. Junaria or B. simplex, but these species were not recorded from the site, whereas B. minganense and B. crenulatum were present. Other moonwort species recorded at the site were B. ascendens and B. lineare, for which we have no RAPD data. Neither of these moonworts typically has rounded pinnae. Because not all of the diagnostic B. minganense bands were present, cLapover.09 does not appear to be an F, hybrid between B. minganense and B. crenulatum, but is more likely a backcross or later generation hybrid derivative. CONCLUSIONS Although many plants of B. minganense and B. crenulatum could not be reliably distinguished by canonical variate analysis of morphology, all sampled plants, except an apparent hybrid, could readily be assigned to species on the basis of RAPD profile. This supports the distinctness of the two species although their morphologies intergrade. Although breeding system cannot be inferred from dominant markers such as RAPDs, higher levels of variability were detected within populations and species than might be predicted from previous genetic data, which suggested a high level of inbreeding. Thus codominant DNA markers such as micro- satellites might be a productive avenue for further research into breeding system and evolutionary processes in Botrychium. ACKNOWLEDGMENTS Thanks to Kathy Ahlenslager, Elroy Burnett, Leslie Ferguson, Joann Harris-Rode, Jerry Hustafa, Kirk Larson, Mick Mueller, Mark Mousseaux, Scott Riley, Faye Streier, Jim Vanderhorst, Kari 268 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) Yanskey, and Gene Yates for help with collecting. For illuminating the mystery of the moonworts, improved by the comments of two reviewers, W. Hauk and R. Moran, but any errors are the responsibility of the authors. This study was supported by the C. R. Stillinger Trust, University of Idaho. LITERATURE CITED AHLENSLAGER, K. and P. Lesica. 1996. pipe rem of Botrychium oso and its putative parent species, B. hesperium and B. paradoxum. Amer. BACHMANN, 28 1997. Nuclear DNA markers in pers biosystematic emer Opera Bot. 132: 137-14 CRAWFORD, 2 J. 1997. Molecular markers for the pea of genetic variation within and between uae of rare plants. Opera Bot. 132:1 Farrar, D. R. 8. Isozyme cies her er moonwort ferns (Botrychium subgenus panei aa their interspecific relationships in northwestern North America. Draft final Gastony, G. J. and M. D. WinpHaM. 1989. Species concepts in tat the treatment and definition of agamosporous species. Amer. Fern J. 79:65— GLeaAson, H. A. and A. Cronquist. 1991. Order ees. tee 13, in Manual of vascular lants of northeastern United States and Adjacent Canada, second edition. The New York ork. Haroic, T. M., S. J. BRUNSFELD, R. S. Fritz, M. Morcan and C. M. Orians. 2000. Morphological and molecular evidence for hybridization and introgression in a willow (Salix) hybrid zone. Mol. Ecol. 9:9-24. Hauk, W. D. 1995. A molecular assessment of relationships Bi oa species of Botrychium ieee Botrychium (Ophioglossaceae). Amer. Fern J. 8 an ER. 1999. Isozyme variability g CTY} i rol Bonychdum subgenus —633. = Hircucock, C. L. and A. Cronquist. 1973. Botrychium. Pp. 44-45, in diana of the Pacific Northwest, an Illustrated Manual. vag 3 of poucsa: jars Press, Seat JOHNsoN, R. A. and D. W. WIcHERN. 1992. sith multivariate oe analysis, third edition. Prentice Hall, Englewood Cliffs, New Jerse JOHNsON-GroH, C. L. and D. R. Farrar. 1996. Effects of leaf loss on moonwort ferns, Botrychium subgenus Botrychium. Amer. J. Bot. 83(6 suppl):127. wane co T. 1994. A synonymized checklist of the vascular flora of the United States, Canada, and eenland, second edition, Vol. 1-Checklist. Timber Press, Portland. R. 198 es ee D. B. 1985. Ophioglossaceae, Botrychium. Pp. 101-115, in A field manual of the ferns and oe -allies of the United States and Canada. Shalarcias Institution Press, Wash- ian See D. “" D. P. Wuirtier and L. M. REILLY. 1985. Spe ag rate of self-fertilization i Tape ap tg aati dissectum. Amer. J. Bot. 7 Ricestan LD 0. Survivorship and sa changes in he eens of Botrychium dissectum in ois Pennsylvania. Amer. Fern J. 80:173-182. Net, M. 1978. Estimation of rc ai ct and genetic distance from a small number of individuals. Genetics 89:583 Paris, C. A., F. S. WAGNER and W. a eke Jk. 1989. Cryptic species, — delimitation, and taxonomic Les in the PRO LNCR RE ferns. Amer. Fern J. 79:46— Ranker, T. A., S. K. FLoyp, M. D. WinpHaM and P. G. Trapp. 1994. Sane biogeography of ple adiontem-nigram (A le A acm in North America an gid mplications for speciation theory in homosporous pteridophytes. Amer. J. Bot. 81:776-78 SWARTZ & BRUNSFELD: MORPHOMETRIC ANALYSIS OF BOTRYCHIUM SPECIES 269 Rout_r, F. ? 1997. NTSYS-pc, Numerical Taxonomy and Multivariate Analysis System, ver. 1.8. Exeter Software, http://www. oe e.com SAS ae 6.11, SAS Institute Inc., ome ay, N Soxtis, D. E. and P. S. Sottis. 1986. E for inbreeding in the fern Botrychium virginianum (Ophioglossaceae). Amer. J. Bot. 73:588-592. Sressins, G. L. 1950. Variation and Evolution in Plants. Columbia University Press, New Work, Nu ¥: Stewart, C. N. Jr. and L. E. Via. 1993. A rapid CTAB DNA megs a useful for RAPD Tryon, R. M. and A. F. Tryon. 1982. Ophioglossaceae. Pp. 25-39, in Ferns and dillivad plate seth oii reference to tropical America. Springer-Verlag, New York. Wacner, D. H. 1992. Guide to the species of Botrychium in Oregon. United States Department of Agriculture oe Service Report. Leaf morphology of the Botrychium lunaria group in Washington and Oregon. United rte ‘Departmen of Agriculture Forest Service Report. WAGNER, F. S. 3. Chromosomes of North American ie ne and moonworts (Ophioglossa- ceae: Bouya) Cont. Univ. ieee ant Herb. Pg WAGNER, W. RB. Jr. 980. A probable new I t [ life li x simplex, from central seu Michigan Bot. 19 9 examples of the moonwort iar Botrychium matricariifolium xX simplex 4. o = Z ® . Ophioglossaceae. Adder’s-Tongue Family. Pp. 98-103, in J. C. ener ed. The Jepson manual eer plants of California. University of California Press, Berke He . 1989. Moonworts (Botrychium: a a in the iemiosvill area, Butte and Tehama counties, California. Madrono 36:131-— and L. P. Lorp. 1956. The morphological and aa distinctness of Botrychium minganense and B. lunaria in Michigan. Bull. Torrey Bot. Club. 83:261-280. _R. C. Moran and C. R. WertH. 1993. Aspleniaceae Newman. Spleenwort Family. Pp. 228— 229, in Flora of North America Editorial Committee, eds. Flora of North America north of Mexico, vol. 2. Oxford University Press, New York. and F. S. WacNer. 1981. New species of moonworts, Botrychium subg. Botrychium sis a emcee ee font North America. Amer. Fern J. 71:20—30 and . 1983. Two moonworts of the Rocky aeons Botrychium hesperium and a new ica = omer confused with it. Amer. Fern J. 7 and 8. Detecting Botrychium hybrids in the jee es elite region. Michigan Bot. 27:75—80. and . 1993. Ophioglossaceae C. Agardh Adder’s-tongue family. Pp. 85-105, in Flora of North America Editorial Committee, eds. Flora of North America north of Mexico, vol. 2. Oxford ane ig’ Press, New York. . BeITEL. 1985. Evidence for a hybridization in eo aon with shee mycoparasitic gametophytes. Proc. Roy. Soc. Edinburg 73- HaurF.er and J. K. pee 1984. A new brent ne od ea «. (Ophioslossacoae Botrychium). Can. J. Bot. 62:629-634 Watano, Y. and N. SaHAsHI. 1992. roar eareie inhreeding and its Sage consequences in a oer fern genus, Sceptridium —— Syst. Bot. 17:486—502. Wituams, G. K., A. R. Kuseuk, K. J. Livak, J. A. Raratskt and S. V. Tincey. 1990. DNA poly- morphisms amplified by arbitrary primers are useful as pete markers. Nuc. Acids Res. 18: 6531-6535. Z1KA, P. F. 1992. The results of a survey for rare Botrychium species (moonworts and grapeferns) July-September 1991 in the Wallowa-Whitman National Forest (draft). Oregon Natural ergs Program report to United States Department of Agriculture Forest Service, Portland. R. B B. N. Newuouse. 1995. Grapeferns and moonworts (Botrychium, zhioalcsoe) in the Columbia Basin. United States Department of Agriculture Forest Service Report. American Fern Journal 92(4):270—287 (2002) Reproductive Behavior of Cloned Gametophytes of Pteridium aquilinum (L.) Kuhn. Forses W. ROBERTSON 41 Braid Farm Road, Edinburgh EH10 6LE, U.K. ABsTRACT.—Spores from single fronds of three different taxa of Pteridium aquilinum (L.) Kuhn were collected at different sites in Scotland, England and Sri Lanka. Gametophytes developed from these spores were treated to produce arrays of genetically identical clones. Sporophyte formation was determined when such clones from th different tophyt erived from the same The breeding behavior of bracken gametophytes presents some unresolved problems. Thus, Wilkie (1956) produced experimental evidence for genetic incompatibility in bracken by recording the frequency of sporophyte formation in combinations of clones from different gametophytes derived from single fronds. The clones were produced by harvesting the prothalli which proliferated from the margins of sectioned gametophytes. The results, in clones prepared from three Scottish populations of bracken, could be plausibly reconciled with the occurrence of two mating types in each population. Since combinations of clones between populations were cross-compatible, it appeared that a single locus, multi-allele system was present so that each sporophyte would be heterozygous for dissimilar alleles. It was also noted that, although well defined, the apparent incompatibility was not absolute since a low, variable frequency of sporophytes occurred among putatively in- compatible combinations. On the other hand, Klekowski (1972) reported that single, isolated gametophytes of bracken from different localities display wide variation in sporophyte formation, ranging from zero to nearly 100%, with most samples from different sites exceeding 30%. Self-fertilization in such gametophytes is the rule and incompatibility is conspicuously absent. The differences in frequency of sporophyte formation per sample were attributed to differences in the frequency of recessive sporophytic lethals for which the parent plants were heterozygous. He also noted that Wilkie’s findings could be explained if the populations concerned carried balanced lethals, whereby each parent plant would be a double heterozygote for recessive sporophytic lethals at two loci linked in repulsion. Haploid gametophytes would carry one or the other of the ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 271 lethals and thereby present the appearance of two mating-types. If the lethality of either homozygous combination were incomplete i.e., the lethals were leaky, apparent cases of breakdown in the incompatibility system, inferred by Wilkie (1956), could be accounted for. Klekowski also indicated the need for further study of Scottish populations that might prove atypical, a suggestion that prompted the present study. he experiments described here were designed to discover whether the appearance of two ‘“‘mating-types’’, whatever their origin, could be detected in bracken populations from Scotland, England and Sri Lanka. It is particularly important to discover whether or not evidence from the British and Sri Lankan populations, geographically separated and belonging to different subspecies, leads to the same conclusions. MATERIAL AND METHODS SAMPLE SiTES.—Seventeen spore samples were obtained from single fronds collected at the sites indicated in Table 1. Ten Scottish spore collections were obtained from Clunie Dam (CD1 to CD5), Black Hill (BH3 and BH11), Temple (T1), Rubery Reservoir (R1) and Edgelaw Reservoir (E1). Spores were also obtained from one English site and two Sri Lankan sites: Farr’s Inn and Bambarakanda Falls. Five samples (SL1 to SL5) were collected at the former and one (SL6) at the latter site. Three taxa are included in these collections. Both Pteridium aquilinum (L.) Kuhn ssp. aquilinum and ssp. fulvum (Kuhn) Page & Mill. are included in the Clunie Dam samples. CD2, CD3, and CD4 belong to ssp. fulvum and CD1, and CD5 to ssp. aquilinum. The stand of ssp. fulvum is roughly triangular with sides of approximately 20 m (Page and Mill, 1994). Three fronds CD2, CD3, and CD4 were collected approximately 10 m apart from the west side of the stand. The ssp. aquilinum fronds, CD1 and CD5 were collected adjacent to, respectively, the north and south sides of the stand of ssp. fulvum, which is surrounded by sporophytes belonging to ssp. aquilinum. The Black Hill site refers to a roughly circular, isolated stand of ssp. aquilinum surrounded b Calluna moor. Two fronds were collected 60 m apart. The English population was from a scattered distribution of aquilinum. From the Sri Lankan populations of ssp. revolutum (Kuhn) Wu Zheng-yi & Raven, which is common in upland areas and the only subspecies in the island, five fronds were collected over a 15 m distance within a fairly continuous stand bordering one side of a road. The other site, (SL6) is several miles away from Farr’s Inn and at a lower elevation by some 1500 m. CuLTURE OF GAMETOPHYTES.—Spores were collected overnight by inverting fertile fronds on paper. The spores were washed three times by centrifugation in sterile water. Single drops of suspended spores were transferred by micro- pipette to petri dishes with 1% agar ( Sigma A7002) made up with Knop’s solution and sterile water (Wilkie, 1956). All cultures were kept at 20°C under a standard fluorescent strip light, except for occasional periods under daylight 272 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) TaBLE 1. Locations of the different single-frond spore collections from three sub-species of Pteridium aquilinum (L.) Kuhn. Nomenclature for British samples follows Page (1982) and Page and Mill, (1994), and for the Sri Lankan samples Wu Zheng-yi and Raven (1999). Map references for British samples refer to the U.K. Ordnance Survey, Landranger Series. Site Identification of spore collections Map reference ssp. aquilinum ie Dam CD1, CD5 GR NN 915 592 Black Hill BH3, BH1 GR NT 184 636 Rubery Reservoir R1 GR NT 311 571 Edgelaw Reservoir El GR NT 306 581 Temple 11 GR NT 312 583 Hutton-le-Hole a GR SE 700 890 ssp. fulvum Clunie Dam CD2, CD3, CD4 GR NN 915 592 ssp. revolutum Farr’s Inn Sit) SP2,/SU4. SLAUSES 80 E496 E 49 Bambarakanda Falls SL6 80 E516N 46 and ambient temperature, which applied equally to all cultures within a set of comparisons. The frequency of sporophyte formation by gametophytes was recorded under the following conditions. When thalli had grown to about 0.5 cm diameter, a random sample was transferred individually to small compart- ments, 2 cm square and 1.7 cm deep, in plastic boxes made up of 25 such compartments, each provided with sterile, washed sand moistened with Knop’s solution. The appearance of a sporophyte was attributed to self- fertilization. Pairs of such gametophytes, whose members were from different sites, were also kept under similar conditions. Any sporophytes which appeared in the latter comparisons may have arisen by selfing or crossing between gametophytes. A different kind of experiment was carried out with cloned gametophytes derived from single thalli. This entailed the combining of the cloned gametophytes in pairs, either according to a regular scheme described below, or randomly combined within or between different spore samples. To produce clones, young spore derived gametophytes were either treated for five minutes with 0.5 M KCl (Dyer, 1979) and then washed with water or they were cut into segments. Most, but not all, gametophytes treated either way and kept thereafter on Knop’s agar substrate produced many small thalli around the margins. These small thalli were removed and grown to produce arrays of genetically identical clones. For each sample of spores from a given collection, 25 randomly chosen gametophytes were used to produce clones. Either method of producing clones led to the same conclusions. Treated gametophytes differed in the rate of formation of daughter clones. When twenty or more clones became available, for at least nine or ten treated gametophytes, the clones were removed to set up the experiments described ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 273 Taste 2. Diallel combinations of cloned gametophytes. The numbers 2 and 4 refer to the potential maximum number of sporophytes for combinations of clones from, respectively, the same or a different gametophyte. Combinations between identical clones occur once, but twice between clones from different gametophytes. Clone numbers 6 — Co NP PB] w& ee ee a NP bP NPP PPP PIN ee oe) Se a co) POON OO FW DH eR ee o below. In the case of CD4, for reasons explained later, a particular test was repeated with sets of clones that had developed later. To compare the behavior of cloned gametophytes they were transferred in appropriate pairs, when about 0.5 cm in diameter, to individual compartments of the plastic boxes under the conditions noted above. After ten to fourteen days they were irrigated with aerated tap water. This was repeated at intervals until two to three weeks had elapsed without the further appearance of sporophytes, when the experiment was terminated. The presence of spor- ophytes was determined by inspection with a low-power binocular micro- scope. To avoid possible damage due to handling and to avoid the risk of contamination, the occurrence of archegonia and antheridia was not followed during these tests. As will be apparent later, that in no way detracts from the significance of the evidence but points to an obvious subject of future enquiry. ANALYTICAL PROCEDURE.—Genetic incompatibility has been claimed to account for the reproductive behavior of cloned gametophytes. An obvious way to check this is to set up a N X N diallel mating system whereby members of the arrays of clones derived from the same frond are combined in pairs in all possible ways, including combinations between sister clones. This results in the mating scheme illustrated in Table 2. For convenience it can be collapsed into the indicated triangular form. The lower diagonal position (diagonal slots) refers to the pairing of sister clones while all the other positions refer to combinations between the arrays (e.g., 1 X 2, 2 X 1). Since a gametophyte, cloned or otherwise, has the capacity to produce a single sporophyte, the maximum number of sporophytes expected in the diagonal slots is two whereas four are expected for all the other, duplicate combinations shown in the diagram. Interpretation of the reproductive behavior of clones depends on the nature of the departure from the numerical distribution shown in Table 2 when they are combined in such a diallel scheme. 274 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) It is first necessary to consider how the presence of a simple genetic incompatibility system would affect the distribution of sporophytes in a diallel test. In the case of two, equally frequent haploid mating types (+ and —) where only the heterozygote (+/—) will give rise to a sporophyte, the results of combining gametophytes from a heterozygous individual can be represented as: + = + - O 2] + or, more succinctlyas O 4] + 2 O! - ot = Extending the same sort of diagrammatic representation to the simplest hypothesis of diallel combinations in which the gametophyte clones in individual arrays are all (+) or all (—), the results can be ordered to display the characteristic pattern shown in Table 3a. Note there is a rectangular set of positions, representing the heterozygotes, with a maximum number of four sporophytes. All the other positions will fall into one of the two triangles that represent either the (++) or the (~—) homozygotes; these do not produce sporophytes. In a diallel test of this kind the practical task is to see whether the order of the paired gametophytes can be arranged to display the characteristic pattern of sporophyte production. Exactly the same kind of pattern can be generated if the parent individual is heterozygous for two recessive sporophytic lethal genes at two loci linked in repulsion, in which case only the double heterozygote will give rise to a sporophyte. This balanced lethal situation (Table 3b) leads to the same pattern of sporophyte production as the case of simple incompatibility, provided the linkage is complete. For either hypothesis the model assumes equal numbers of the alternative genotypes among the gametophytes which give rise to the clones used in any diallel test. In practice there will be chance variation about the 1:1 ratio and this will lead to corresponding departure from the precisely symmetrical pattern of the theoretical distribution illustrated in Tables 3a and 3b. Where it is necessary to test for departure from a 1:1 ratio the Chi-Square test has been used. RESULTS SPOROPHYTE PRODUCTION IN NON-CLONED GAMETOPHYTES.—Young gametophytes were removed from the agar plate at random and allowed to develop either on their own (Table 4) or in the proximity of another gametophyte derived from a different frond of the same or a different taxon (Table 5). In the first situation the comparisons include samples from ssp. aquilinum (CD1, CD5, E1, R1, and T1), from ssp. fulvum (CD2, and CD4) and from ssp. revolutum (SL1, SL3, SL4, SL5, and SL6). Among these isolated gametophytes, the frequency of sporophyte production ranged from 0.16 to 0.76, with an average of 0.43. These data are consistent with the variation reported by Klekowski (1972). In the second design, the combination of gametophytes from different sources resulted in a much higher frequency of sporophyte production. Almost all (0.95) of a total of 598 such combinations produced at least one and often two ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 275 TABLE 3. The potential maximum numbers of sporophytes produced by crossing, in all possible ways, clones from eight different gametophytes derived from a sporophyte heterozygous for either: a) (+) and (—) genotypes or b) balanced lethals linked in repulsion. It is assumed that alternative haploid genotypes are equally frequent in each situation. The diagonals refer to the single combinations of identical clones; all other combinations occur twice. a) gametophytes derived from a sporophyte heterozygous for (+) and (—) genotypes Clone numbers and genotype Af 2 3 4 5 6 + a + + - - - = 0 0 0 0 4+ 4 4 4 ab 1 0 0 0 a + 4 4 + Z ) 0 4 4 4 4 + 3 0 4 4 4 4 + 4 0 0 0 0 = 5 0 0 0 = 6 0 0 = f 0 = 8 b) gametophytes derived from a sporophyte heterozygous for balanced lethals linked in repul- sion Clone numbers and genotype 1 2 3 4 5 6 7 8 at! = is a a os = Lan 0 0 0 4 4 4 4 +1 1 0 0 0 4 4 4 4 +1 2 0 0 4 4 4 4 cad 3 0 4 4 4 4 +1 4 0 0 0 0 1+ 5 0 0 0 fe 6 0 0 1+ 7 0 1+ 8 sporophytes, whether the combinations were within or between taxa (Table 5 The lower production of sporophytes in isolates suggests a high incidence of recessive, sporophytic lethals at different loci. Given full penetrance of sporophytic lethals and heterozygosity in the source sporophytes for one, two, or three different, independently assorting lethals, frequencies of respectively 0.5, 0.25 and 0.125 are expected in random samples of isolated, selfed gametophytes. All the tests referred to in Table 4 can be reconciled with heterozygosity for either one or two lethals except T1 and SL5 in which the proportion of combinations with a sporophyte significantly exceeds 0.5 ( 01). However, for environmental or genetic reasons, not all lethals may be fully expressed when homozygous. Lethals may be “leaky” so that sporopyhte occurrence is higher than would otherwise be predicted, possibly so in T1 and SL5. The high frequency of sporophyte production with gametophytes from different sources (Table 5) is also consistent with the occurrence of recessive lethals at different loci. 276 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) TABLE 4. Sporophyte frequency in single, isolated, non-cloned gametophytes derived directly from spores. N refers to the number and + and 0 to those with or without a sporophyte. Origin N + 0 % fertile ssp. aquilinum CD1 50 14 36 0.28 CD5 25 15 10 0.60 E1 99 50 49 0.51 R1 100 55 45 0:55 Hig 99 66 33 0.67 ssp. fulvum CD2 50 12 38 0.24 CD4 26 16 9 0.64 ssp. revolutum SL1 25 rs 18 0.28 SL3 25 5 20 0.20 SL4 25 4 21 0.16 SL5 25 19 6 0.76 SL6 25 7 18 0.28 CLONED GAMETOPHYTES OF SssP. FULVUM.—Tables 6 and 7 show the diallel combinations for the Clunie Dam samples derived from the stand of fulvum (CD2, CD3, and CD4). The original gametophytes used to produce clones were arbitrarily assigned numbers e.g., CD1 to CD25 to identify the clones derived from a particular gametophyte. These are the axes in Tables 6 to 9 and 11 to 13. For each analysis, the combinations have been arranged to best compare the observed distribution of sporophytes with the patterns illustrated in 3a and 3b of Table 3. The distribution of sporophyte production in Table 6 for pair combinations can be explained by the presence of two genotypes for which the parent frond was heterozygous. When members of a gametophyte pair belong to the same genotype, sporophyte formation does not occur but does so when they differ in this respect. All sister clone pairs, derived from the same gametophyte (the diagonal slots), failed to produce sporophytes and the pattern of presence or absence of sporophytes corresponds to the pattern in both Tables 3a and 3b. Thus, for CD2 clone numbers 1, 4, 5, 6, 15, 2, and 9 did not produce a sporophyte or only rarely did when paired with a member of that set. Similar results are seen for the members of the other set, clone numbers 3, 10 and 14. However, when members of different sets were combined at least one and often three or four sporophytes were produced. Occasionally, one or two sporophytes occur where, according to either the incompatibility or balanced lethal model, none are predicted. The same pattern is encountered with CD3. The two categories or classes of clone include numbers 7,9, 20 and 21 on the one hand, and numbers 4, 1, 9,12, 16 and 17 on the other, with the same qualifications as noted for CD2. The diallel test with CD4 was carried out twice (Table 7). In the first test, performed at the same time as the tests with CD2 and CD3, all nine series of AMERICAN Pace FERN 2002 JOURNAL QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Editor R. James Hickey Botany Department, Miami University, Oxford, OH 45056 hickeyrj@muohio.edu Associate Editors Gerald J. Gastony, Department of Biology, Indiana University, Bloomington, IN 47405-6801 Christopher H. Haufler, Department of Botany, University of Kansas, Lawrence, KS 66045-2106 Robbin C. Moran, New York Botanical Garden, Bronx, NY 10458-5126 James H. Peck, Department of Biology, University of Arkansas—Little Rock, Little Rock, AR 72204 Table of Contents for Volume 92 (A list of articles arranged alphabetically by author) ADAMKEWICZ, Li. (Se6'@ i. sEPIOREY <6 5 ccs cus aia ieke Ge hk bie de RR dcr PE ee os nn had Sek ek i oe oR Res wR ee Bk bnew Romie, ed: fee a, A ee eh ce FG Se Ke Re Ao ww en bdednwe Cuiou, W.-L., D. R. FARRAR, & T. A. RANKER. The Mating Systems of Some Epiphytic Poly- PP ates aici te aie in oe a PP eae Gued GOW awed Bs Cg Bh COs I os eed 6 3b ae dw be eo ee na ks Cusick, A. W. A Binomial for the Hybrid Polypodium of Eastern North America........ EAE i TO, BUR ose hg i ee wd Rw oh da eo ie nad hwaeweees FARRAR, D. R. Obituary: Warren H. Wagner, Jr. (1920-2000) ...............-.0000. PROGR, Be Eee i) hse sw en Ke o a5 oe Pup aken ae vor PE GRA Gort igs, J. E. Lycopodium lagopus New in West Virginia .....................-.. GRANT J. R. (see W. H. WAGNER, Jr.) HILDEBRAND, T. J., C. H. HAUFLER, J. P. THERRIEN, C. WALTERS, & P. HAMMOND. A New ybrid Polypodium Provides Insights Concerning the Systematics of Polypodium scouleri and its Sympatric Congeners Houston, H. A. (see T. A. RANKER) JOHNSON-GROH, C., C. RIEDEL, L. SCHOESSLER, & K. SKOGEN. Belowground Distribution and bundance of Botrychium Gametophytes and Juvenile Sporophytes sommiscet-tsmon, C, (are A ©. SrenevOn ny). os kc ok wb ee dk hl Sok es veo ken KELLoFF, C. L., J. SkoG, L. ADAMKEWwICcz, & C. R. WERTH. Differentiation of Eastern North American Athyrium filix-femina Taxa: Evidence From Allozymes and Spores KNEPPER, D. A., D. M. JOHNSON, & L. J. MUSSELMAN. Marsilea mutica in Virginia LELLINGER, D. B. Bibliography of Warren Herbert Wagner, Jr. LELLINGER, D. B. Additions to the Fern Flora of Saba, Netherlands Antilles BRELTAGER, TP, TD, Ga TRAIN ok i 2 4 GWA hin 444.64 bla ok 6 da uA one een wk, MENDOZzA, A., B. PEREZ-GarciA, & R. RIBA. Comparative Research of Gametophytes of Olfersia alata and Olfersia cervina (Dryopteridaceae) MIcKEL, J. T. (see J. E. SkoG) Montcomery, J. D. Pteridophytes of Upper Katanga (Democratic Republic of Congo) .. . Morrow, A. C. & R. R. DuTe. Crystals Associated with the Intertracheid Pit Membrane of es Wet Pictnl PPCM HURT 6.6 se 60a Haw Wn 649 Wh OR OME RRS Ce AO MUSSELMAN, L. J. (see D. A. KNEPPER) PAaCcHECO, L. Trichomanes ribae (Hymenophyllaceae), a New Filmy Fern from Costa Rica and Panama PALMER, D. D. Taxonomic Notes on Hawaiian Pteridophytes....................0-. PERE GARCIA, .B (see A. NENDOAA) is ow. inca cua Jadhede oie. oe ew eins ae xin ee PRADO, J. & D. B. LELLINGER. Adiantum argutum, an Unrecognized Species of the A. lati- FOE PM at icunk nnn accuse bpadens CAO p oe eo eat be ones renee PrApo, J. & A. R. SmitH. Novelties in Pteridaceae from South America .............. RANKER, T. A. & H. A. Houston. 7 acu Sexuality in the Laboratory a Good Pre- mate te Dena ey i, SUE gous 55. veka cdc oda ee eee eeews oe eed wane or RANEER 1 A2(SEe Wists. OHIOU jiixal saca wie bd Oe ee ee es aha ee ae RST ER oi Sees ee VIL POA ano Riri es DID Fee ena Ally bats SOL OD ack Mine Te FS Es I 5 6 a 6 aw eG Pine PGW NEM a oie oa Hewisede bh des ROBERTSON, F. W. Reproductive Behavior of Cloned Gametophytes of Preridium aquilinum (L.) Kuhn SANCHEZ, C. A New Filmy Fern from the Dominican Republic ..................... SCHOESSLER: L. ((see' ©. JOHNSON*GROH) 14.2 oeo oo dd Ped Boos See beh hee icks DHBPRIEUESy Ei (hee- ds MAN) ahs ou Seas a oon al ei ae eee eee ote wos SIMAN, S. E. & E. SHEFFIELD. Polypodium vulgare Plants Sporulate Continuously in a Non- seasonal Glasshouse Environment .................000c ce eecccceeccuee Skoa, J. E., E. A. ZIMMER, & J. T. MICKEL. Additional Support for Two Subgenera of Anemia (Schizaeaceae) from Data for the Chloroplast Intergenic Spacer Region trnL-F and MOL PHOlOR yao a seins sects us aes lO is BAe eho GA hue ie PROG (See Ge Pew ROIs tn oad toting etl Materiel (Gl ott, tet ee a SROGEN, ic. (see © JOHNSON=GRON) oi ons ha tatiana Shag abo seh sae Situ, A. R. & R. B. CRANFILL. Intrafamilial Relationships of the Thelypteroid Ferns (Thel- PONE CAG CAE) octet ain hate wre Hoe Ad nln id vied, AME eds Gs Oe Wie he har aoa Sea Re (see 5, PRADO io:04 eh oee BER ee 8 hale wee be eaw eked baubce STENSVOLD, M. C., D. R. FARRAR, & C. JOHNSON-GROH. Two New Species of Moonworts (Botryvechian subs. Boirychiun) fom Alaska. sans 2 ec Sere sas CONES awe ee Swartz, L. M. & S. J. BRUNSFLED. The Morphological and Genetic Distinctness of Botrych- ium minganense and B. crenulatum as Assessed by Morphometric Analysis and RAPE ARES oad nsed eg es ero ota ea omnes ts creek eens taal ates eae TALBOT, S. L & S. S. TaLBor. A New Population of Aleutian Shield Fern (Polystichum aleu- own C..Christens:) on Adak Island, Alaska s20% occancatcucsedseeane tees PALBOE. Gc. (See Sr ls PAEBON) 6 4 cahe crys enabling Hodcln as. cee em AWE sense metas od TAYLOR, W. C. Isoétes 3 herb-wagneri, an Interspecific Hybrid of . bolanderi 3 I. echino- Cd | ee ee ee ee ee re. eee eee et TPHERRIEN IP sei 0 it DEH RAINES Win oa.c to feed trea as toe lee oe. hdteae "WAGNERY Ec ciy (S60 Eby CIR AN ore Gits 1) pie yrs ete as Sst ees Sa wk os Bets Oak WAGNER, JR., W. H., & J. R. GRANT. Botrychium alaskense, a New Moonwort from the In- MAE en Eo ei ks eg ae eo oo we 1k a a WAGNER, W. H. (see P. F. ZIKA) WALTERS, C. (see T. J. HILDEBRAND) WarkIns, JR., J. E. & D. R. FARRAR. A New Name for an Old Fern from North Alabama .... WERTH, C. R. (see C. L. KELLOFF) WILSON, K. A. Continued Pteridophyte Invasion of Hawaii ZIKA, P. F., E. R. ALVERSON, W. H. WAGNER, & F. S. WAGNER. Botrychium hesperium in the Wallowa Mountains of Oregon ZIMMER, A. (see J. E. SKOG) Volume 92, Number |, January—March, pages 1—38, issued 7 March 2002 Volume 92, Number 2, April—June, pages 39-184, issued 30 July 2002 Volume 92, Number 3, July-September, pages 185-246, issued 25 October 2002 Volume 92, Number 4, October-December, pages 247-304, issued 30 December 2002 2239 119 ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 277 TaBLE 5. Frequency of at least one sporophyte per pair (+) when non-cloned gametophytes from different locations are combined in pairs. In all the Clunie Dam (CD) pairs one gametophyte belongs to aquilinum and the other to fulvum. The Sri Lankan pairs were derived from putatively different individuals of revolutum, while in the remaining combinations (SL3 and BH11) one gametophyte belongs to aquilinum and the other to revolutum. N refers to the number of pairs. — Pair combinations N ~ ) % fertile ssps. aquilinum and fulvum CD1 and CD2 174 170 4 0.97 CD1 and CD3 100 85 15 0.85 CD5 and CD2 Z 25 a) 1.00 CD5 and CD4 124 122 2 0.98 ssps. revolutum and revolutum SL1 and SL3 25 25 0 1.00 SL1 and SL4 25 25 0 1.00 SL3 and SL5 25 22 3 0.88 SL3 and SL6 25 24 el 0.96 SL4 and SL6 25 22 3 0.88 ssps. aquilinum and revolutum 11 and SL 50 48 2 0.95 Total 598 568 30 0.95 clones failed to produce sporophytes or did so only very rarely. As noted above, although the model assumes two equally frequent classes among the gametophytes, derived from the single frond that gave rise to the clones, chance will cause variation about a 1:1 ratio in a random sample. However, it seems improbable, but not impossible, that nine gametophytes would belong to a single genotype. The test with CD4 was repeated with a second series of clones which had developed later. This second test is in accord with the data from the previous CD2 and CD3 tests, suggesting that the first test with CD4 was non- representative. In the second test, one genotype included clone numbers 10, 14, 23, and 25 while the other included clone numbers 24, 1, 8, 11, 13, and 15. To identify which of the two genotypes is the one represented in the first test, clone numbers 18 and 19 of the first test were combined with all but one of the different clones used in the second test. Table 8 indicates that all the clones used in the first test belong to the same genotype as clone numbers 10, 14, 23, and 25 of the second test. Although unlikely, these data support a departure from a 1:1 ratio of genotypes among the cloned gametophytes of the first test. It was noted earlier that CD2, CD3, and CD4 were derived from a single stand of ssp. fulvum. It is therefore of interest to ascertain the genetic comparability among them. Clones belonging to the alternative genotypes of respectively CD2 and CD3, CD2 and CD4, and CD3 and CD4 were combined in pairs. To compensate for the occasional shortage of replicates, clones not represented in the original test were included e.g., CD3 number 18 and CD4 number 3. One combination was lost due to algal infection. Table 9 indicates identity of the 278 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) TaBLE 6. Diallel tests with the Clunie Dam samples (CD2 and CD3) of cloned gametophytes of ssp. fulvum. The clone numbers have been arranged to reveal the pattern of combinations that do or do not produce sporophytes, as in Tables 3a and 3b. CD2 clone numbers zi 4 5 6 15 2 9 3 10 14 0 0 0 0 0 L 2 1 | 3 t 0 0 0 0 x | 1 a 4 4 4 0 0 0 0 0 ca 3 2 a 0 0 0 0 3 3 2 6 0 di 0 4 2 eS: 15 0 1 4 3: 4 2 0 4 3 4 9 0 0 0 3 0 3 10 0 14 CD3 clone numbers 7 19 20 aA | 4 1 9 12 16 17 0 0 Z Z ve 2 3 3 3 3 7 0 0 0 a a 2, 3 2 2 19 0 0 al Zz Zi 2 4 4 20 0 3 97 | a 4 2 ZA 0 0 0 0 0 0 4 0 0 0 0 0 ‘l 0 1 0 0 9 0 0 0 12 0 0 16 0 17 two genotypes in the three sets of cloned gametophytes. Pairings within genotype failed to produce sporophytes or did so only rarely so whereas pairing between genotypes often yielded the maximum of two sporophytes. By cross referencing, the total number of clones listed in the tests described in Tables 6 to 9 can be assigned to two genotypes comprising 24 and 16 samples respectively. This is not significantly different from a 1:1 ratio (y2 = 1.6, p> 0.1). This evidence makes it likely that the stand of ssp. fulvum, from which CD2, CD3, and CD4 were collected, constitutes a single individual. This conclusion is also consistent with the results of randomly combined, paired, cloned gametophytes either from the same or different fronds from the stand of ssp. fulvum (Table 10, Sections i and ii). Among these pairs within fronds there is a 1:1 ratio of combinations that produce sporophytes and those that fail to do so. The same is generally true for combinations of clones derived from different fronds, except for a statistically significant excess of pairs which produce sporophytes when clones belonging to CD2 and CD3 were combined (y* = 13.5, p < 0.01). This is in sharp contrast to the combinations of cloned gametophytes between taxa, Sections iii and iv of Table 10. In these combinations, there is a consistently high incidence of sporophyte formation and often the maximum is produced. ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 279 TABLE 7. The repeat diallel test with the Clunie Dam sample (CD4). CD4 i) 2 6 16 17 18 19 20 Pa | 22 0 0 1 0 0 0 0 0 1 2 0 0 0 0 0 0 1 0 6 0 0 0 0 0 0 0 16 u | 0 0 0 0 0 17 0 0 1 0 0 18 0 0 0 0 19 0 0 0 20 0 0 21 0 22 CD4 ii) 10 14 23 25 24 1 8 11 13 15 0 0 10) 0 4 3 3 3 2 3 10 0 0 0 3 + 3 3 3 4 14 0 1 2 4 4 3 3 e 23 0 4 3 4 2 4 1 25 0 0 Z 0 0 0 24 0 0 0 0 0 1 0 0 0 0 8 0 0 0 gta 0 0 13 0 15 CLONED GAMETOPHYTES OF SSP. AQUILINUM—These experiments involve two samples collected adjacent to and on opposite sides of the stand of ssp. fulvum. Pairing among clones of either CD1 or CD5 were carried out in the same manner as described for fulvum and the results are shown in Table 11. In the CD1 diallel combinations the situation is similar to that already seen in fulvum. The diagonal slots are all zero. It appears that clone numbers 1, 3, 7, 11, 14 and possibly 12 belong to one genotype and 2, 5, 8, and 9 belong to the other. In CD5, although there is again evidence for the presence of two genotypes, there are more exceptions to the predicted occurrence of sporophytes. Thus, three of the ten pairings of sister clones give rise to one or two sporophytes. Clone numbers 2, 5, 13, 17, and 10 probably belong to one genotype and clone numbers 11, 12, 20, 4 and 6 to the other. The two Black Hill samples (Table 12) also differ to some extent from each other. In BH3, clone numbers 1, 3 and 9 appear to belong to one category and numbers 4, 6, 12, 14, and 19 to the other; clone numbers 15 and 20 are exceptions. In their case sister clone pairing leads to the appearance of either one or two sporophytes in the diagonal slots. Also both numbers 15 and 20 produce sporophytes when paired with a member of either of the two sets: (1, 3, and 9) or (4, 6, 12, 14, and 19). In the balanced lethal model, this could occur if linkage is incomplete so that recombination in the parent sporophyte produces gametophytes which do not carry either of the sporophytic lethals. Alternatively, interactions with genes at other loci might be responsible for 280 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) + £ + CTV\A ABLE 8, g i identity 4 clone numbers 18 or 19 of the first test with the CD4 clones used in the second diallel test. Each combination was represented by a single pair so the maximum predicted number of sporophytes CD4 clones of second test 10 14 23 25 1 8 5G 13 15 CD4 clones of first test 0 0 0 0 7) 2 i z 2 18 0 0 0 0 Z i | 1 a | Zz 19 allowing sporophytes to develop when they would not otherwise do so. BH11 behaves like fulvum with two categories of cloned gametophytes that distinguish, respectively, clone numbers 11, 21, 25, 22 and 18 from clone numbers 24, 14 and 9. In the small York sample with five sets of clones there is again evidence of two categories, numbers 4, 5 and 16 and numbers 1 and 22. CLONED GAMETOPHYTES OF SSP. REVOLUTUM.—The diallel results for SL1, SL2, SL3, and SL4 are shown in Table 13. All the diagonal slots are empty, except for one exception in both SL2 and SL3, and there is the now familiar pattern of two alternative sets of clones. For SL1, one set or genotype includes clone numbers 3, 4, 5, 6, 8, and 10 and the other includes clone numbers 1 and 2. In SL3 the alternative sets comprise clone numbers 2, 3, 4, 6, 7, 8, 9, and 15, on the one hand, and clone numbers 1 and 5 on the other. In SL4 the alternative sets or genotypes are clone numbers 1, 3, and 4 and clone numbers 5, 2, 6, 7, 8, and 9. None of these distributions depart significantly departs from a 1:1 ratio. SL2 is inconsistent and illustrates the kind of exception referred to in CD5 and BH3. Thus, six of the clones fall into two categories, numbers 3, 4, and 5 and numbers 1, 7, and 9. Clone numbers 2 and 8 are exceptions. These clones produce sporophytes when combined with a sister clone or a clone belonging to either set (Table 13). DISCUSSION Three features of these experiments are of particular significance. Firstly, with very few exceptions, sister clone pairs fail to produce sporophytes. Pooling the data over all tests, only 12 sporophytes formed out of a total of 118 pairs of sister clones. Such infertility is not due to loss of potential to develop the hermaphroditic condition on the part of cloned gametophytes since the pairing of such clones derived from different spores collected at different sites or from different taxa regularly led to a high and often maximum rate of sporophyte production (Table 10). Secondly, the evidence from the diallel experiments indicates that clones derived from a single frond belong to one of two classes such that, sprorophyte formation depends generally on the joint presence of a cloned gametophyte from each class (Tables 6 to 9 and 11 to 13). It is assumed that the difference between classes is genetic. ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 281 TABLE 9. Tests of genetic identity among clones of the Clunie Dam samples of fulvum (CD2, CD3, and CD4). Cloned gametophytes are combined in pairs. One combination (1 X 16) was lost due to fungal infection. CD2 clones 14 10 3 15 9 6 5 4 1 2 0 0 1 1 2 1 af 2 u 1 1 zt 1 0 z Zz 4 Zz L 1 1 9 0 1 0 1 2 1 Hi uN 1 A | 12 0 0 0 1 P 2 A 1 - e 16 @) 0 0 2 2 1 al 1 2 1 18 CD3 clones 0 BI ut 1 Zz 7) 2 2 if Z 4 i | HF 0 0 < lt 0 0 0 Z aI Z at 0 0 0 0 0 0 el 19 aT 1 1 0 0 0 0 0 0 0 20 2 Z 2 0 0 0 0 0 0 0 21 CD2 clones 14 10 3 15 9 6 5 4 7 2 0 0 0 1 7 2 4 2 z y a 0 1 0 2 2 1 2 2 1 v4 6 0 0 0 a | 2 AE 1 2 y 2 16 0 0 0 1 2 2 1 1 2 1 tT? 0 ) L, 2 1 1 1 2 2 1 18 CD4 clones 0 0 0 Zz 1 g | z | 2 Z ? 19 0 | 0 | P4 i! : 3 a z 1 20 0 0 0 Z 1 1 Z 2 Z, a | 21 0 1 0 1 1 2 2 Z t 1 22 CD3 clones - Z 19 20 yt | yi 9 +2 16 17 1 0 1 1 1 0 0 0 0 0 2 2 2 if ‘s | 0 0 4 0 0 6 A 2 1 0 1 0 0 0 0 0 16 0 Z Z Z t 0 0 0 0 0 47 1 2 1 2 z 0 0 0 0 (8) 18 CD4 clones bi 1 0 2 1 0 0 0 0 0 19 0 1 Z 2 1 0 At 0 0 0 20 - Z 1 2 a 0 0 0 0 0 21 Ps 1 0 1 a‘. 0 0 0 0 0 22 Thirdly, when young, non-cloned gametophytes are isolated, self-fertilisa- tion regularly occurs and the variable lack of complete fertility may be ascribed to the incidence of sporophytic lethals for which the parent plant is het- erozygous (Table 4). Thus, cloned and non-cloned gametophytes appear to differ dramatically in their capacity for self fertilisation. It was this contrast that prompted Klekowski (1972) to invoke the hypothesis of balanced lethals and suggest the possibility of atypical behavior in the populations studied by Wilkie (1956). Since the behavior of the cloned gametophytes is essentially the same in samples belonging to different taxa, or derived from geographically 282 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) TABLE 10. Frequency of formation of at least one sporophyte in random combinations of pairs of cloned gametophytes derived from the same or different spore samples. Sporophyte production Samples N + 7 Frequency i. fulvum—Clunie Dam CD2 50 24 26 0.48 CD3 50 24 26 0.48 CD4 50 30 20 0.60 Total 150 78 72 0.52 ii. fulvum—Clunie Dam CD2 and CD3 50 38 12 0.76 CD2 and CD4 50 22 28 0.44 CD3 and CD4 49 28 21 0.56 Total 149 88 61 0.59 iii. aquilinum and fulvum—Clunie Dam CD5 and CD2 25 25 0 1.00 CD5 and CD4 25 24 1 0.96 CD1 and CD5 25 21 4 0.84 Total 79 69 5 0.92 iv. aquilinum and revolutum BH3 and SL4 50 48 2 0.96 remote sites, it is likely that it holds for the Pteridium complex generally under the conditions provided in these experiments. At first sight the appearance of multi-allelic incompatibility looks like the obvious interpretation of the results of the diallel tests and this interpretation certainly cannot be excluded, although it encounters the embarrassing evidence for self fertilisation on the part of isolated, single non-cloned gametophytes. If gametic incompatibility does account for the reproductive behavior of cloned gametophytes it appears necessary to infer that the process of cloning has altered physiology or development to uncover an incompatibility system which is not normally expressed in non-cloned gametophytes. However, it is necessary to enquire whether the apparent contradiction in the behavior of cloned and non-cloned gametophytes might be resolved within the framework of what is known about the reproductive behavior of gametophytes. Naf (1958) concluded that gametophytes form antheridia in response to antheridogen secreted into the medium by other, more rapidly growing gametophytes. If this external stimulus is absent, antheridia do not form although archegonia do. Hence, if a gametophyte develops from a single spore in isolation, it will be unable to undergo self-fertilisation. This appears to hold generally although exceptions may occur, especially after long periods of isolation. It may be assumed that a cloned gametophyte behaves the same way and similarly requires an external stimulus to produce antheridia. The minimum requirement is the presence of another gametophyte that can provide the ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 283 TABLE 11. Diallel tests on cloned gametophytes of Clunie Dam samples (CD1 and CDs) of aquilinum. One combination (1 X 7) was lost due to algal infection. CD1 clones A! 3 7 11 14 AZ Z 5 8 9 0 0 - 0 0 2 = a zZ 4 | 0 0 Z 0 0 4 2 3 2 2 0 0 0 2 1 4 1 4 is 0 0 2 4 4 1 1 a 0 0 3 2 Z z 14 0 2 2 Ps A 12 0 0 0 0 2 0 0 0 a 0 1 8 0 9 CD5 clones 2 5 13 ae 10 Tt 12 20 4 16 0 0 0 zt Zz 3 3 1 it 2 0 0 0 1 3 3 a) | a 5 0 0 2 2 4 a 3 - 13 2 di 3 4 Zz | 3 17 0 2 3 3 = 2 10 0 1 2 0 0 ii 1 | 0 2 12 1 0 Hf 20 0 Zz 4 0 16 stimulus, but not any gametophyte can provide it. The diallel tests suggest that the stimulus is generally present only when the two gametophytes concerned are genetically different. Since antheridogen is accepted as the agent which induces antheridia formation, one might wonder whether, within a species, it constitutes a single chemical entity or might occur in different forms which can be recognized by a gametophyte as different from the form it secretes, in which case only exposure to a different form would allow antheridia formation and hence the possibility of sporophyte formation. An alternative model could be envisaged in which the antheridogen is constant but other compounds take over the role just suggested, except that they would be responsible for determining whether or not a gametophyte responded positively to the presence of antheridogen. At least, this hypothesis may have the merit of resolving the discrepancy between the behavior of sister clones and isolated non-cloned gametophytes and removes the need to invoke balanced lethals. Although separated at an early age, it is likely that the latter have already been primed to produce antheridia via exposure to the chemical stimuli contributed by genetically different gametophytes on the agar plate. number of Wilkie’s (1956) results can be accommodated within this general scheme. Thus single, isolated gametophytes, derived from single spores by micro-manipulation, only rarely gave rise to a sporophyte. When 284 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) TABLE 12. Diallel tests with cloned gametophytes of aquilinum (BH3, BH11, and Y1). — i) = > eal © o oO Oo © ©; 0 NNwWhN CRN NM WwW] SF SOON NNNWTOD FP OONNN DN & ONRP OR WWD & Orooowwwwr bs BH11 = lo} bo > = n= co) eos Soo Ss eo eo o/* ee oo o> COoWWW PF Pw OR WWNHWS WE Coco WwWRN WWW! O i) ht Oo ooo OW FN] RB me NN & = ior) such gametophytes from the same frond were combined in pairs, only half of the combinations produced sporophytes, an observation which suggested a role for incompatibility, and which is also consistent with the diallel analyses and the results of randomly combining cloned gametophytes from different arrays derived from the same frond. The analysis of clone properties, in Wilkie’s case, followed a procedure that differed from that used in the present experiments. The capacity to form a sporophyte was ascertained for single cloned gametophytes, when combined with sperm suspensions prepared from alternative clones derived from the same frond. Two classes of clone were detected and sporophyte formation was attributed to the union of genetically different gametes. However, it seems not unlikely that the antheridia inducing agent(s) will be included in such sperm suspensions, thereby inducing antheridia formation that would not otherwise occur and making self- 285 ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES Diallel tests with cloned gametophytes of revolutum. TABLE 13. SL1 SL2 ce) SL3 SL3 <<} o oS 286 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) fertilization possible. Hence, the assumption of obligatory cross-fertilization is open to doubt. It would be informative to repeat the experiment, using also a fraction of the original sperm suspension from which sperm had been removed by filtration. To distinguish between self- and cross-fertilization requires genetic markers. It would be feasible to make such a distinction by comparing appropriate, variable DNA sequences in tissues of sporophytes and the gametophytes used in such experiments. There remains, however, a problem in Wilkie’s report of the production of sperm by cloned gametophytes. According to the hypothesis developed here that would not be expected. One can only speculate whether the process of producing sperm suspensions introduces special conditions not encountered in the present experiments. It was noted in the diallel tests, that there are occasional instances in which a cloned gametophyte can produce a sporophyte when combined with any clone of either of the two classes derived from a single frond. This suggests there are additional genetic or environmental factors that can influence antheridia formation or genetic incompatibility, if that is operating. It must also be asked whether the process of cloning might introduce reactions, apart from the speculative suggestion noted above, that would not occur in a normal, uncloned gametophyte. Dopp (1959) in his studies of antheridia and archegonia formation in Pteridium, noted that when thallus wings were sectioned, antheridia arose only behind a cut in the area furthest from the meristem which is responsible for a diffusible inhibitor of antheridia formation. It is conceivable that clones that develop from different regions of the original gametophyte may not be equivalent in their capacity to form antheridia and this could contribute to a reduction in sporophyte number below the maximum predicted in some instances. Such an effect would introduce a stochastic element that would tend to obscure underlying regularity in the recognition of two classes of clone per frond. If present, such an effect appears comparatively unimportant since the regularity is not obscured, whereas pairing of clones between taxa or sites, where it should equally apply, regularly results in high rates of sporophyte formation. One might also enquire whether the evidence from the experiments described here is relevant to the natural history of gametophytes in the wild. Little information exists on this score, although Voeller and Weinberg (1969) have drawn attention to the natural scarcity of both gametophytes and young plants. If antheridia formation depends on genetically determined stimuli, of the kind suggested in the interpretation of the diallel tests, then the behavior of gametophytes in the wild will be density dependent. If their frequency is low, as seems to be often the case, then such gametophytes will behave in the same way as gametophytes isolated from the spore stage: self-fertilisation will be unlikely, or at least infrequent. At intermediate density, fertility will largely depend on the joint presence of genetically different classes responsible for reciprocal induction of antheridia formation and this will tend to promote variability. One might also wonder how far a similar system might apply more widely in ferns. ROBERTSON: CLONED PTERIDIUM GAMETOPHYTES 287 Clearly many problems remain for future study, and the present experi- ments, with their limited objectives cannot settle the issues raised. In a different context, there is an unexpected by-product of the diallel analysis. Systematic combination of cloned gametophytes in such a scheme provides an empirical way of determining whether a particular stand of bracken is made up of more than one individual. For example, analysis of the samples collected from different locations within the Clunie Dam stand of ssp. fulvum (Table 9) suggested that it comprised a single individual which had spread to occupy a substantial area. ACKNOWLEDGMENTS Thanks are due to Drs. C. N. Page and A. Dyer for general information about ferns and especially to Dr. Page for bringing the Clunie Dam situation to my notice and for collecting the fronds from which the spores were obtained. I am indebted to Dr. Dyer for critical comment on the manuscript and to Professor John Thomson for stimulating discussion. LITERATURE CITED Dopp, W. 1959 . Uber eine hemmende und eine fordernde Substanz in den Prothallien von Dyer : : KiekowskI, E. J. 1972. Evidence rsa FY og self-incompatibility in the homosporous Fern Pteridium aquilinum. Evolution & —73. Nar, U. 1958. On the physiology of id formation in the bracken fern Pteridium aquilinum Pace, C. N. 1982. The Ferns of Britain and Ireland. Cambridge University Press, Cambridge. and R. MILL. 1994. Scottish Bracken (Pteridium): new taxa and a new combination. Bot. J. Scotland 47:139-140. VoOELLER, B. and E. S. WEINBERG. 1969. gegen, and physiological aspects of antheridia formation in ferns. Pp 77-93, in J. E nckel (ed.), Current Topics in Plant Science. ¥ as Wik, D. 1956. Incompatibility 4 in bracken. Heredity 10:247—256. Wu ZHENG-y1, W. and P. H. Raven. 1999. Flora of China. Missouri Botanical Garden Press. (St. Louis). American Fern Journal 92(4):288—293 (2002) A New Population of Aleutian Shield Fern (Polystichum aleuticum C. Christens.) on Adak Island, Alaska SANDRA LOOMAN TALBOT Alaska Science Center, U.S. Geological Survey, 1011 East Tudor Road, Anchorage, AK 99503 STEPHEN S. TALBOT U.S. Fish and Wildlife Service, 1011 East Tudor Road, Anchorage, AK 99503 TRACT.—We report and describe a new population of the endangered Aleutian shield fern (Polystichum aleuticum C. Christens.) discovered on Mount Reed, Adak Island, Alaska. The new because it increases the total number of known populations and individuals for the species. The Aleutian shield fern, Polystichum aleuticum C. Christens., is one of the most restricted and rare ferns in North America (Smith, 1985); it is listed as an endangered species (U.S. Department of Interior, 1988). In a circumpolar assessment of the rare vascular plants of the Arctic, the fern was classified according to the IUCN Red List threat categories as ‘‘endangered”’ (Talbot et al., 1999). The species was first collected by W. J. Eyerdam, an assistant to Eric Hultén, in 1932 on Atka Island, one of the Andreonof Islands located in the center of the Aleutian Island chain, Alaska (Fig. 1). Eyerdam’s holotype specimen (W. J. Eyverdam 1086, 5 July 1932) is accessioned at S; isotypes are accessioned at CAS, DS, and US. The species was first described by Christensen (1938). Attempts to relocate the original collection site on Atka in years subsequent to the species’ description were unsuccessful (Smith and Davison, 1988). Then, in 1975, a population of the species was discovered, this time on the northeast arm of Mt. Reed, on Adak Island (Fig. 1), also of the Andreonof Islands of the Aleutian Island chain, Alaska (Smith, 1985). Several subsequent searches, performed from 1986 to 1988, on various islands of the Aleutian Island chain (Adak, Kagalaska, Atka and Attu; see Talbot et al., 1995 for review) failed to locate additional populations. Finally, in 1988 and 1993, we discovered a second and third population, respectively, again on Mt. Reed (Talbot et al., 1995). Three populations comprising approximately 117 individual fern clumps are known for the species (Tande, 1989; Talbot et al., 1995). In August 1999, genetic studies of Polystichum aleuticum were initiated to assess the species’ relationship to P. lachenense (Hook.) Bedd. of Asia, as recommended in Talbot et al. (1995). While collecting samples as part of this genetic research, as well as ongoing systematic monitoring of the three known populations (Anderson, 1992), we discovered a fourth population of P. aleuticum, located approximately 142 m below the two populations found on TALBOT & TALBOT: POLYSTICHUM ALEUTICUM IN ALASKA 289 176°45°W 176°. oe ; 2 | 9 %, , b NRE @ Qs SW | P a) 9 ey O 4 Co 2 : OY ? g pS Lake 2g al : @. Am — Moun Resd Sering Sea 5 fea < Be a Ada Rp D 0 ~ QQ Island } “isicchaie 51°48'N i 176°45'W Polystichum aleuticum populations 0) Southern site on northeast arm of Mt. Reed afittu Unimak ce Aleutian Islands s @) Northern site on northeast arm of Mt. Reed - N ® Site on northwest arm of Mt. Reed Adak -* Va \@ "Reps —wRe * —— @ se aca ) It. Reed 1. Location of the four known populations of the Aleutian shield fern, Polystichum oa. on Mt. Reed, Adak Island, Aleutian Islands, Alaska. the northeast arm of Mt. Reed, Adak Island, at an elevation of 338 m. As was the case for the other three sites found on Mt. Reed, the fourth site was at the base of a steep rock outcrop on a northeast-facing slope. The slope angles at this site ranged from 60° to 90°. Notably, the site is located approximately 22 m below the lowest of the three previous populations (Population 3, Fig. 1, Table 1), in an area considered too treacherous to survey during earlier efforts due to steep, unstable, slippery slopes. This finding expands the elevational range of P. aleuticum, placing the species at elevations between 338 m and 525 m (Table 1). Also, notably, the fourth site is located on a northeast-facing slope. Climatologic records based on observations from 1950 to 1982 indicate wind direction on Adak is predominantly from the west-southwest, averag- ing 10.5 knots per hour, from June to November, the period of time during which the habitat would likely be free of snow (U.S. Department of the Navy, 1989). Thus, the occurrence of all known populations on northeast-facing TABLE 1. Population characteristics of Polystichum aleuticum and associated geographical variables. Latitude/ Island Location Pop. # Longitude Elevation (m) Aspect # Individuals Year Relocated Atka unknown, “within v unknown unknown unknown unknown 1932! N ie the village e Atha” very rare Adak Reed, NE A uh 51° 49.640' N 475.5—-525.8 NE 98 1975” xd 176° 41.861’ W Adak Mt. Reed, NE Arm 2 5 9.491’ N 457.2-469.4 NE 14 1988° ed 176° 41.776’ W Adak Mt. Reed, NW Arm 3 51° 49.960’ N 360.4 NE 5 1993° Y. 176° 44.141’ W Adak Mt. Reed, NE Arm 4 51° 49.378' N 338.0 NE 14+ 19994 Y 176° 41.733’ W 1 Christensen (1938). * Smith (1985) * Talbot et al. (1995). * Present study. 062 (Z00z) * YAAWON 26 AWN'TIOA “TVNYNOL NYA NVOMANVY TALBOT & TALBOT: POLYSTICHUM ALEUTICUM IN ALASKA 291 slopes suggests habitats supporting populations of P. aleuticum are those of- fering protection from high winds during critical growth and reproductive periods. At least 14 clumps constituted the new population; however, due to the inaccessibility of some of the vertical rock faces, a complete count of individual clumps was not possible. This number therefore underestimates the size of the population. Among the 14 clumps counted, five are associated with rock grottos; the remaining clumps are found on ledges on steep rock faces. Unlike the other locales, no clumps were associated with herb meadows. Clumps in the grottos and ledges comprise from five to 30 fronds, with all clumps containing fronds with sori. Clumps associated with steep rock faces comprise 12—40 fronds, again with all clumps examined containing fronds with sori. However, for safety reasons, not all clumps on vertical rock faces were examined to determine the number of fronds or presence of sori. The presence of sori on some P. aleuticum fronds suggests this population may be useful if spores were used for controlled propagation. We recorded vascular plants associated with the new population; nomen- clature follows USDA, NRCS (2001). Vascular plants associated with grotto and ledge clumps were the creeping dwarf shrub Salix rotundifolia Trautv.; the forbs Achillea millefolium L. var. borealis (Bong.) Farw., Anemone narcissi- flora L., Arnica unalaschcensis Less., Campanula lasiocarpa Cham., Conio- selinum gmelinii (Cham. & Schlecht.) Steud., Huperzia chinensis (Christ) Czern., Lycopodium alpinum L., Pedicularis verticillata L., Platanthera sp. L. C. Rich., Polygonum viviparum L., Polystichum lonchitis (L.) Roth, Potentilla villosa Pallas ex Pursh, and Valeriana acutiloba Rydb.; and the graminoids Carex macrochaeta C. A. Mey., Carex circinata C. A. Mey., Poa sp. (L.) Roth (viviparous) and Tofieldia coccinea Richards. Vascular plants associated with vertical rock faces included the forbs Achillea millefolum var. borealis, Epilobium hornemannii Reichenb., and Viola langsdorfii Fisch. ex Gingins; and the graminoids Carex circinata, Poa sp. (viviparous) and Tofieldia coccinea. Comparison of this list of associates from the other sites (Lipkin, 1985; Talbot et al., 1995) indicates a high degree of similarity in species composition among the four known populations. Using a hand-held Global Positioning System, we recorded precise geo- graphic coordinates for this population as 51° 49.578’ N, 176° 41.733’ W. The discovery of this new population increases the number of known individuals to 131. We note here that this population was discovered accidentally while we were disoriented in the fog on Adak, and the area would, under normal circumstances, have been considered too treacherous to survey. It is very possible that additional populations of P. aleuticum inhabit Mt. Reed or nearby mountains, in areas too dangerous to survey without risk of injury, and we suggest the number of individuals representing P. aleuticum on Adak is likely underestimated, despite extensive surveys undertaken throughout the island from the mid-1980s to the mid-1990s. This discovery is significant because it increases the number of known individuals by approximately 12%, and the number of known populations 292 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) from three to four, thus providing an increased buffer against loss of either individuals or populations of this rare and endangered species. It also expands the known elevational range within which future searches for new populations should be targeted. Despite this new finding on Adak, however, P. aleuticum continues to be one of the rarest ferns in North America. While conducting botanical studies on other Aleutian islands between 1985 to 2001, the second author searched rocky outcrops similar to those on Adak that support the four known populations, without finding any new populations. These islands include: Adugak, Aiktak, Amlia, Buldir, Chagulak, Davidof, Kasatochi, Khvostof, Kiska, Nizki, and Uliaga islands. Additional surveys by both authors of Attu Island in 2000 and Simeonof Island of the Shumagin Island group, eastern Aleutians, in 1997 also failed to yield discoveries of new populations. Thus, biologists have searched for P. aleuticum on 16 islands in the Aleutian chain during the past fifteen years, and have found the fern only on the northern slopes of Mt. Reed, Adak Island. ACKNOWLEDGMENTS Financial support was provided by the Western Area Ecological Services (WAES) and Department of Refuges, Region 7, U.S. Fish and Wildlife Service, and the Alaska Science Center, U.S. Geological Survey, Biological Resources Division. We thank Art Davenport of WAES and the staff of the Aleutian Islands Unit, Alaska Maritime NWR, Adak Island, particularly Jenna Mueller and Jeff Williams, for field support. Terri Morganson (Division of Realty, U.S. Fish and Wildlife Service) prepared Figure 1. We thank Judy R. Gust, Dirk V. Derksen, R. James Hickey, and an anonymous reviewer for valuable comments on the manuscript. LITERATURE CITED AnpeERSON, B. L. 1992. Aleutian shield fern (Polystichum aleuticum C. Chr. in Hultén) recovery lan. Unpublished Report. U.S. Fish and Wildlife Service, Anchorage, Alaska. CHRISTENSEN, C. 1938. Polystichum aleuticum C. Chr., a new North American species. Amer. Fern J. 28(3):111-112. LipKIN, R. 1985. Status Report on Polystichum aleuticum C. Chr. Unpublished report, U.S. Fish and Wildlife Service, Office of Endangered Species, Anchorage, Alaska. SmitH, D. K. 1985. Polystichum aleuticum from Adak Island, a second locality for the species. Amer. Fern J. 75(2):72. Davison, P. G. 1988. Polystichum aleuticum C. Chr. in Hultén. Site survey of Atka Island, Alaska, 1988. Unpublished field report, U.S. Fish and Wildlife Service, Fish and Wildlife Enhancement, Anchorage, Alaska. Ta.sor, S. S., Tabor, S. L., and ScHoFIELD, W. B. 1995. Contribution toward an understanding of Polystichum aleuticum C. Chr. on Adak Island, Alaska. Amer. Fern J. 85(3):83-88 » YURTSEV, B. A., Murray, D. F., Arcus, G. A., Bay, C., and E.tvepaxk, A. 1999. Atlas of rare endemic vascular plants of the Arctic. Conservation of Arctic Flora and Fauna (CAFF) Technical Report No. 3. U.S. Fish and Wildlife Service, Anchorage, Alaska. TaNnbDe, G. F. 1989. Aleutian shield-fern (Polystichum aleuticum C. Chr.). Field studies for 1989: establishment of permanent population monitoring plots and habitat characterization. Unpublished report, U.S. Fish and Wildlife Service, Office of Ecological Services, Anchorage, Alaska. TALBOT & TALBOT: POLYSTICHUM ALEUTICUM IN ALASKA 293 USDA, NRCS. 2001. The PLANTS Database, lean 3:1 — //plants.usda.gov). National Plant Data Center, Baton Rouge, nena 70874-4490, U U.S. DEPARTMENT OF INTERIOR. 1988. termination of Final Rule. Federal Register Borat an 626-4630. U.S. DeparTMENT Or THE Navy, Nava OcEANOGRAPHY COMMAND DETACHMENT. 1989 . Summary of meteorological prone aay nee 1949-1987 Adak Naval Air Station. U. S. Navy, Adak, Alaska. gered status for Polystichum aleuticum. American Fern Journal 92(4):294—295 (2002) SHORTER NOTES Trichomanes ribae (Hymenophyllaceae), a New Filmy Fern from Costa Rica and Panama.—The Hymenophyllaceae in Costa Rica and Panama are well known due to the works of Lellinger (Pteridologia 2: 185-228. 1989) and Pacheco (in G. Davidse, M. Sousa S., and S. Knapp, eds. Flora Mesoamericana. vol. 1. Psilotaceae a Salviniaceae. Univ. Nacional Aut6noma de México, México, D. F. Pp. 62-83. 1995). However, as a result of additional studies during a recent trip to Costa Rica, a new species has been identified. Trichomanes (Trichomanes) ribae Pacheco, sp. nov.—TYPE. Panama: Panama, 5-10 km NE of Altos de Pacora, on trail at end road, 750 m, 7 Mar 1975, S. Mori & J. Kallunki 4964, (holotype, MO). (Fig. 1) Rhizoma repens, 0.1 cm diametro, trichomatibus catenatis; folia remota, 4.8— 11.5 X 2.7—3.8 cm; petioli 0.08—0.5 x 0.05 cm, trichomatibus catenatis; laminae 4.7-11 cm lanceolatae, 2-pinnatifidae, apice pinnatifidae, basi subtruncatae; rachis alata; pinnae sche can once imbricatae; sori 1-4 per pinnam, in- volucris in lobis immersi, 0.2—0.2 .2cm, exserto. Rhizome long creeping, 0.1 cm in diameter with catenate trichomes; leaves distant, 4.8-11.5 < 2.7—3.8 cm; petioles 0.08—0.5 x 0.05 cm, nonalate, loosely and deciduously clothed with brown catenate trichomes, similar trichomes on rachis and veins; blade lanceolate, 4.7—-8.3 cm, 2-pinnatifid, chartaceous, apex pinnatifid, base subtruncate; rhachis alate, wings 0.07—0.08 cm wide on either side; pinnae oblong-lanceolate, 11-18 on a side below the apex, 1.2—1.6 X 0.7, overlapping at in right angles to rhachis, their apices pinnatifid; segments oblong, 0.15-0.18 x 0.1—0.18 cm, apex obtuse to bifid, plane, margins complete; venation open, anadromous, pinnate, veins 2-furcate, not reaching the apices of the lobes; lamina cells almost isodiametric, translucent; sori lateral on the pinnae, 1—4 per pinna, involucres immersed, 0.2-0.25 < 0.2 cm campanulate, apex wide-flaring; receptacle exserted. PARATYPE.—COSTA RICA: Limon; Siquirres, Las Brisas de margen izquierda de Quebrada Jesus, afluente innominado, Camino a Cerro Tigre. 09° 56’ 40” N; 83° 25’ 15” W, 800 m, 22 Mar. 1996, G. Herrera 8849 & G. Valverde (CR). Trichomanes ribae belongs to subgenus Trichomanes as evidenced by the anadromous venation, sori lateral on the pinnae, distant leaves, and 2- pinnatifid lamina. Its nearest relative is T. rupestre (Raddi) Bosch from which it differs by shorter leaves, 1-4 sori per pinna, and campanulate immersed involucres with wide-flaring apex. This species is always epiphytic while T. rupestre is epipetric or terrestrial, but not epiphytic. This species is dedica- ted to Ramon Riba y Nava Esparza. This work was supported by the Instituto Nacional de Biodiversidad, Costa Rica, Nelson Zamora provided financial and logistic help. I thank Rolando RANKER ET AL.: GAMETHOPHYTE SEXUALITY & AG, 4 5d Spr N GE SiH A si ? Sieh NY ie, Ms es eT si “s SOUR 2 = (es SF So DSSS Aig @, Ve 5 oi 5) Shee TEAR ig Ms j ; Wee USE SENS =. ‘a r,. F Tt PEGA AAG Mec OE BCI OE NES Sis BS yk ee he J) de cS & iD oh z WE Sout AY SW yas a $Y B27] le 1g Ys fs \ Y N le \ 5 } ‘ ¥ Ny 5, i SYP a} ie 2 XS ys i & ») Sh we Wests Se BP BSS SEL. gic | ~ & Was = PO a oe - fF K 4 . es "9 i 3 J i : RZ ME ; PE Ponce M3 is ND ut a H od Qe oh ee Pma.y®2 SIRF PRN "hee eee oy ge Fic. 1 Trichomanes ribae. a) habit; b) general view of leaf; c) pinna with campanulate immersed involucres with wide-flaring apices. Trichomanes rupestre. d) pinna with involucres. Jiménez for the drawings that illustrate this new species. I am also grateful to Fernando Chiang for the Latin translation of the species diagnosis.—Leticia Pacueco. Departamento de Biologia-Botanica Estructural y Sistematica Vegetal, Universidad Autonoma Metropolitana-Iztapalapa, Apdo. Postal 55-535, 09340 México D. F., México. American Fern Journal 92(4):296 (2002) Referees for 2002 All papers submitted to the journal are peer reviewed. Members of the editorial board and the continued success. The American Fern Society and I extend our thanks to the following reviewers for their assistance, diligence, and patience in the year 2002. Special thanks are given to Dr. David B. Lellinger for assuming the task of editing the Wagner Memorial Issue—R. James Piakiey. DaviD BARRINGTON Davin M. JOHNSON THOMAS RANKER ALD R. F PAULO LaBIAK CHRISTINA ROLLERI Davip FRANCKO Joun T. MICKEL Jupy Ski ERALD GASTONY NICHOLAS MONEY Y SMALL GARY ER JAMES D. MONTGOMERY ALAN R. SMITH CHRISTOPHER HAUFLER Rossin C. Moran W. Cari TAYLOR Kerry, HEAFNER AMES PECK, FLORENCE S. WAGNER TerRRY A. HILDEBRAND KATHLEEN PRYOR KENNETH A. WILSON ALFREDO HUERTA R. RABELER GEORGE YATSKIEVYCH Index to Volume 92 3,8-C-arabinosylluteolin, a New Flavonoid om Pteris hee A Binomial for the a Polypodium of 0 astern iertt America, 24 A new Hybrid Polypodiu rovides Insights ncerning the Systematics of Polypodi- m scouleri and its Sympatric Congeners, A New Population of Aleutian Shield F (Polystichum aleuticum Chri sae on Asak ot Alaska, 8 Acystopteris 138, 139; japonica, 136, 143 ADAMKEWICZ, L. (see C FF) Adenophorus pinnatifidus, 97; pinnatifidus var. pinnaifidus, 97; pinnatifidus var. rockii, 97; sarmentosus var. rockii, 97 Adiantumi, 23; argut 23-25, 27; diogoano, ‘Edwinii’, 180; 7; tenerum, 180; viviesii, 23 Adams 237, 257 ave, 17 yoncanl aaurantiaca, 239 Alnus shale ‘ees viridis subsp. sinuata, 156 Alsophila, 2 ALVERSON, E aba e P. F, Zika) ie ioe 142, 144-146, oligocar- , 14 Apel fe pense, as rapa 138 : sebicennié 128; 128; semihirsuta, 121, 123, 126; subg Ane 21. : ; subg. Anemior- rhiza, 119- , 124, 125, 127-129; subg nderwoodiana, 121, 123 it 21. eon 124, 126: are 123, 126 Anemone drummondii var. lithophila, 156; narcissifl 139, 143, 145, 224; bipinnatum, 98; pei 136, 143; enatum, 98; enatum mmiparum, 99; fragile, 99; fragile var. insulare, 99; Hate 99; septentrionale, ig tricho- S, 42, 45, 251; triphyllum, 9 Penny major, Athyrium 8 01, 139; angustum, 185-213 naked 185-213; filix-femina, 113, rum, i cen ssp. sitchense, 186; Chinn ae Azolla filiculoides, ee 180 BARRINGTON, D. S. (see G. J. Gastony) erberis pinnata, 21 Betula alleghaniensis, 87; neoalaskana, 169 Blechnaceae, 144 Blechnum, 144, 179; appendiculatum, 179, 180; gl meprer 179; occidentale, 179; spi- cant, 116, 117 ain tricophyllum 246 Botrychium, 10, 18, 41, 43, 81-83, 85-91, 150, 153-155, 157, 164, 169; alaskense, 164— 164; campestre, 83-90; crenulatum, 151, 249-263, 265, 267; dissectum, 11, 14; 250; mormo, 83, 90; Hh 08 um, 25 ? + pemadasara, 250 pet bg. Scepteri- » 152, 150-158; virgin- ianum, 11, Th. 83-88, 90, 164; Xwatertonense, 251; ya — 157, 83, 85-87, 150, 151 153- 15 Botryychium hesperium in the . Moun- tains of oe egon, 239 Broussaisia, BRUNSFELD, S. : ck. Swartz) Calamagrostis nutkaensis, 216; rubescens, 239 Campanula lasiocarpa, isin sce angus Nip 67, 68-70, 73- phyllitidis, 67-70, 72-76 Caneaea , 124 iomanes, 124, 125, 127; reniforme, 120, 121, 124, 126 Carex ae EO aT 46; concinnoies, 239; i, 239; macrochaeta, 291 80 Cheilanthes, 224; bradei, 110; castanea, 242: is Sap 291; eatonii, 242; venusta, 110; 8 Ghokeaenen 4 Chingia, 139, ng oe 134, 143 Cuiou, W-L., D. FARRAR, and T. A. RANKER. The eer Systems of Some Epiphytic olypodiceae. Choral pea ee var. divaricatum, Sesame 100, 139, 141, 144, 145; acuminata, 138; arida, 134, 143; augescens, 134, 138, 141, 143; boydiae, 100; cyatheoides, 100; AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) da 100, hispidula, 134, 143; Xincesta, 99; x apr 99, 97; para- sie 100; wailele, 100, Cibotium, 101 Peco ch sinensis, 245; junos, 245 Comparative Research at Gametophytes of Olfersia alata - Olfersia cervina (Dry- Bhs gape e), 2 NT, D. S. (see G. } er Coryphoptrs, an, 144-146; seemannii, 13 B. (see A. R. Smiru) Cupressus macrocarpa, 216 A. W CUusIcK, Psedns for the Hybrid De rth America, 240 Cyanea, 101 Cyathea arborea, 94; cooperi, 180; muricata, 94 14 9, 141, 142, 145; Xincestus, es Kiaclovsardlings 99, 100; interruptus, ule, Cyrtomium, 229, 234, 236, 237; falcatum, 180 Costoptrs 138, 139, 143, 145, 224; protrusa, , 143 Deparia lancea, 136; 143; petersenii, 180 Dictyocline, 138, 139, 141, 142; griffithii, 134, Didymochlaena, 138, 139; truncatula, 136, 143, : , 236, Differentiation of Eastern North American pers filis-femina xed Evidence neg and Spores, 1 Biersidonns rao ae digitatum, 242; habereri, 2; tristachyum, Diplazium, 101, 139: cristatum, 95; esculentum, nal Doodia Dryopteridaceae, 142 sora, 100, alboviridis, 100, 97, ; glabra var nii, 101, 97, 102; glabra var. hobdya- na, 101, 102; rg 101; 102; x leedsii, 161; podosora, Dute, R. R. (see A. C. Morr ek INDEX TO VOLUME 92 Eicer, J. E. Lycopodium lagopus New in West irginia, 241 Elaphoglossum, 124 lymus glaucus, 239; trachycaulus subsp. vio- ceus, 15 Epilobium hornemannii, 289 Equisetum, 43 Erigeron caespitosus, 156; compostus, 239; Jaucus, 21 Eucalyptus globulus, 216 Farrar, D. R. Obituary: Warren H. Wagner, Jr. RRAR, D. see W-L Cuiou) Festuca déeideablas 239; rubra, 154, 159 agro chiloensis, 154, 159; virginiana, 169, Galium boreale, ro Gastony, G. J. S. BARRINGTON and D CONANT. ital Rolla Milton Tryon, Jr. 1916-2001), = lis =] Soa 5 25 EA 4 Goniopteris, 138, 139, 142, 144; poiteana, 134, 3 Grammitis, 46 Grant, J. R. (see W. H. WAGNER, JR.) ait — 139, 143; ovamense, 137 Hammonb, P. (see T. J. HILDEBRAND) Haplodictyum, 132 HAUuFLeR, C. “| (see T. J. HILDEBRAND) Heliconia, eracleum arn Hickey, R. = Review: Ms Flora of Illinois. Ferns. 2nd ed., ces abiloram, a ILDEBRAND, T. J., C. H. HAUuFLER, J. P. ne RRIEN, Watters, and P. HamMonp. A New Hybrid Polypodium Prariiee Insights Concerning the Systematics of Polypodi- um scouleri and its Sympatric Congeners, 214 malosorus, 1 137; pyenocarpos, 137 2 Huper : Hydrilla a 243 Hydrocotyle verticillata, 243 e , 128 coides, 20; integri- a ta seo 20, 21; oe. a ge 139, 146 crenatum, anal hawaiiensis, 102; hawaiiensis var. mauiensis, 102, 97 Illustrated Flora of Illinois. Ferns. 2nd ed. Re 24 evie IMPERATO, Fr. 3,8-C-arabinosylluteolin, a New Flavonoid from Pteris vittata, 244 Intsia, 10 Isoétes, 163; bolanderi X echinospora, 163; olanderi, 161-163; echinospora, 162- 63; Xherb-wagneri, 162-163, 161 Johnson, D. M. (see D. A. Knepper) Johnson-Groh, C. (see M. C. Stensvold) Johnson-Groh, C., C. Riedel, L. Schoessler, and K. Skogen. Belowground Distribution and Abundance of Botrychium Gametophytes and Juvenile Sporophytes, 80 Kalmia lIatifolia, 242 KELLoFF, C. L., J. Skoc, L. ADAMKEWiIcz and C. R. Wert. Differentiation of Eastern North American Athyrium filis-femina Taxa: sage from Allozymes and Spores, 185 Knepper, D. OHNSON and L, aeetoee Mestion mutica in Virginia, Larix laricina, 241 Lastreopsis, 131 ectotypification: Adiantum argutum, 23 Kea D. B. Additions to the a Flora of a, Netherlands Antilles, 9 ing D. B. tia a of =a Herbert ie D. 'B, ak Prado Le oe 131, 132, 138, 139, 141; pilosa, tottoides, 134, 143 nia ce 125; Pa ls 31 Lorinseria, 139, 142, 144, 145; areolata, 137 upinus nootkatensis, 154, 159 Lycopodiella, 241 Lycop diiu [pi 289; clavatum, 241, 242: clavatum var. monostachyon, 241; den- droideum, 242; =— 242; lagopus, 241-242; obscurum, 2 300 Lycpodium lagopus New in West Virginia, 241 Lygodium, 119, 120, 124, 125, 127; flexuosum, 121, 126; japonicum, 179, 180, 182; micro- phyllum, 121, 126 Macrothelypteris, pe 139, 141, 144-146; tor- resiana, 134, 143, 145; 180 Marsilea Pete a 0-182; —— 181, 182; mutica 243; quadrifolia, rsilea pasate in Virginia, Sr Matoniaceae, 119 Matteuccia, 139, 144 struthiopteris, 137 01 _ — PEREZ- at wl R. Ripa. arative Research of Gametophytes Olfersia sae and Olfersia cervina (Dry- opteridaceae), 229 Meniscium, 131, 134, 139, 142-144 Meniso 132 Me seatesa Pe 139, 141, 144 crassifolium, 135, Mesopteris, 132 Meta thes 138-140, 143, 145, 146; dayi, MICKEL, i x a oc) Microgramma heterophylla, 67-70, 73-76 Microlepia mauiensis, 102; strigosa, 102; stri- osa var. mauiensis, 102, 97 Microsorum spectrum var. pentadactylum, 103, 97 Siloingin ge pi 159 Mohria, 119, 120, 123, 125, 127 128-129; ene cafforum, 121, 28, i. Montcomery, J. Review: ophytes of Upper Katanga eee ratte ic of Congo), 2 Morrow, A. a ee Crystals Associated with the Intertracheid Pit Membrane of Woody Fern Botrychium multifidum, 1 MussELMAN, L. J. (se a A. KNEPPER) Myriophyllum pinnatum, 243 oo 139, 141, 146; aoristisora, 143 Nephroeps 94; falcata, lai hirsutula, 180; multi Pe on abineitte 7a Obituary: Rolla Milton Tryon, Jr. (1916-2001), 1 Olfersia alata, 229-238; cervina, Onoclea, 139, 142, 144; sensibilis, 116, 137 nocleopsis 139; hintonii, Oreopteris, 139, 140, 145, 146; limbosperma, 135, 143 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) wings 125, 224; cinnamomea, 30, 31, 36; claytonia Osmundaceae, a 125 Osmundopteris, 83 Oxytropis campestris, 154, 159 PacHEco, L. Trichomanes ribae (Hymenophylla- ceae), a New Filmy Fern from Costa Rica axonomic Notes on Hawaiian Pteridophytes, 97 sider 138-146; beddomei, 138; neva- » ABO, hibakes verticilata, 291 ne —76, 18 Phymatosorus grossus, 180; ie: 67- Picea shnicn, 169; rubens, 242; sitchensis, 154, 156 Pinus contorta, 252 a austroamericana, 180; calomela- Side tiens clavellata, 242; sp., 291 Platycerium bifurcatum, 180 Plenasium, 125 Plesioneuron archboldiae, 135, 139, 143 Pleurozia conchifolia, Pneumatopteris, 139, 141, 144; ecallosa, 135, sei Poa sp., 291 paba titi, 10, 14 Polygonum viviparum, 291 olypodiaceae, 142 Polypodiales, 133 olypodium, 32; rphum, 32, 214, 216; sheeleaaee ; net a ; oe e dun var ecmnine 97, 104; pellucidum var. ““gamma’’, 103; pellucidum var. pel- lucidum, 75, 103, 104; rockii, 97; sax- imontanum, 214, 216; scouleri, 214-228; sibiricum, 214, 216; spec penta- actyium, 103; spectrum var. spectrum, INDEX TO VOLUME 92 sain Riley 240; vulgare, 32-36, 15222, Pavers oo 143, , (226; 229, 235—2397; acrostichoides, 225; ae cern x lon- itis, 225; aleuticum, 288-293; i ns, 208, 209; lonchitis, 225, 291; munitum, 201, 208, Populus balsamifera, 169; canescens, 46; grandidentata, 46; tremuloides, 169 Potentilla norvegica, 169; villosa, 2 R and D. B. LELLINGER. Adiantum argutum, an Unrecognized Species of the A; a Group, . R. Situ. Novelties in Pter- ica, Pronephrium, 139, 141, 142, 144, 146; simplex, 135, 143 Protomarattia, 193 seudocyclosorus 139; gost 135, 143 sarge eg Lge 9, 141, 144-146; , 136, Preceny LZ Pteridaceae, 105 Pter. ium 31, 36, 37; aquilinum, 30, 201, 210, 287; aquilinum ssp. aquilinum, : 9 ri poe um aa 271, ni ; pearcei, 110; = ig a 208; peioanfite, = vitiate, -24 Racomitrion canescens, 154, 159 RANKER, T. A. (se OU Ranker, T. M. and H. z ae Is Gameto- phyte Senile in the Laboratory a Good edictor of Sexuality in Nature? 112 Reproductive Behavior of Cloned Gameto- phytes of Pteridium aquilinum (L.) Kuhn., 270 Rhinanthus crista-galli, 154, 159 Rhododendron maximun, 242 agra squarrosus, 154, 159 A, R. (se ENDOZA a Oe a a JoHNSON-GrRoH) Ropertson, F. W. Reproductive Behavior of oe sat a of Pteridium aqui- m (L.) K Rosa ae wise a a a te 263; stellatus, sea 113-117, 139, 144, 145; * hereon 3-115, 117, 137; pallida, 113-115, , 114; unisora, 114 Salix alozesis 169; bebbiana, 169; rotundifo- 29 pee ar he 180-182 SANCHEZ, Cc. New age Fern from the Dominican Republic Schizaea, se 1 aes 127; elegans, 121, a 124, 126 chizaeaceae, 120; 124, 125, 128 SCHOESSLER, e C. JOHNSON-GROH edum eater tie ssp. Pp ago ia 182; kraussiana, 180; sellata, veal mbrosa, 180 eens ples 156; as 239 Silene uralensis 7 pees ensis, 156 SIMAN, E. and E ea Polypodium vu oe Plants ‘Sporulate Continuously ina do Seasonal Glasshouse Environ- SKOG, 1 peers . KELLOFF) BKOG; Ae oe esis and J. T. MICKEL. Additional aes for Two Subgenara of Anemia (Schizaeaceae) from Data for the Mcnues Intergenic Spacer Region trnL-F and Morphology, 119 SKOGEN, K. (see C. JoHNSON-GROH) H, A. R. (see J. PRapo) ae and R B. CranFILu. Intrafamilial Relatigastaps ef the o_o roid Ferns (Thelypteridaceae), 1 Spharotephanes a5, a 141, 144; penn 136, 137, 143; iehutincinniin, 13 Sphagn 242 gi > apis a en 140, 141, 144; pilosa, ions 132, 139-141, 144; leprieurii, 136, 143 Stenochlaena 139, 144; milnei, 137 302 a = ie D. R. Farrar and C. JoHNson- o New Species of Moonworts eee subg. Botrychium) from Alas- a, 150 Swartz, L. M. and S. J. BRUNSFELD. The Mor- eal: ne Genetic Distinctness of tum as Assessed by Morphometric Anal- ysis and RAPD Markers, 249 TaALBoT, S. L. and S. S. Ta.sor. A New Population of Aleutian Shield Fern (Poly- stichum aleuticum C. Christens.) on Asak 6 9 ‘soétes eee gneri, an In- terspecific Hybr id of I. bolanderi x I. echinospora (Isoétaceae), 161 Tectaria incisa, 180 The Morphological and Genetic pesuinaes of Botrychium minganense an nula- tum as Assessed by sai pr RAPD Markers, 2 Thelypteridaceae, 131, 132, 138, 140, 142-145 Thelypteris 131, 138-141, 144-146; burksio- Tofieldia coccinea, 291 Torreya yunnanesis, 10 AMERICAN FERN JOURNAL: VOLUME 92 NUMBER 4 (2002) pads 48, 193; krausii, 95; membrana- ceum, 95; punctatum subsp punctatum, ten e€ 294-295; rupestre, 292 aie ge euged be enophyllaceae), A Film n from Costa Rica and Panama, 294 Trichoneuron, 131, 6 Trigonospora 139, a cline 136, 143 Tsuga canaden Tsuga haere ale 265 Utriculuaria, 2 Vaccinium 242; uliginosum, 169 satan acutiloba, 291 iola adunca, 239; langsdorfii, 291 he 48; appalachiana, 242 Vittaria-type, 229, 230, 234, 235, 236, 237 Wacner, F. S. (see P. F. Zika) -»» & J. R. Grant. Botrychium alaskense. a New eceiait from the Interior of Alaska, 16 Wa ter, C. (see T. J. Hons Watkins, J. E., Jk. and D. R. Rarrar. A New seed - an old Fern from North Alaba- dene Pa i (see C. L. KELLOFF WiLson, K. A. Continued Pteridophyte Invasion oO waii, 179 ica, 124 Woodsia 139, 145, 224; polystichoides, 137 a P. F., E. R. ALVerson, W. H. WAGNER, jr. FP. S. Wacner. Botrychium hesperium “i Wallowa “pa of Oregon, 239 tila E. A. (see J. E. Skoc) g - xara - : - o oa 2 i Tate a ae —_ iN, ey - * yj oS . i OeOme) a Sis 7 = “ou a S 3 | ; _ on. ae lay 7 ales eae Ht in i ae. ; oe 7 : - Ee c= Se : ee a 7 oe) - i, OD ihe ; | - - p _ othe \J97 7 ; : - - ai) = | | SS ! : oo _ ; : 5 =a) i i o ad 7 5 leg a. a oo 7 7 Br a "2 ° a x - eal es 7 - oF am cs Bia, Gy, oe oe eh © om A ” raat on sare) ey ——- Penagle, ay 4... 4 : 2 = Ripe icc ae YJ ; a a ach aren. Oia ‘ a he oe Oe i +) : 7 4 : : - K i ' i ff 7 ri - ia = ; ce, 7 ay 7 bis ra 7 7 z = © he Ul! PAR in Bia 0 we syed 2 a, ak a a aera ss Te att 0° ees ee Saclaiinseeiliael nm. % a is 1 eu i Y a : : i Se Yee ee ee eS ia. | Saree Vcr ae eA S30 9S Ret ow Fee ? : : - ie, ae i pe f se " - me - - ¥ 5 7 _ re : % : 4 a S. : ‘ ra = ie Fs one 7 " q _—— . & S a on 7 7 = 7 Yen mis 2 a as oo | : " 5 + 2B a : a5) | _ : ! in x i! ae > ae a < nt : ie = ee a h WA ee ; = a te - i 7 ON ee 7 oe i bi ; tg Non 1 aes con ae a a Tis Vin Sas 7 = 7 >. e a - ve : : ws 7 2 “i 7 a 7 i 7 ae oh tc : 7 ot bask : ; 7 Fan ee 7 - ne a at) ‘2 — a 7 7 > poo YS m , 7 ae ‘1 oe bala Mae Pas _ : ‘piey i oe ne mee - aes. = : ho Seen < a ee ave a ee Ve ec 7 wot aren | ‘ f a oe : Sy eee Ol. ' 7 bag a > G = x - / =i. - ae yee : ere See ae. en eae pea fae a 7 _ ee See! i > as Sa, hate ie ie ; ae —_ _ = ve ane Mee on rl ete ee To ere : Pr sas me’. 2." : io: ag | be ee tn a ae : ae . . _ 7 | ee 7 : > 7 - = 7 aes a la ae jen ee ar be stiine 7 ra i Hi If) ti INFORMATION FOR Au : Ml Authors are encouraged to submit manuscripts pertinent to pteridology for pub- lication in the American Fern Journal. Manuscripts should be sent to the Editor. Acceptance of papers for publication depends on merit as judged by two or more referees. Authors are encouraged to contribute toward publishing costs; howev: the payment or non-payment of page charges will affect neither the cane of manuscripts nor the date of publication. Authors should adhere to the following guidelines; manuscripts not so prepared may be returned for revision prior to review. Submit manuscripts in triplicate (xerocopies acceptable), including review copies of illustrations and originals of illustrations. After review, submission of final versions of manuscripts on diskette (in PC- or Mac-compatible formats) is strongly encouraged. Use standard 8% by 11 inch paper of good quality, not “erasable” paper. Double space manuscripts throughout, including title, authors’ names and addresses, short, informative ab- stract, text (including heads and keys), literature cited, tables (separate from text), and figure captions (grouped as consecutive paragraphs separate from figures). Arrange parts of manuscript in order just given. Include author’s name and page number in upper right corner of every sheet. Provide margins of at least 25 mm all around on typed pages. Do not submit right-justified copy, avoid footnotes, and do not break words at ends of lines. Make table headings and figure captions self-explanatory. Use S.I. (metric) units for all measures (e.g., distance, elevation, weight) unless quoted or cited from another source (e.g., specimen citations). For nomenclatural matter (i.e., synonymy and typification), use one paragraph per basionym (see Regnum Veg. 58:39—40. 1968). Abbreviate titles of serial publi- cations according to Botanico-Periodicum-Huntianum (Lawrence et al., 1968, Hunt Botanical Library, Pittsburgh) and its supplement (1991). References cited only as part of nomenclatural matter are not included in literature cited. For shorter notes and reviews, omit the abstract and put all references parenthetically in text. Use Index Herbariorum (Regnum Veg. 120:1—693. 1990) for designations of her- a. Illustrations should be proportioned to fit page width with caption on the same . Provide margins of at least 25 mm on all illustrations. For continuous-tone illustrations, design originals for reproduction without reduction or by uniform amount. In composite blocks, abut edges of adjacent photographs. Avoid com- bining continuous-tone and line-copy in single illustrations or blocks. Coordinate sequence and numbering of figures (and of tables) with order of citation in text. Explain scales and symbols in figures themselves, not in captions. Include a scale and reference to latitude and longitude in each map. Proofs and reprint order forms are sent to authors by the printer. Authors should send corrected proofs to the editor and reprint orders to the printer. Authors will be assessed charges for extensive alterations made after type has been set. For other matter of form or style, consult recent issues of American Fern Jour- nal and The Chicago Manual of Style, 14th ed. (1993, Univ. Chicago Press, Chicago). Occasionally, departure from these guidelines may be justified. Authors are encouraged to consult the editor for assistance with any aspect of manuscript preparation. Papers longer than 32 printed pages may be sent to the Editor of Preridologia (Memoir Editor, see cover 2). PTERIDOLOGIA ISSUES IN PRINT The following issues of Pteridologia, the memoir series of the American Fern Society, are available for purchase: 1. Wagner, David H. 1979. Systematics of Polystichum in Western North America North of Mexico. 64 pp. $10.00 postpaid. 2A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Choco (Part 1: Psilotaceae through Dicksoniaceae). 364 pp. $32.00 postpaid. 3. Lellinger, David B. 2002. A Modern Multilingual Glossary for Taxonomic Pteridology. 263 pp. $28.00 postpaid. Send your order with a check or money order to: American Fern Society, Inc., c/o U.S. National Herbarium MRC-166, Smithsonian Institution, Washing- ton, DC 20560. AMERICAN FERN JOURNAL ON MICROFICHE Volumes 1—61 of the American Fern Journal are available as archival quality, silver positive microfiches. Single volumes or the entire run may be purchased. The fiches are easily read with 10 or greater magnification (using a dissecting microscope and transmitted illumination or a fiche reader). Silver negative micro- fiches of vols. 1-50 are also available. The price is $4.00 per volume or $244. per set of 61 volumes, postpaid. Send your inquiry or order with a check or money order to: American Fern Society, Inc., c/o Dr. James D. Montgomery, Ecology III, Inc., R.D. 1, Box 1795, Berwick, PA 18603. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://www.amerfernsoc.org/