Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-qks25 Total loading time: 0 Render date: 2024-08-12T23:08:23.771Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 September 2012

Thomas H. Nash, III
Affiliation:
Arizona State University
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Lichen Biology , pp. 364 - 461
Publisher: Cambridge University Press
Print publication year: 2008

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aarrestad, P. A. and Aamlid, D. (1999). Vegetation monitoring in South-Varanger, Norway – species composition of ground vegetation and its relation to environmental variables and pollution impact. Environmental Monitoring and Assessment, 58, 1–21.CrossRefGoogle Scholar
Abba', S., Ghignone, S. and Bonfante, P. (2006). A dehydration-inducible gene in the truffle Tuber borchii identifies a novel group of dehydrins. BMC Genomics, 7, 39.CrossRefGoogle ScholarPubMed
Acharius, E. (1798). Lichenographiae Svecicae Prodromus. Linköping: D.G. Björn.CrossRefGoogle Scholar
Acharius, E. (1803). Methodus qua omnus detectos Lichenes. Stockholm: F.D.D. Ulrich.Google Scholar
Acharius, E. (1810). Lichenographia universalis. Göttingen: F. Dandewerts.Google Scholar
Acharius, E. (1814). Synopsis methodica lichenum. Lund.Google Scholar
Adams, G. C. and Kropp, B. R. (1996). Athelia arachnoidea, the sexual state of Rhizoctonia carotae, a pathogen of carrot in cold storage. Mycologia, 88, 459–472.CrossRefGoogle Scholar
Adler, M. T. and Calvelo, S. (2002). Parmeliaceae s. str. (lichenized Ascomycetes) from Tierra del Fuego (southern South America) and their world distribution patterns. Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg, 30–32, 9–24.Google Scholar
Adler, M. T., Fazio, A., Bertoni, M. D., et al. (2004). Culture experiments and DNA verification of a mycobiont isolated from Punctelia praesignis (Parmeliaceae, lichenized Ascomycotina). Bibliotheca Lichenologica, 88, 1–8.Google Scholar
Aguiar, L. W., Martau, L., Oliveira, M. L. A. A. and Martins-Mazzitelli, S. M. A. (1998). Efeitos do dióxido de enxofre (SO2) em liquens, Rio Grande do Sul, Brasil. Iheringia Botanica, 50, 67–73.Google Scholar
Ahmadjian, V. (1967 a). The Lichen Symbiosis. Toronto: Blaisdell Publishing.Google Scholar
Ahmadjian, V. (1967 b). A guide to the algae occurring as lichen symbionts: isolation, culture, cultural physiology and identification. Phycologia, 6, 127–160.CrossRefGoogle Scholar
Ahmadjian, V. (1973). Methods of isolating and culturing lichen symbionts and thalli. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 653–659. London: Academic Press.Google Scholar
Ahmadjian, V. (1988). The lichen alga Trebouxia: does it occur free-living?Plant Systematics and Evolution, 158, 243–247.CrossRefGoogle Scholar
Ahmadjian, V. (1993). The Lichen Symbiosis. New York: John Wiley.Google Scholar
Ahmadjian, V. (1995). Lichens are more important than you think. BioScience, 45, 123–124.CrossRefGoogle Scholar
Ahmadjian, V. (2001). Trebouxia: reflections on a perplexing and controversial lichen photobiont. In Symbiosis, ed. Seckbach, J., pp. 373–383. Dordrecht: Kluwer Academic.Google Scholar
Ahmadjian, V. and Heikkilä, H. (1970). The culture and synthesis of Endocarpon pusillum and Staurothelse clopima. Lichenologist, 4, 259–267.CrossRefGoogle Scholar
Ahmadjian, V. and Jacobs, J. B. (1981). Relationship between fungus and alga in the lichen Cladonia cristatella Tuck. Nature (London), 389, 169–172.CrossRefGoogle Scholar
Ahti, T. (1999). Biogeography. Nordic Lichen Flora, 1, 7–8.Google Scholar
Ahti, T. (2000). Cladoniaceae. Flora Neotropica Monograph, 78, 1–362.Google Scholar
Alebic-Juretic, A. and Arko-Pijevac, M. (1989). Air pollution damage to cell membranes in lichens – results of simple biological test applied in Rijeka, Yugoslavia. Water, Air, and Soil Pollution, 47, 25–33.CrossRefGoogle Scholar
Alexopoulos, C. J., Mims, C. W. and Blackwell, M. (1996). Introductory Mycology. 4th edn. New York: John Wiley.Google Scholar
Allan, A. and Fluhr, R. (1997). Two distinct sources of elicited reactive oxygen species in tobacco epidermal cells. Plant Cell, 9, 1559–1572.CrossRefGoogle ScholarPubMed
Allen, J. F., Mullineaux, C. W., Sanders, C. E. and Melis, A. (1989). State transitions, photosystem stoichiometry adjustment and non-photochemical quenching in cyanobacterial cells acclimated to light absorbed by photosystem I or photosystem II. Photosynthesis Research, 22, 157–166.CrossRefGoogle ScholarPubMed
Alpert, P. (1988). Survival of a desiccation-tolerant moss, Grimmia laevigata, beyond its observed microdistributional limits. Journal of Bryology, 15, 219–227.CrossRefGoogle Scholar
Alscher, R. (1984). Effects of SO2 on light-modulated enzyme reactions. In Gaseous Air Pollutants and Plant Metabolism, ed. Koziol, M. J. and Whatley, F. R., pp. 181–200. London: Butterworths.Google Scholar
Alstrup, V. (1992). Effects of pesticides on lichens. Bryonora, 9, 2–4.Google Scholar
Alstrup, V. and Hansen, E. S. (1977). Three species of lichens tolerant of high concentrations of copper. Oikos, 29, 290–293.CrossRefGoogle Scholar
Alstrup, V. and Hawksworth, D. L. (1990). The lichenicolous fungi from Greenland. Meddelelser om Grønland, Bioscience, 31, 1–90.Google Scholar
Amthor, J. S. (1995). Higher plant respiration and its relationships to photosynthesis. In Ecophysiology of Photosynthesis, ed. Schulze, E. D. and Caldwell, M. M., pp. 71–101. Berlin: Springer.Google Scholar
Anagnostidis, K. and Komárek, J. (1985). Modern approach to the classification system of cyanophytes. 1 – Introduction. Archiv für Hydrobiologie, Supplementband, 71 (Algological Studies, 38/39), 291–302.Google Scholar
Anagnostidis, K. and Komárek, J. (1988). Modern approach to the classification system of cyanophytes. 3 – Oscillatoriales. Archiv für Hydrobiologie, Supplementband, 80 (Algological Studies, 50/53), 327–472.Google Scholar
Anagnostidis, K. and Komárek, J. (1990). Modern approach to the classification system of cyanophytes. 5 – Stigonematales. Algological Studies, 59, 1–73.Google Scholar
Anand, M., Laurence, S. and Rayfield, B. (2005). Diversity relationships among taxonomic groups in recovering and restored forests. Conservation Biology, 19, 955–962.CrossRefGoogle Scholar
Andersen, H. L. and Ekman, S. (2005). Disintegration of the Micareaceae (lichenized Ascomycota): a molecular phylogeny based on mitochondrial rDNA sequences. Mycological Research, 109, 21–30.CrossRefGoogle ScholarPubMed
Anonymous (2004). Vagrant lichen charged in elk deaths. Castilleja, 23, 1.
Antoine, M. E. (2004). An ecophysiological approach to quantifying nitrogen fixation by Lobaria oregana. Bryologist, 107, 82–87.CrossRefGoogle Scholar
Antoine, M. E. and McCune, B. (2004). Contrasting fundamental and realized ecological niches with epiphytic lichen transplants in an old-growth Pseudotsuga forest. Bryologist, 107, 163–173.CrossRefGoogle Scholar
Aptroot, A. (1987). Terpenoids in tropical Pyxinaceae (lichenized fungi). In XIV International Botanical Congress, Abstracts, Berlin, ed. Greuter, W., Zimmer, B. and Behnke, H.-D., p. 5–04–7.Google Scholar
Aptroot, A. (2001). Lichenized and saprobic fungal biodiversity of a single Elaeocarpus tree in Papua New Guinea, with the report of 200 species of ascomycetes associated with one tree. Fungal Diversity, 6, 1–11.Google Scholar
Aptroot, A. and Seaward, M. R. D. (2003). Freshwater lichens. In Freshwater Mycology, ed. Tsui, C. K. and Hyde, K. D., pp. 101–110. Hong Kong: Fungal Diversity Press.Google Scholar
Aptroot, A. and Sipman, H. J. M. (1997). Diversity of lichenized fungi in the tropics. In Biodiversity of Tropical Microfungi, ed. Hyde, K. D., pp. 93–106. Hong Kong: Hong Kong University Press.Google Scholar
Aptroot, A. and Sparrius, L. B. (2006). Additions to the lichen flora of Vietnam, with an annotated checklist and bibliography. Bryologist, 109, 358–371.CrossRefGoogle Scholar
Archer, A. W. and Elix, J. A. (1993). Additional new taxa and a new report of Pertusaria (lichenised Ascomycotina) from Australia. Mycotaxon, 49, 143–150.Google Scholar
Archer, D., Eggink, G., Schweizer, M. Stymne, S. and Rathledge, G. (1999). Manipulation of Lipid Metabolism Aimed at the Production of Fatty Acids and Polyketides. Final report, Contr. No. Air 2-CT94-967. Internet Publication www.biomatnet.org.secure/Air/S1045.htm.
Archibald, P. A. (1975). Trebouxia DePuymaly (Chlorophyceae, Chlorococcales) and Pseudotrebouxia (Chlorophyceae, Chlorosarcinales). Phycologia, 14, 125–137.CrossRefGoogle Scholar
Armaleo, D. (1993). Why do lichens make secondary products? In XV International Botanical Congress Abstracts, ed. Furuya, M., p. 11. Yokohama, Japan: International Union of Biological Sciences.Google Scholar
Armaleo, D. and Clerc, P. (1990). Lichen chimeras: DNA analysis suggests that one fungus forms two morphotypes. Experimental Mycology, 15, 1–10.CrossRefGoogle Scholar
Armaleo, D. and Clerc, P. (1995). A rapid and inexpensive method for the purification of DNA from lichens and their symbionts. Lichenologist, 27, 207–213.CrossRefGoogle Scholar
Armstrong, R. A. (1974). A comparison of the growth-curves of the foliose lichen Parmelia conspersa determined by a cross-sectional study and by direct measurement. Environmental and Experimental Botany, 32, 221–227.CrossRefGoogle Scholar
Armstrong, R. A. (1988). Substrate colonization, growth, and competition. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 3–16. Boca Raton: CRC Press.Google Scholar
Armstrong, R. A. (1992). Soredial dispersal from individual soralia in the lichen Hypogymnia physodes (L.) Nyl. Environmental and Experimental Botany, 32, 55–63.CrossRefGoogle Scholar
Armstrong, R. A. (1993). Factors determining lobe growth in foliose lichen thalli. New Phytologist, 124, 675–679.CrossRefGoogle Scholar
Armstrong, R. A. and Smith, S. N. (1992). Lobe growth variation and the maintenance of symmetry in foliose lichen thalli. Symbiosis, 12, 145–158.Google Scholar
Aronson, J. M. (1977). Cell walls and intracellular polysaccharides of Leptomitales. Abstracts Second International Mycological Congress A–L, ed. Bigelow, H. E. and Simmons, E. G., p. 19. Tampa: IMC-2, Inc.Google Scholar
Arup, U. and Grube, M. (1998). Molecular systematics of Lecanora subgenus Placodium. Lichenologist, 30, 415–425.CrossRefGoogle Scholar
Arup, U., Ekman, S., Grube, M., Mattsson, J. E. and Wedin, M. (2007). The sister group relation of Parmeliaceae (Lecanorales, Ascomycota). Mycologia, 99, 42–49.CrossRefGoogle Scholar
Arup, U., Ekman, S., Lindblom, L. and Mattsson, J. E. (1993). High performance thin layer chromatography (HPTLC), an improved technique for screening lichen substances. Lichenologist, 25, 61–71.CrossRefGoogle Scholar
Asahina, Y. and Shibata, S. (1954). Chemistry of Lichen Substances. Tokyo: Japan Society for the Promotion of Science.Google Scholar
Ascaso, C., Wierzchos, J. and los Rios, A. (1995). Cytological investigations of lithobiontic microorganisms in granitic rocks. Botanica Acta, 108, 474–481.CrossRefGoogle Scholar
Aspray, T., Jones, E., Whipps, J. and Bending, G. (2006). Importance of mycorrhization helper bacteria cell density and metabolite localization for the Pinus sylvestris / Lactarius rufus symbiosis. FEMS Microbiology Ecology, 56, 25–33.CrossRefGoogle ScholarPubMed
Augusto, S., Pinho, P., Branquinho, C., et al. (2004). Atmospheric dioxin and furan deposition in relation to land-use and other pollutants: a survey with lichens. Journal of Atmospheric Chemistry, 49, 53–65.CrossRefGoogle Scholar
Avise, J. C. (1998). The history and purview of phylogeography: a personal reflection. Molecular Ecology, 7, 371–379.CrossRefGoogle Scholar
Bacci, E., Calamari, D., Gaggi, C., et al. (1986). Chlorinated hydrocarbons in lichen and moss samples from the Antarctic Peninsula. Chemosphere, 15, 747–754.CrossRefGoogle Scholar
Baćkor, M. and Dzubaj, A. (2004). Short-term and chronic effects of copper, zinc and mercury on the chlorophyll content of four lichen photobionts and related alga. Journal of the Hattori Botanical Laboratory, 95, 271–284.Google Scholar
Baćkor, M. and Fahselt, D. (2004). Physiological attributes of the lichen Cladonia pleurota in heavy metal-rich and control sites near Sudbury (Ont., Canada). Environmental and Experimental Botany, 52, 149–159.CrossRefGoogle Scholar
Baćkor, M. and Váczi, P. (2002). Copper tolerance in the lichen photobiont Trebouxia erici (Chlorophyta). Environmental and Experimental Botany, 48, 11–20.CrossRefGoogle Scholar
Baćkor, M. and Zetikova, J. (2003). Effects of copper, cobalt and mercury on the chlorophyll content of lichens Cetraria islandica and Flavocetraria cucullata. Journal of the Hattori Botanical Laboratory, 93, 175–187.Google Scholar
Baćkor, M., Dvorsky, K. and Fahselt, D. (2003). Influence of invertebrate feeding on the lichen Cladonia pocillum. Symbiosis, 34, 281–291.Google Scholar
Baddeley, M. S., Ferry, B. W. and Finegan, E. J. (1973). Sulphur dioxide and respiration in lichens. In Air Pollution and Lichens, ed. Ferry, B. W., Baddeley, M. S. and Hawksworth, D. L., pp. 299–313. Toronto: University of Toronto Press.Google Scholar
Badger, M. R., Pfanz, H., Büdel, B., Heber, U. and Lange, O. L. (1993). Evidence for the functioning of photosynthetic CO2 concentration mechanisms in lichens containing green algal and cyanobacterial photobionts. Planta, 191, 57–70.CrossRefGoogle Scholar
Bailey, R. H. (1966). Studies on the dispersal of lichen soredia. Journal of the Linnean Society, Botany, 59, 479–490.CrossRefGoogle Scholar
Bailey, R. H. (1976). Ecological aspects of dispersal and establishment in lichens. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 215–247. New York: Academic Press.Google Scholar
Balaguer, L. and Manrique, E. (1995). Factors which determine lichen response to chronic fumigations with sulphur dioxide. Cryptogamic Botany, 5, 215–219.Google Scholar
Balaguer, L., Manrique, E., los Rios, A., et al. (1999). Long-term responses of the green algal lichen Parmelia caperata to natural CO2 enrichment. Oecologia, 119, 166–174.CrossRefGoogle ScholarPubMed
Balaguer, L., Valladares, F., Ascaso, C., et al. (1996). Potential effects of rising tropospheric concentrations of CO2 and O3 on green-algal lichens. New Phytologist, 132, 641–652.CrossRefGoogle Scholar
Bargagli, R., Iosco, F. P. and Barghigiani, C. (1987). Assessment of mercury dispersal in an abandoned mining area by soil and lichen analysis. Water, Air, and Soil Pollution, 36, 219–225.CrossRefGoogle Scholar
Barghigiani, C., Bargagli, R., Siegel, B. Z. and Siegel, S. M. (1990). A comparative study of mercury distribution on the Aeolian volcanoes, Vulcano and Stromboli. Water, Air, and Soil Pollution, 53, 179–188.CrossRefGoogle Scholar
Barkman, J. J. (1958). Phytosociology and Ecology of Cryptogamic Epiphytes. Assen: Van Gorcum.Google Scholar
Barreno, E. (1991). Phytogeography of terricolous lichens in the Iberian Peninsula and the Canary Islands. Botanika Chronika, 10, 199–210.Google Scholar
Barreno, E., Grube, M., Bois, L., et al. (1998). Forum discussion. Lichens: a special case in biogeographical analysis. International Lichenologial Newsletter, 31, 18–24.Google Scholar
Barták, M., Solhaug, K. A., Vráblíková, H. and Gaulaa, Y. (2006). Curling during desiccation protects the foliose lichen Lobaria pulmonaria against photoinhibition. Oecologia, 149, 553–560.CrossRefGoogle ScholarPubMed
Bartók, K. (1999). Pesticide usage and epiphytic lichen diversity in Romanian orchards. Lichenologist, 31, 21–25.Google Scholar
Bates, J. W., Bell, J. N. B. and Farmer, A. M. (1990). Epiphyte recolonisation of oaks along a gradient of air pollution in south-east England, 1979–1990. Environmental Pollution, 68, 81–99.CrossRefGoogle Scholar
Bates, J. W., Bell, J. N. B. and Massara, A. C. (2001). Loss of Lecanora conizaeoides and other fluctuations of epiphytes on oak in S. E. England over 21 years with declining SO2 concentrations. Atmospheric Environment, 35, 2557–2568.CrossRefGoogle Scholar
Bauer, H. (1984). Net photosynthetic CO2 compensation concentrations of some lichens. Zeitschrift für Pflanzenphysiologie, 114, 45–50.CrossRefGoogle Scholar
Beard, K. H. and DePriest, P. T. (1996). Genetic variation within and among mats of the reindeer lichen, Cladina subtenuis. Lichenologist, 28, 171–182.CrossRefGoogle Scholar
Beck, A. (1999). Photobiont inventory of a lichen community growing on heavy-metal-rich rock. Lichenologist, 31, 501–510.CrossRefGoogle Scholar
Beck, A. (2002). Photobionts: diversity and selectivity in lichen symbioses. International Lichenological Newsletter, 35, 18–24.Google Scholar
Beck, A. and Koop, H. U. (2001). Analysis of the photobiont population in lichens using a single-cell manipulator. Symbiosis, 31, 57–67.Google Scholar
Beck, A., Friedl, T. and Rambold, G., (1998). Selectivity of photobiont choice in a defined lichen community: inferences from cultural and molecular studies. New Phytologist, 139, 709–720.CrossRefGoogle Scholar
Beck, A., Kasalicky, T. and Rambold, G. (2002). Myco-photobiontal selection in a Mediterranean cryptogam community with Fulgensia fulgida. New Phytologist, 153, 317–326.CrossRefGoogle Scholar
Beckelhimer, S. L. and Weaks, T. E. (1986). Effects of water transported sediment on corticolous lichen communities. Lichenologist, 18, 339–347.CrossRefGoogle Scholar
Becker, V. E. (1980). Nitrogen fixing lichens in forests of the southern Appalachian Mountains of North Carolina. Bryologist, 83, 29–39.CrossRefGoogle Scholar
Becker, V. E., Reeder, J. and Stetler, R. (1977). Biomass and habitat of nitrogen fixing lichens in an oak forest in the North Carolina Piedmont. Bryologist, 80, 93–99.CrossRefGoogle Scholar
Beckett, A. (1981). Ascospore formation. In The Fungal Spore: Morphogenetic Controls, ed. Turian, G. and Hohl, H. R., pp. 107–129. London: Academic Press.Google Scholar
Beckett, P. J., Boileau, L. J. R., Padovan, D., Richardson, D. H. S. and Nieboer, E. (1982). Lichens and mosses as monitors of industrial activity associated with uranium and lead accumulation patterns. Environmental Pollution (Series B), 4, 91–107.CrossRefGoogle Scholar
Beckett, R. P. (1995). Some aspects of the water relations of lichens from habitats of contrasting water status studied using thermocouple psychrometry. Annals of Botany, 76, 211–217.CrossRefGoogle Scholar
Beckett, R. P. (1997). Pressure-volume analysis of a range of poikilohydric plants implies the existence of negative turgor in vegetative cells. Annals of Botany, 79, 145–152.CrossRefGoogle Scholar
Beckett, R. P. and Brown, D. H. (1983). Natural and experimentally-induced zinc and copper resistance in the lichen genus Peltigera. Annals of Botany, 52, 43–50.CrossRefGoogle Scholar
Beckett, R. P. and Brown, D. H. (1984 a). The control of cadmium uptake in the lichen genus Peltigera. Journal of Experimental Botany, 35, 1071–1082.CrossRefGoogle Scholar
Beckett, R. P. and Brown, D. H. (1984 b). The relationship between cadmium uptake and heavy metal uptake tolerance in the lichen genus Peltigera. New Phytologist, 97, 301–311.CrossRefGoogle Scholar
Beckett, R. P. and Minibayeva, F. V. (2007). Rapid breakdown of exogenous extracellular hydrogen peroxide by lichens. Physiologia Plantarum, 129, 588–596.CrossRefGoogle Scholar
Beckett, R. P., Marschall, M. and Laufer, Z. (2005 a). Hardening enhances photoprotection in the moss Atrichum androgynum during rehydration by increasing fast rather than slow-relaxing quenching. Journal of Bryology, 27, 7–12.CrossRefGoogle Scholar
Beckett, R. P., Mayaba, N., Minibayeva, F. V. and Alyabyev, A. J. (2005 b). Hardening by partial dehydration and ABA increase desiccation tolerance in the cyanobacterial lichen Peltigera polydactylon. Annals of Botany, 96, 109–115.CrossRefGoogle ScholarPubMed
Beckett, R. P., Minibayeva, F. V., Vylegzhanina, N. V. and Tolpysheva, T. (2003). High rates of extracellular superoxide reduction by lichens in the Suborder Peltigerineae correlate with indices of high metabolic activity. Plant, Cell and Environment, 26, 1827–1837.CrossRefGoogle Scholar
Bedeneau, M. (1982). Reproduction in vitro des effets de la pollution par le dioxyde de soufre sur quelques lichens. Annales des Sciences Forestieres, 39, 165–178.CrossRefGoogle Scholar
Bedford, D. J., Schweizer, E., Hopwood, D. A. and Khosla, C. (1995). Expression of a functional fungal polyketide synthase in the bacterium Streptomyces coelicolor A3(2). Journal of Bacteriology, 177, 4544–4548.CrossRefGoogle Scholar
Begora, M. and Fahselt, D. (2001). Photolability of secondary compounds in some lichen species. Symbiosis, 31, 3–22.Google Scholar
Belandria, G., Asta, J. and Nurit, F. (1989). Effects of sulphur dioxide and fluoride on ascospore germination of several lichens. Lichenologist, 21, 79–86.CrossRefGoogle Scholar
Belnap, J. (2001). Factors influencing nitrogen fixation and nitrogen release in biological soil crusts. In Biological Soil Crusts: Structure, Function, and Management, ed. Belnap, J. and Lange, O. L., pp. 241–261. Berlin: Springer.Google Scholar
Belnap, J. (2002). Nitrogen fixation in biological soil crusts from southeast Utah, USA. Biological Fertility of Soils, 35, 128–135.CrossRefGoogle Scholar
Belnap, J. and Lange, O. L. (eds.) (2003). Biological Soil Crusts: Structure, Function, and Management: Ecological Studies 150. Berlin: Springer.CrossRefGoogle Scholar
Belnap, J. and Lange, O. L. (2005 a). Lichens and microfungi in biological soil crusts: community structure, physiology, and ecological functions. In The Fungal Community: Its Organization and Role in the Ecosystem, Vol. 3, ed. Dighton, J., White, J. F. and Oudemans, P., pp. 117–138. Boca Raton: CRC Press.CrossRefGoogle Scholar
Belnap, J. and Lange, O. L. (2005 b). Biological soil crusts and global changes: what does the future hold? In The Fungal Community: Its Organization and Role in the Ecosystem, Vol. 3, ed. Dighton, J., White, J. F. and Oudemans, P., pp. 697–712. Boca Raton: CRC Press.CrossRefGoogle Scholar
Benedict, J. B. (1991). Experiments on lichen growth. II. Effects of a seasonal snow cover. Arctic and Alpine Research, 23, 189–199.CrossRefGoogle Scholar
Benner, J. W. and Vitousek, P. M. (2007). Development of a diverse epiphyte community in response to phosphorous fertilization. Ecological Letters, 10, 628–636.CrossRefGoogle Scholar
Benner, J. W., Conroy, S., Lunch, C. K.Toyoda, N. and Vitousek, P. M. (2007). Phosphorus fertilization increases the abundance and nitrogenase activity of the cyanolichen Pseudocyphellaria crocata in Hawaiian montane forests. Biotropica, 39, 400–405.CrossRefGoogle Scholar
Bennett, J. P. and Wetmore, C. M. (1999). Geothermal elements in lichens of Yellowstone National Park, USA. Environmental and Experimental Botany, 42, 191–200.CrossRefGoogle Scholar
Berger, F. and Aptroot, A. (2003). Further contributions to the flora of lichens and lichenicolous fungi of the Azores. Arquipélago, 19A, 1–12.Google Scholar
Bergman, D. E. and Ebinger, J. E. (1990). Cyanogenesis in the lichen genus Dermatocarpon. Castanea, 55, 207–210.Google Scholar
Beschel, R. E. (1961). Dating rock surfaces by lichen growth and its application to glaciology and physiography (lichenometry). In Geology of the Arctic, Vol. 2, ed. Raasch, G. O., pp. 1044–1062. Toronto: University of Toronto Press.Google Scholar
Bewley, J. D. (1979). Physiological aspects of desiccation tolerance. Annual Review of Plant Physiology, 30, 195–238.CrossRefGoogle Scholar
Bewley, J. D. and Krochko, J. E. (1982). Desiccation tolerance. In Physiological Plant Ecology. Vol. II: Water Relations and Carbon Assimilation, ed. Lange, O. L., Nobel, P. S., Osmond, C. B. and Ziegler, H., pp. 325–378. Encyclopedia of Plant Physiology 12B. Berlin: Springer.CrossRefGoogle Scholar
Bilger, W., Rimke, S., Schreiber, U. and Lange, O. L. (1989). Inhibition of energy-transfer to photosystem II in lichens by dehydration: different properties of reversibility with green and blue-green phycobionts. Journal of Plant Physiology, 134, 261–268.CrossRefGoogle Scholar
Bingle, L. E., Simpson, T. J. and Lazarus, C. M. (1999). Ketosynthase domain probes identify two subclasses of fungal polyketide synthase genes. Fungal Genetics and Biology, 26, 209–223.CrossRefGoogle ScholarPubMed
Bischoff, H. W. and Bold, H. C. (1963). Phycological studies. IV. Some soil algae from Enchanted Rock and related algal species. University of Texas Publication, 6318, 1–95.Google Scholar
Bjelland, T. and Ekman, S. (2005). Fungal diversity in rock beneath a crustose lichen as revealed by molecular markers. Microbial Ecology, 49, 598–603.CrossRefGoogle ScholarPubMed
Bjerke, J. W. (2005). Synopsis of the lichen genus Menegazzia (Parmeliaceae, lichenized Ascomycotina) in South America. Mycotaxon, 91, 423–454.Google Scholar
Bjerke, J. W., Elvebakk, A., Dominguez, B. and Dahlbäck, A. (2005). Seasonal trends in usnic acid concentrations of Arctic, alpine and Patagonian populations of the lichen Flavocetraria nivalis. Phytochemistry, 66, 337–344.CrossRefGoogle ScholarPubMed
Bjerke, J. W., Lerfall, K. and Elvebakk, A. (2002). Effects of ultraviolet radiation and PAR on the content of usnic and divaricatic acids in two arctic-alpine lichens. Photochemical and Photobiological Sciences, 1, 678–685.CrossRefGoogle ScholarPubMed
Bjerke, J. W., Zielke, M. and Solheim, B. (2003). Long-term impacts of simulated climatic change on secondary metabolism, thallus structure and nitrogen fixation activity in two cyanolichens from the Arctic. New Phytologist, 159, 361–367.CrossRefGoogle Scholar
Björkman, O. (1981). Responses to different quantum flux densities. In Physiological Plant Ecology. Vol. I: Responses to the Physical Environment, ed. Lange, O. L., Nobel, P. S., Osmond, C. B. and Ziegler, H., pp. 57–108. Encyclopedia of Plant Physiology 12A. Berlin: Springer.CrossRefGoogle Scholar
Björkman, O., Boardman, N. K., Anderson, J. A., et al. (1972). Effect of light intensity during growth of Atriplex patula on the capacity of photosynthetic reactions, chloroplast components and structure. Carnegie Institution Washington Yearbook, 71, 115–135.Google Scholar
Black, M. and Pritchard, H. W. (2002). Desiccation and Survival in Plants: Drying Without Dying. Oxon: CABI Publishing.CrossRefGoogle Scholar
Blaha, J., Baloch, E. and Grube, M. (2006). High photobiont diversity associated with the euryoecious lichen-forming ascomycete Lecanora rupicola (Lecanoraceae, Ascomycota). Biological Journal of the Linnean Society, 88, 283–293.CrossRefGoogle Scholar
Blanco, O., Crespo, A., Divakar, P. K., Elix, J. A. and Lumbsch, H. T. (2005). Molecular phylogeny of parmotremoid lichens (Ascomycota, Parmeliaceae). Mycologia, 97, 150–159.CrossRefGoogle Scholar
Blanco, O., Crespo, A., Elix, J. A., Hawksworth, D. L. and Lumbsch, H. T. (2004). A molecular phylogeny and a new classification of parmelioid lichens containing Xanthoparmelia-type lichenan (Ascomycota: Lecanorales). Taxon, 53, 959–975.CrossRefGoogle Scholar
Blanco, O., Crespo, A., Ree, R. H. and Lumbsch, H. T. (2006). Major clades of parmelioid lichens (Parmeliaceae, Ascomycota) and the evolution of their morphological and chemical diversity. Molecular Phylogenetics and Evolution, 39, 52–69.CrossRefGoogle ScholarPubMed
Blum, O. B. (1973). Water relations. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 381–400. New York: Academic Press.Google Scholar
Boardman, N. K. (1977). Comparative photosynthesis of sun and shade plants. Annual Review of Plant Physiology, 28, 355–377.CrossRefGoogle Scholar
Boardman, N. K., Anderson, J. M., Thorne, S. E. and Björkman, O. (1972). Photochemical reactions of chloroplasts and components of the photosynthetic electron transport chain in two rainforest species. Carnegie Institution Washington Yearbook, 71, 107–114.Google Scholar
Boileau, L. J. R., Beckett, P. J., Lavoie, P., Richardson, D. H. S. and Nieboer, E. (1982). Lichens and mosses as monitors of industrial activity associated with uranium mining in northern Ontario, Canada. I. Field procedures, chemical analyses and interspecies comparisons. Experimental Pollution (Series B), 4, 69–84.CrossRefGoogle Scholar
Boison, G., Mergel, A., Jolkver, H. and Bothe, H. (2004). Bacterial life and dinitrogen fixation at a gypsum rock. Applied and Environmental Microbiology, 70, 7070–7077.CrossRefGoogle Scholar
Boissière, M.-C. (1982). Cytochemical ultrastructure of Peltigera canina: some factors relating to its symbiosis. Lichenologist, 14, 1–28.CrossRefGoogle Scholar
Boissière, M.-C. (1987). Ultrastructural relationship between the composition and the structure of the cell wall of the mycobiont of two lichens. Bibliotheca Lichenologica, 25, 117–123.Google Scholar
Bold, H. C. and Wynne, M. J. (1985). Introduction to the Algae. Stucture and Reproduction. 2nd edn. Englewood Cliffs: Prentice Hall.Google Scholar
Boonpragob, K., Nash, T. H., III, and Fox, C. A. (1989). Seasonal deposition patterns of acidic ions and ammonium to the lichen Ramalina menziesii Tayl. in southern California. Environmental and Experimental Botany, 29, 187–197.CrossRefGoogle Scholar
Borecký, J. and Vercesi, A. (2005). Plant uncoupling mitochondrial protein and alternative oxidase: energy metabolism and stress. Bioscience Reports, 25, 271–286.CrossRefGoogle Scholar
Boreham, S. (1992). A study of corticolous lichens on London plane Platanus × hybrida trees in West Ham Park, London. London Naturalist, 71, 61–71.Google Scholar
Bothe, H. and Loos, E. (1972). Effect of far red light and inhibitors on nitrogen fixation and photosynthesis in the blue-green alga Anabaena cylindrica. Archiv für Microbiologie, 86, 241–254.CrossRefGoogle Scholar
Bothe, H., Distler, E. and Eisbrenner, G. (1978). Hydrogen metabolism in blue-green algae. Biochimie, 60, 277–289.CrossRefGoogle ScholarPubMed
Bottomley, P. J. and Stewart, W. D. P. (1977). ATP and nitrogenase activity in nitrogen-fixing heterocystous blue-green algae. New Phytologist, 79, 625–638.CrossRefGoogle Scholar
Boucher, V. L. and Nash, T. H., III (1990 a). Growth pattern in Ramalina menziesii in California: coastal vs. inland populations. Bryologist, 93, 295–302.CrossRefGoogle Scholar
Boucher, V. L. and Nash, T. H., III (1990 b). The role of the fruticose lichen Ramalina menziesii in the annual turnover of biomass and macronutrients in a blue oak woodland. Botanical Gazette, 151, 114–118.CrossRefGoogle Scholar
Boustie, J. and Grube, M. (2005). Lichens – a promising source of bioactive secondary metabolites. Plant Genetic Resources, 3, 273–287.CrossRefGoogle Scholar
Bowker, M. A., Belnap, J., Davidson, D. W. and Phillips, S. L. (2005). Evidence for micronutrient limitation of biological soil crusts: importance to arid-lands restoration. Ecological Applications, 15, 1941–1951.CrossRefGoogle Scholar
Branquinho, C., Brown, D. H. and Catarino, F. (1997). The cellular location of Cu in lichens and its effects on membrane integrity and chlorophyll fluorescence. Environmental and Experimental Botany, 38, 165–179.CrossRefGoogle Scholar
Brightman, F. H. and Seaward, M. R. D. (1977). Lichens of man-made substrates. In Lichen Ecology, ed. Seaward, M. R. D., pp. 253–293. London: Academic Press.Google Scholar
Broady, P. A. and Ingerfeld, M. (1993). Three new species and a new record of chaetophoracean (Chlorophyta) algae from terrestrial habitats in Antarctica. European Journal of Phycology, 28, 25–31.CrossRefGoogle Scholar
Brochmann, C., Gabrielsen, T. M., Nordal, I., Landvik, J. Y. and Elven, R. (2003). Glacial survival or tabula rasa? The history of North Atlantic biota revisited. Taxon, 52, 417–450.CrossRefGoogle Scholar
Brock, T. D. (1978). Thermophilic Microorganisms and Life at High Temperatures. New York: Springer.CrossRefGoogle Scholar
Brodo, I. M. (1973). Substrate ecology. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 401–441. New York: Academic Press.Google Scholar
Brodo, I. M. (1978). Changing concepts regarding chemical diversity in lichens. Lichenologist, 10, 1–11.CrossRefGoogle Scholar
Brodo, I. M. and Richardson, D. H. S. (1978). Chimeroid associations in the genus Peltigera. Lichenologist, 10, 157–170.CrossRefGoogle Scholar
Brodo, I. M., Sharnoff, S. D. and Sharnoff, S. (2001). Lichens of North America. New Haven: Yale University Press.Google Scholar
Brooks, D. R. (2004). Reticulations in historical biogeography: the triumph of time over space in evolution. In Frontiers of Biogeography: New Directions in the Geography of Nature, ed. Lomolino, M. V. and Heaney, L. R., pp. 125–144. Sunderland: Sinauer Associates.Google Scholar
Brouwer, R. (1962). Distribution of dry matter in the plant. Netherland Journal of Agricultural Sciences, 10, 399–408.Google Scholar
Brown, D. H. (1972). The effect of Kuwait crude oil and a solvent emulsifier on the metabolism of the marine lichen, Lichina pygmaea. Marine Biology, 12, 309–315.CrossRefGoogle Scholar
Brown, D. H. (1992). Impact of agriculture on bryophytes and lichens. In Bryophytes and Lichens in a Changing Environment, ed. Bates, J. W. and Farmer, A. M., pp. 259–283. Oxford: Clarendon Press.Google Scholar
Brown, D. H. and Beckett, R. P. (1984). Uptake and effect of cations on lichen metabolism. Lichenologist, 16, 173–188.CrossRefGoogle Scholar
Brown, D. H. and Beckett, R. P. (1985). The role of the cell wall in the intracellular uptake of cations by lichens. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 247–258. New York: Plenum Press.CrossRefGoogle Scholar
Brown, D. H., Ascaso, C. and Rapsch, S. (1987). Ultrastructural changes in the pyrenoid of the lichen Parmelia sulcata stored under controlled conditions. Protoplasma, 136, 136–144.CrossRefGoogle Scholar
Brown, D. H., MacFarlane, J. D. and Kershaw, K. A. (1983). Physiological-environmental interactions in lichens. XVI. A re-examination of resaturation respiration phenomena. New Phytologist, 93, 237–246.CrossRefGoogle Scholar
Brown, D. H., Standell, C. J. and Miller, J. E. (1995). Effects of agricultural chemicals on lichens. Cryptogamic Botany, 5, 220–223.Google Scholar
Brown, M. J., Jarman, S. K. and Kantvilas, G. (1994). Conservation and reservation of non-vascular plants in Tasmania, with special reference to lichens. Biodiversity and Conservation, 3, 263–278.CrossRefGoogle Scholar
Brown, R. M. Jr. and Bold, H. C. (1964). Comparative studies of the algal genera Tetracystis and Chlorococcum. Phycological Studies V. University of Texas Publications, 6417, 1–213.Google Scholar
Brunauer, G. and Stocker-Wörgötter, E. (2005). Culture of lichen fungi for future production of biologically active compounds. Symbiosis, 38, 187–201.Google Scholar
Brunauer, G., Grube, M., Muggia, L. and Stocker-Wörgötter, E. (2006). Gene bank of PKS from the mycobiont of Xanthoria elegans. Published in NCBI Database.
Brunner, U. and Honegger, R. (1985). Chemical and ultrastructural studies on the distribution of sporopollenin-like biopolymers in 6 genera of lichen phycobionts. Canadian Journal of Botany, 63, 2221–2230.CrossRefGoogle Scholar
Bruns, T. D., White, T. J. and Taylor, J. W. (1991). Fungal molecular systematics. Annual Review of Ecology and Systematics, 22, 525–564.CrossRefGoogle Scholar
Bruteig, I. E. (1993). The epiphytic lichen Hypogymnia physodes as biomonitor of atmospheric nitrogen and sulphur deposition in Norway. Environmental Monitoring and Assessment, 26, 27–47.CrossRefGoogle ScholarPubMed
Bryant, J. P., Chapin, F. S. III and Klein, D. R. (1983). Carbon nutrient balance of boreal plants in relation to vertebrate herbivory. Oikos, 40, 357–368.CrossRefGoogle Scholar
Bubrick, P. and Galun, M. (1980 a). Proteins from the lichen Xanthoria parietina which bind to phycobiont cell walls. Correlation between binding patterns and cell wall cytochemistry. Protoplasma, 104, 167–173.CrossRefGoogle Scholar
Bubrick, P. and Galun, M. (1980 b). Symbiosis in lichens: differences in cell wall properties of freshly isolated and cultured phycobionts. FEMS Microbiology Letters, 7, 311–313.CrossRefGoogle Scholar
Bubrick, P., Galun, M. and Frensdorff, A. (1981). Proteins from the lichen Xanthoria parietina which bind to phycobiont cell walls. Localization in the intact lichen and cultured mycobiont. Protoplasma, 105, 207–211.CrossRefGoogle Scholar
Bubrick, P., Galun, M. and Frensdorff, A. (1984). Observations on free-living Trebouxia de Puymaly and Pseudotrebouxia Archibald, and evidence that both symbionts from Xanthoria parietina (L.) Th. Fr. can be found free-living in nature. New Phytologist, 97, 455–462.CrossRefGoogle Scholar
Buchauer, M. J. (1973). Contamination of soil and vegetation near a zinc smelter by zinc, cadmium, copper, and lead. Environmental Science and Technology, 7, 131–135.CrossRefGoogle Scholar
Büdel, B. (1982). Phycobionten der Lichinaceen. Diplom-Thesis. Marburg: Universität Marburg.
Büdel, B. (1987). Zur Biologie und Systematik der Flechtengattungen Heppia und Peltula im südlichen Afrika. Bibliotheca Lichenologica, 23, 1–105.Google Scholar
Büdel, B. (1990). Anatomical adaptations to the semiarid/arid environment in the lichen genus Peltula. Bibliotheca Lichenologica, 38, 47–61.Google Scholar
Büdel, B. (1992). Taxonomy of lichenized procaryotic blue-green algae. In Algae and Symbioses, ed. Reisser, W., pp. 301–324. Bristol: Biopress Limited.Google Scholar
Büdel, B. and Henssen, A. (1983). Chroococcidiopsis (Cyanophyceae), a phycobiont in the lichen family Lichinaceae. Phycologia, 22, 367–375.CrossRefGoogle Scholar
Büdel, B. and Henssen, A. (1988). Zwei neue Peltula-Arten von Südafrika. International Journal of Mycology and Lichenology, 2, 235–249.Google Scholar
Büdel, B. and Lange, O. L. (1991). Water status of green and blue-green phycobionts in lichen thalli after hydration by water vapor uptake: do they become turgid?Botanica Acta, 104, 361–366.CrossRefGoogle Scholar
Büdel, B. and Lange, O. L. (1994). The role of cortical and epicortical layers in the lichen genus Peltula. Cryptogamic Botany, 4, 262–269.Google Scholar
Büdel, B. and Wessels, D. C. (1986). Parmelia hueana Gyeln., a vagrant lichen from the Namib Desert, SWA/Namibia. I. Anatomical and reproductive adaptations. Dinteria, 18, 3–15.Google Scholar
Bull, W. B. (1996). Dating San Andreas fault earthquakes with lichenometry. Geology, 24, 111–114.2.3.CO;2>CrossRefGoogle Scholar
Bull, W. B. and Brandon, M. T. (1998). Lichen dating of earthquake-generated regional rockfall events, Southern Alps, New Zealand. Geological Society of America Bulletin, 110, 60–84.2.3.CO;2>CrossRefGoogle Scholar
Bull, W. B., King, J., Kong, F., Moutoux, T. and Phillips, W. M. (1994). Lichen dating of coseismic landslide hazards in alpine mountains. Geomorphology, 10, 253–264.CrossRefGoogle Scholar
Bunce, H. W. F. (1996). Methods of monitoring smelter emission effects on a temperate rain forest. Fluoride, 29, 241–251.Google Scholar
Bungartz, F., Garvie, L. A. J. and Nash, T. H., III (2004). Anatomy of the endolithic Sonoran Desert lichen Verrucaria rubrocincta Breuss: implications for biodeterioration and biomineralization. Lichenologist, 36, 55–73.CrossRefGoogle Scholar
Burkholder, P. R. and Evans, A. W. (1945). Further studies on the antibiotic activity of Lichens. Bulletin of the Torrey Botanical Club, 72, 157–164.CrossRefGoogle Scholar
Burkholder, P. R., Evans, A. W., McVeigh, I. and Thornton, H. K. (1944). Antibiotic activity of Lichens. Proceedings of the National Academy of Sciences, USA, 30, 250–255.CrossRefGoogle ScholarPubMed
Buschbom, J. and Barker, D. (2006). Evolutionary history of vegetative reproduction in Porpidia s.l. (lichen-forming Ascomycota). Systematic Biology, 55, 417–484.CrossRefGoogle Scholar
Buschbom, J. and Mueller, G. M. (2004). Resolving evolutionary relationships in the lichen-forming genus Porpidia and related allies (Porpidiaceae, Ascomycota). Molecular Phylogenetics and Evolution, 32, 66–82.CrossRefGoogle Scholar
Buschbom, J. and Mueller, G. M. (2005). Testing “species pair” hypotheses: evolutionary processes in the lichen-forming species complex Porpidia flavocoerulescens and Porpidia melinodes. Molecular Biology and Evolution, 23, 574–586.CrossRefGoogle ScholarPubMed
Butin, H. (1954). Physiologisch-ökologische Untersuchungen über den Wasserhaushalt und die Photosynthese bei Flechten. Biologisches Zentralblatt, 73, 459–502.Google Scholar
Butler, M. J. and Day, A. W. (1998). Fungal melanins: a review. Canadian Journal of Microbiology, 44, 1115–1136.CrossRefGoogle Scholar
Bychek-Guschina, I. A., Kotlova, E. R. and Heipieper, H. (1999). Effects of sulfur dioxide on lichen lipids and fatty acids. Biochemistry (Moscow), 64, 61–65.Google ScholarPubMed
Calatayud, A., Sanz, M.-J., Calvo, E., Barreno, E. and Valle-Tascon, S. (1996). Chlorophyll a fluorescence and chlorophyll content in Parmelia quercina thalli from a polluted region of northern Castellón (Spain). Lichenologist, 28, 49–65.CrossRefGoogle Scholar
Calatayud, A., Tempe, P. J. and Barreno, E. (2000). Chlorophyll a fluorescence emission, xanthophylls cycle activity, and net photosynthetic responses to ozone in some foliose and fruticose lichen species. Photosynthetica, 38, 281–286.CrossRefGoogle Scholar
Caldwell, C. F., Turano, F. J. and McMahon, M. B. (1998). Identification of two cytosolic ascorbate peroxidase cDNAs from soybean leaves and characterization of their products by functional expression in E. coli. Planta, 204, 120–126.Google ScholarPubMed
Calvelo, S. and Liberatore, S. (2001). Checklist of Argentinian lichens (version 2). Online: www.biologie.uni-hamburg.de/checklists.argen_12.htm.
Campbell, D. (1996). Complementary chromatic adaptation alters photosynthetic strategies in the cyanobacterium Calothrix. Microbiology, 142, 1255–1263.CrossRefGoogle Scholar
Campbell, D., Hurry, V., Clarke, A. K., Gustafsson, P. and Öquist, G. (1998). Chlorophyll fluorescence analysis of cyanobacterial photosynthesis and acclimation. Microbiology and Molecular Biology Reviews, 62, 667–683.Google ScholarPubMed
Cane, D. E., Walsh, C. T. and Khosla, C. (1998). Harnessing the biosynthetic code: combinations, permutations, and mutations. Science, 282, 63–68.CrossRefGoogle ScholarPubMed
Cardinale, M., Puglia, A. M. and Grube, M. (2006). Molecular analysis of lichen-associated bacterial communities. FEMS Microbiology Ecology, 57, 484–495.CrossRefGoogle ScholarPubMed
Carlberg, G. E., Ofstad, E. B., Drangsholt, H. and Steinnes, E. (1983). Atmospheric deposition of organic micropollutants in Norway studied by means of moss and lichen analysis. Chemosphere, 12, 341–356.CrossRefGoogle Scholar
Carter, N. E. A. and Viles, H. A. (2003). Experimental investigations into the interactions between moisture, rock surface temperatures and an epilithic lichen cover in the bioprotection of limestone. Building and Environment, 38, 1225–1234.CrossRefGoogle Scholar
Carter, N. E. A. and Viles, H. A. (2005). Bioprotection explored: the story of a little known earth surface process. Geomorphology, 67, 273–281.CrossRefGoogle Scholar
Case, J. W. and Krouse, H. R. (1980). Variations in sulphur content and stable sulphur isotope composition of vegetation near a SO2 source at Fox Creek, Alberta, Canada. Oecologia, 44, 248–257.CrossRefGoogle Scholar
Casely, A. F. and Dugmore, A. J. (2004). Climate change and ‘anomalous’ glacier fluctuations: the southwest outlets of Myrdalsjokull, Iceland. Boreas, 33, 108–122.CrossRefGoogle Scholar
Casselman, K. D. (2001). Lichen Dyes: The New Source Book. Mineola, NY: Dover Publications.Google Scholar
Castenholz, R. W. and Waterbury, J. B. (1989). Group I. Cyanobacteria. In Bergey's Manual of Systematic Bacteriology, Vol. 3, ed. Staley, J. T., Bryant, P., Pfennig, N and Holt, J. G., pp. 1710–1806. Baltimore: Williams and Wilkins.Google Scholar
Cavalier-Smith, T. (1987). The origin of fungi and pseudofungi. In Evolutionary Biology of the Fungi, ed. Rayner, A. D. M., Brasier, C. N. and Moore, D., pp. 339–353. Cambridge: Cambridge University Press.Google Scholar
Chamberlain, A. C. (1970). Interception and retention of radioactive aerosols by vegetation. Atmospheric Environment, 4, 57–78.CrossRefGoogle ScholarPubMed
Chapin, F. S. III (1991). Integrated responses of plants to stress. BioScience, 41, 29–36.CrossRefGoogle Scholar
Chapin, F. S. III, Bloom, A. J., Field, C. B. and Waring, R. H. (1987). Plant responses to multiple environmental factors. BioScience, 37, 49–57.CrossRefGoogle Scholar
Chapman, R. L. (1984). An assessment of the current state of our knowledge of the Trentepohliaceae. In Systematics of the Green Algae. ed. Irvine, D. E. G. and John, D. M., pp. 233–250. London: Academic Press.Google Scholar
Chen, G.-X., Kazmir, J. and Cheniae, G. M. (1992). Photoinhibition of hydroxylamine-extracted photosystem II membranes: studies of the mechanism. Biochemistry, 31, 11 072–11 083.CrossRefGoogle ScholarPubMed
Chen, S., Wu, D. and Wu, J. (1989). Using lichen communities as SO2 pollution monitors. Journal of Nanjing Normal University (Natural Science), 12, 77–82.Google Scholar
Chooi, Y. H., Stalker, D., Louwhoff, S. and Lawrie, A. (2006). The search for a polyketide synthase gene producing beta-orsellinic acid and methylphloroacetophenone as precursors of depsidones and usnic acid in the lichen Chondropsis semiviridis. Poster presentation, International Mycological Congress, IMC 8, Cairns, Australia.
Cislaghi, C. and Nimis, P. L. (1997). Lichens, air pollution and lung cancer. Nature, 387, 463–464.CrossRefGoogle ScholarPubMed
Clarke, A. K., Campbell, D., Gustafsson, P. and Öquist, G. (1995). Dynamic responses of photosystem II and phycobilisomes to changing light in the cyanobacterium Synechococcus sp. PCC 7942. Planta, 197, 553–562.CrossRefGoogle Scholar
Clauzade, G. and Roux, C. (1976). Les Champignons Lichénicoles non Lichénisés. Montpellier: Université des Sciences et Techniques du Languedoc.Google Scholar
Clayden, S. R. (1992). Chemical divergence of eastern North American and European populations of Arctoparmelia centrifuga and their sympatric usnic acid-deficient chemotypes. Bryologist, 95, 1–4.CrossRefGoogle Scholar
Clayden, S. R. (1997). Intraspecific interactions and parasitism in an association of Rhizocarpon lecanorinum and R. geographicum. Lichenologist, 29, 533–545.CrossRefGoogle Scholar
Clerc, P. (2006). Synopsis of Usnea (lichenized Ascomycetes) from the Azores with additional information on the species in Macaronesia. Lichenologist 38, 191–212.CrossRefGoogle Scholar
Codogno, M., Poelt, J. and Puntillo, D. (1989). Umbilicaria freyi spec. nova und der Formenkreis von Umbilicaria hirsuta in Europa. Plant Systematics and Evolution, 165, 55–69.CrossRefGoogle Scholar
Cohn, F. (1853). Untersuchungen über die Entwicklungsgeschichte microskopischer Algen und Pilze. Novorum actorum academiae caesareae leopoldinae-carolinae naturae curiosorum, 24, 101–256.Google Scholar
Coley, P. D. (1988). Effects of plant growth rate and leaf lifetime on the amount and type of anti-herbivore defense. Oecologia, 74, 531–536.CrossRefGoogle ScholarPubMed
Collins, C. R. and Farrar, J. F. (1978). Structural resistances to mass transfer in the lichen Xanthoria parietina. New Phytologist, 31, 71–78.CrossRefGoogle Scholar
Common, R. S. (1991). The distribution and taxonomic significance of lichenan and isolichenan in the Parmeliaceae (lichenized Ascomycotina), as determined by iodine reactions. I. Introduction and methods. II. The genus Alectoria and associated taxa. Mycotaxon, 41, 67–112.Google Scholar
Cook, L. G. and Crisp, M. D. (2005). Directional asymmetry of long-distance dispersal and colonization could mislead reconstructions of biogeography. Journal of Biogeography, 32, 741–754.CrossRefGoogle Scholar
Cook, L. M. (2003). The rise and fall of the carbonaria form of the peppered moth. Quarterly Review of Biology, 78, 399–417.CrossRefGoogle ScholarPubMed
Cowan, D. A., Green, T. G. A. and Wilson, A. T. (1979 a). Lichen metabolism. 1. The use of tritium labelled water in studies of anhydrobiotic metabolism in Ramalina celastri and Peltigera polydactyla. New Phytologist, 82, 489–503.CrossRefGoogle Scholar
Cowan, D. A., Green, T. G. A. and Wilson, A. T. (1979 b). Lichen metabolism. 2. Aspects of light and dark physiology. New Phytologist, 83, 761–769.CrossRefGoogle Scholar
Cowan, I. R., Lange, O. L. and Green, T. G. A. (1992). Carbon-dioxide exchange in lichens: determination of transport and carboxylation characteristics. Planta, 187, 282–294.CrossRefGoogle ScholarPubMed
Cowie, R. H. and Holland, B. S. (2006). Dispersal is fundamental to biogeography and the evolution of biodiversity on oceanic islands. Journal of Biogeography, 33, 193–198.CrossRefGoogle Scholar
Cox, C. B. and Moore, P. D. (2005). Biogeography: An Ecological and Evolutionary Approach. Oxford: Blackwell Publishing.Google Scholar
Cox, R. J., Hitchman, T. S., Byron, K. J., et al. (1997). Post-translational modification of heterologously expressed Streptomyces Type II polyketide synthase acyl carrier proteins. FEBS Letters, 405, 267–272.CrossRefGoogle ScholarPubMed
Coxson, D. S. (1988). Recovery of net photosynthesis and dark respiration on rehydration of the lichen, Cladina mitis, and the influence of prior exposure to sulphur dioxide while desiccated. New Phytologist, 108, 483–487.CrossRefGoogle Scholar
Coxson, D. S. (1991). Impedance measurement of thallus moisture content in lichens. Lichenologist, 23, 77–84.CrossRefGoogle Scholar
Coxson, D. S. and Curteanu, M. (2002). Decomposition of hair lichens (Alectoria sarmentosa and Bryoria spp.) under snowpack in montane forest, Cariboo Mountains, British Columbia. Lichenologist, 34, 395–402.CrossRefGoogle Scholar
Coxson, D. S. and Nadkarni, N. M. (1995). Ecological roles of epiphytes in nutrient cycles of forest ecosystems. In Forest Canopies, ed. Lowman, M. D. and Nadkarni, N. M., pp. 495–543. London: Academic Press.Google Scholar
Coxson, D. S., Webber, M. R. and Kershaw, K. A. (1984). The thermal operating environment of corticolous and pendulous tree lichens. Bryologist, 87, 197–202.CrossRefGoogle Scholar
Craw, R. C., Grehan, J. R. and Heads, M. J. (1999). Panbiogeography: Tracking the History of Life. Oxford Biogeography Series. Oxford: Oxford University Press, 11, 1–238.Google Scholar
Crespo, A., Bridge, P. D., Hawksworth, D. L., Grube, M and Cubero, O. F. (1999). Comparison of rRNA genotypic variability in the lichen-forming fungus Parmelia sulcata from long established and recolonizing sites following sulfur dioxide amelioration. Plant Systematics and Evolution, 217, 177–183.CrossRefGoogle Scholar
Crespo, A., Lumbsch, H. T., Matsson, J.-E., et al. (2007). Testing morphology-based hypotheses of phylogenetic relationships in Parmeliaceae (Ascomycota) using three ribosomal markers and the nuclear RPB-1 gene. Molecular Phylogenetics and Evolution, 44, 812–824.CrossRefGoogle Scholar
Crews, T. E., Kurina, L. M. and Vitousek, P. M. (2001). Organic matter and nitrogen accumulation and nitrogen fixation during early ecosystem development in Hawaii. Biogeochemistry, 52, 259–279.CrossRefGoogle Scholar
Crisci, J. V., Katinas, L. and Posadas, P. (2003). Historical Biogeography: An Introduction. Cambridge: Harvard University Press.Google Scholar
Crisci, J. V., Sala, O. E., Katinas, L. and Posadas, P. (2006). Bridging historical and ecological approaches to biogeography. Australian Systematic Botany, 19, 1–10.CrossRefGoogle Scholar
Crittenden, P. D. (1975). Nitrogen fixation on the glacial drift of Iceland. New Phytologist, 74, 41–49.CrossRefGoogle Scholar
Crittenden, P. D. (1983). The role of lichens in the nitrogen economy of subarctic woodlands: nitrogen loss from the nitrogen-fixing lichen Stereocaulon paschale during rainfall. In Nitrogen as an Ecological Factor, ed. Boddy, L., Marchant, R. and Read, D. J., pp. 43–68. Oxford: Blackwell Scientific Publications.Google Scholar
Crittenden, P. D. (1989). Nitrogen relations of mat-forming lichens. In Nitrogen, Phosphorus and Sulphur Utilization by Fungi, ed. Boddy, L., Marchant, R. and Read, D. J., pp. 243–268. Cambridge: Cambridge University Press.Google Scholar
Crittenden, P. D. (1996). The effect of oxygen deprivation on inorganic nitrogen uptake in an Antarctic macrolichen. Lichenologist, 28, 347–354.CrossRefGoogle Scholar
Crittenden, P. D. (1998). Nutrient exchange in an Antarctic macrolichen during summer snowfall snow melt events. New Phytologist, 139, 697–707.CrossRefGoogle Scholar
Crittenden, P. D. and Kershaw, K. A. (1978). A procedure for the simultaneous measurement of net CO2-exchange and nitrogenase activity in lichens. New Phytologist, 80, 393–401.CrossRefGoogle Scholar
Crittenden, P. D. and Kershaw, K. A. (1979). Studies on lichen-dominated systems. XXII. The environmental control of nitrogenase activity in Stereocaulon paschale in spruce-lichen woodland. Canadian Journal of Botany, 53, 236–254.CrossRefGoogle Scholar
Crittenden, P. D., Kalucka, I. and Oliver, E. (1994). Does nitrogen supply limit the growth of lichens?Cryptogamic Botany, 4, 143–155.Google Scholar
Crittenden, P. D., Llimona, X. and Sancho, L. (2004). Nitrogenase activity in Thyrea spp. – preliminary results. In Book of Abstracts of the 5th IAL Symposium, Lichens in Focus, ed. Randlane, T. and Saag, A., p. 44. Tartu: Tartu University Press.Google Scholar
Crowe, J. H., Crowe, L. M. and Chapman, D. (1984). Preservation of membranes in anhydrobiotic organisms: the role of trehalose. Science, 223, 701–703.CrossRefGoogle Scholar
Culberson, C. F. (1972). Improved conditions and new data for the identification of lichen products by a standardized thin-layer chromatographic method. Journal of Chromatography, 72, 113–125.CrossRefGoogle ScholarPubMed
Culberson, C. F. (1986). Biogenetic relationships of the lichen substances in the framework of systematics. Bryologist, 89, 91–98.CrossRefGoogle Scholar
Culberson, C. F. and Ammann, K. (1979). Standardmethode zur Dünnschichtchromatographie von Flechtensubstanzen. Herzogia, 5, 1–24.Google Scholar
Culberson, C. F. and Culberson, W. L. (1976). Chemosyndromic variation in lichens. Systematic Botany, 1, 325–339.CrossRefGoogle Scholar
Culberson, C. F. and Elix, J. A. (1989). Lichen substances. In Methods in Plant Biochemistry, Vol. 1: Plant Phenolics, ed. Dey, P. M. and Harborne, J. B., pp. 509–535. London: Academic Press.Google Scholar
Culberson, C. F. and Johnson, A. (1976). A standardized two-dimensional thin-layer chromatographic method for lichen products. Journal of Chromatography, 128, 253–259.CrossRefGoogle Scholar
Culberson, C. F., Culberson, W. L. and Johnson, A. (1981). A standardized TLC analysis of β-orcinol depsidones. Bryologist, 84, 16–29.CrossRefGoogle Scholar
Culberson, C. F., Culberson, W. L. and Johnson, A. (1985). Orcinol-type depsides and depsidones in the lichens of the Cladonia chlorophaea group (Ascomycotina, Cladoniaceae). Bryologist, 88, 380–387.CrossRefGoogle Scholar
Culberson, C. F., Culberson, W. L. and Johnson, A. (1988). Gene flow in lichens. American Journal of Botany, 75, 1135–1139.CrossRefGoogle Scholar
Culberson, C. F., Hale, M. E. Jr., Tønsberg, T. and Johnson, A. (1984). New depsides from the lichens Dimelaena oreina and Fuscidea viridis. Mycologia, 76, 148–160.CrossRefGoogle Scholar
Culberson, C. F., Nash, T. H. III and Johnson, A. (1979). 3-α-Hydroxybarbatic acid, a new depside in chemosyndromes of some Xanthoparmeliae with β-orcinol depsides. Bryologist, 82, 154–161.CrossRefGoogle Scholar
Culberson, W. L. (1967). Analysis of chemical and morphological variation in the Ramalina siliquosa species complex. Brittonia, 19, 333–52.CrossRefGoogle Scholar
Culberson, W. L. (1986). Chemistry and sibling speciation in the lichen-forming fungi: ecological and biological considerations. Bryologist, 89, 123–131.CrossRefGoogle Scholar
Culberson, W. L. and Culberson, C. F. (1967). Habitat selection by chemically differentiated races of lichens. Science, 158, 1195–1197.CrossRefGoogle ScholarPubMed
Culberson, W. L. and Culberson, C. F. (1968). The lichen genera Cetrelia and Platismatia (Parmeliaceae). Contributions from the United States National Herbarium, 34, 449–558.Google Scholar
Culberson, W. L. and Culberson, C. F. (1994). Secondary metabolites as a tool in ascomycete systematics: lichenized fungi. In Ascomycetes Systematics: Problems and Perspectives in the Nineties, ed. Hawksworth, D. L.. pp. 155–163. New York: Plenum Press.CrossRefGoogle Scholar
Culberson, W. L., Culberson, C. F. and Johnson, A. (1977). Pseudevernia furfuracea – olivetorina relationships and ecology. Mycologia, 69, 604–614.CrossRefGoogle Scholar
Culberson, W. L., Culberson, C. F. and Johnson, A. (1993). Speciation in the lichens of the Ramalina siliquosa complex (Ascomycotina, Ramalinaceae): gene flow and reproductive isolation. American Journal of Botany, 80, 1472–1481.CrossRefGoogle Scholar
Curtis, C. J., Emmett, B. A., Grant, H., et al. (2005). Nitrogen saturation in UK moorlands: the critical role of bryophytes and lichens in determining retention of atmospheric N deposition. Journal of Applied Ecology, 42, 507–517.CrossRefGoogle Scholar
Czehura, S. J. (1977). A lichen indicator of copper mineralization, Lights Creek District, Plumas County, California. Economic Geology, 72, 796–803.CrossRefGoogle Scholar
Dahl, E. (1998). The Phytogeography of Northern Europe (British Isles, Fennoscandia and Adjacent Areas). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Dahl, E. and Krog, H. (1973). Macrolichens of Denmark, Finland, Norway and Sweden. Oslo: Universitetsforlaget.Google Scholar
Dahlkild, A., Kallersjo, M., Lohtander, K. and Tehler, A. (2001). Photobiont diversity in the Physciaceae (Lecanorales). Bryologist, 104, 527–536.CrossRefGoogle Scholar
Dahlman, L. and Palmqvist, K. (2003). Growth in two foliose tripartite lichens Nephroma arcticum and Peltigera aphthosa – empirical modeling of external versus internal factors. Functional Ecology, 17, 821–831.CrossRefGoogle Scholar
Dahlman, L., Näsholm, T. and Palmqvist, K. (2002). Growth, nitrogen uptake, and resource allocation in the two tripartite lichens Nephroma arcticum and Peltigera aphthosa during nitrogen stress. New Phytologist, 153, 307–315.CrossRefGoogle Scholar
Dahlman, L., Persson, J., Näsholm, T. and Palmqvist, K. (2003). Carbon and nitrogen distribution in the green algal lichens Hypogymnia physodes and Platismatia glauca in relation to nutrient supply. Planta, 217, 41–48.Google ScholarPubMed
David, J. C. (1987). Studies on the genus Epigloea. Systema Ascomycetum, 6, 217–221.Google Scholar
David, K. A. and Fay, P. (1977). Effects of long term treatment with C2H2 on N2-fixing microorganisms. Applied Environmental Microbiology, 34, 640–646.Google Scholar
Davis, W. C., Gries, C. and Nash, T. H. III (2000). The ecophysiological response of the aquatic lichen Hydrothyria venosa to nitrate in terms of weight and photosynthesis over long periods of time. Bibliotheca Lichenologica, 75, 201–208.Google Scholar
Davison, J. (1988). Plant beneficial bacteria. Nature Bio/Technology, 6, 282–286.CrossRefGoogle Scholar
Deason, T. R. and Bold, H. C. (1960). Phycological studies. I. Exploratory studies of Texas soil algae. University of Texas Publications, 6022, 1–70.Google Scholar
Debuchy, R. and Turgeon, B. (2006). Mating-type structure, evolution and function in Euascomycetes. In Growth, Differentiation and Sexuality, The Mycota, Vol. 1, ed. Kües, U. and Fischer, R., pp. 293–323. Heidelberg: Springer.CrossRefGoogle Scholar
De Kok, L. J. and Stulen, I. (1993). Role of glutathione in plants under oxidative stress. In Sulfur Nutrition and Assimilation in Higher Plants, ed. Kok, L. J., Stulen, I., Rennenberg, H., Brunold, C. and Rauser, W. E., pp. 295–313. The Hague: SPB Academic Publishing.Google Scholar
Ríos, los A. and Grube, M. (2000). Host-parasite interfaces of some lichenicolous fungi in the Dacampiaceae (Dothideales, Ascomycota). Mycological Research, 104, 1348–1353.CrossRefGoogle Scholar
Ríos, los A., Wierzchos, J., Sancho, L. G., Green, T. G. A. and Ascaso, C. (2005). Ecology of endolithic lichens colonizing granite in continental Antarctica. Lichenologist, 37, 383–395.CrossRefGoogle Scholar
Deltoro, V. I., Gimeno, C., Calatayud, A. and Barreno, E. (1999). Effects of SO2 fumigations on photosynthetic CO2 gas exchange, chlorophyll a fluorescence emission and antioxidant enzymes in the lichens Evernia prunastri and Ramalina farinacea. Physiologia Plantarum, 105, 648–654.CrossRefGoogle Scholar
Dembitsky, V. M. and Tolstikov, G. A. (2005). Organic Metabolites of Lichens. Novosibirsk: Publishing House of SB RAS.Google Scholar
Demmig-Adams, B. (2006). Linking the xanthophyll cycle with thermal energy dissipation. Photosynthesis Research, 76, 73–80.CrossRefGoogle Scholar
Demmig-Adams, B., Máguas, C., Adams, W. W. III, et al. (1990). Effect of high light on the efficiency of photochemical energy conversion in a variety of lichen species with green and blue-green phycobionts. Planta, 180, 400–409.CrossRefGoogle Scholar
Denison, W. C. (1973). Life in tall trees. Scientific American, 228, 74–80.CrossRefGoogle Scholar
Denison, W. C. (1979). Lobaria oregana, a nitrogen-fixing lichen in old-growth Douglas fir forests. In Symbiotic Nitrogen Fixation in the Management of Temperate Forests, ed. Gordon, J. C., Wheeler, C. T. and Perry, D. A., pp. 266–275. Corvallis: Oregon State University.Google Scholar
DePriest, P. T. (1993 a). Variation in the Cladonia chlorophaea complex. I. Morphological and chemical variation in Southern Appalachian populations. Bryologist, 96, 555–563.CrossRefGoogle Scholar
DePriest, P. T. (1993 b). Variation in southern Appalachian populations of the Cladonia chlorophaea complex. II. ribosomal DNA variation. Bryologist, 97, 117–126.CrossRefGoogle Scholar
DePriest, P. T. (1993 c). Small subunit rDNA variation in a population of lichen fungi due to optional group-I introns. Gene, 134, 314–325.CrossRefGoogle Scholar
DePriest, P. T. (2004). Early molecular investigations of lichen-forming symbionts: 1986–2001. Annual Review of Microbiology, 58, 273–301.CrossRefGoogle ScholarPubMed
DePriest, P. T., Ivanova, N. V., Fahselt, D., Alstrup, V. and Gargas, A. (2000). Sequences of psychrophilic fungi amplified from glacier-preserved ascolichens. Canadian Journal of Botany, 78, 1450–1459.CrossRefGoogle Scholar
Queiroz, A. (2005). The resurrection of oceanic dispersal in historical biogeography. Trends in Ecology and Evolution, 20, 68–73.CrossRefGoogle ScholarPubMed
des Abbayes, H. (1953). Travaux sur les lichens parus de 1939 à 1952. Bulletin Sociétié Botanique de France, 100, 83–123.CrossRefGoogle Scholar
Vera, J.-P., Horneck, G., Rettberg, P. and Ott, S. (2003). The potential of the lichen symbiosis to cope with extreme conditions of outer space. I. Influence of UV radiation and space vacuum on the vitality of lichen symbiosis and germination capacity. International Journal of Astrobiology, 1, 285–293.CrossRefGoogle Scholar
Vera, J.-P., Horneck, G., Rettberg, P. and Ott, S. (2004). The potential of the lichen symbiosis to cope with extreme conditions of outer space. II. Germination capacity of lichen ascospores in response to simulated space conditions. Advances in Space Research, 33, 1236–1243.CrossRefGoogle ScholarPubMed
Dibben, M. J. (1971). Whole-lichen culture in a phytotron. Lichenologist, 5, 1–10.CrossRefGoogle Scholar
Diederich, P. (1996). The lichenicolous heterobasidiomycetes. Bibliotheca Lichenologica, 61, 1–198.Google Scholar
Diederich, P. (2003). Review. Bryologist, 106, 629–630.Google Scholar
Dietz, S., Büdel, B., Lange, O. L. and Bilger, W. (2000). Transmittance of light through the cortex of lichens from contrasting habitats. In Aspects in Cryptogamic Research. Contributions in Honour of Ludger Kappen, ed. Schroeter, B., Schlensog, M. and Green, T. G. A., pp. 171–182. Berlin-Stuttgart: Gebrüder Borntraeger Verlagsbuchhandlung.Google Scholar
Divakar, P. K., Crespo, A., Blanco, O. and Lumbsch, H. T. (2006). Phylogenetic significance of morphological characters in the tropical Hypotrachyna clade of parmelioid lichens (Parmeliaceae. Ascomycota). Molecular Phylogenetics and Evolution, 40, 448–458.CrossRefGoogle Scholar
Döbbeler, P. (1984). Symbiosen zwischen Gallertalgen und Gallertpilzen der Gattung Epigloea (Ascomycetes). Beihefte zur Nova Hedwigia, 79, 203–239.Google Scholar
Döbbeler, P. and Poelt, J. (1981). Arthropyrenia endobrya spec nov., eine hepaticole Flechte mit intrazellulärem Thallus aus Brasilien. Plant Systematics and Evolution, 138, 275–281.CrossRefGoogle Scholar
Döring, H. and Lumbsch, H. T. (1998). Ascoma ontogeny: is this character set of any use in the systematics of lichenized ascomycetes? Lichenologist, 30, 489–500.CrossRefGoogle Scholar
Döring, H. and Wedin, M. (2004). Infraspecific variation within Stereocaulon species complexes – genetic markers, individuals, populations and species. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 26. Tartu: Tartu University Press.Google Scholar
Drew, E. A. (1966). Some aspects of the carbohydrate metabolism of lichens. Ph.D. thesis. Oxford: University of Oxford.
Duchesne, L. C. and Larson, W. (1989). Cellulose and the evolution of plant life. BioScience, 4, 238–241.CrossRefGoogle Scholar
Duguay, K. J. and Klironomos, J. M. (2000). Direct and indirect effects of enhanced UV-B radiation on the decomposing and competitive abilities of saprobic fungi. Applied Soil Ecology, 14, 157–164.CrossRefGoogle Scholar
Durrell, L. W. and Newsom, I. E. (1939). Colorado's Poisonous and Injurious Plants. Fort Collins: Colorado Experiment Station.Google Scholar
Dyer, P. S., Murtagh, G. J. and Crittenden, P. D. (2001). Use of RAPD-PCR DNA fingerprinting and vegetative incompatibility tests to investigate genetic variation within lichen-forming fungi. Symbiosis, 31, 213–229.Google Scholar
Easton, R. M. (1994). Lichens and rocks: a review. Geoscience Canada, 21, 59–76.Google Scholar
Ebach, M. C. and Humphries, C. J. (2003). Ontology of biogeography. Journal of Biogeography, 30, 959–962.CrossRefGoogle Scholar
Edmands, S. (1999). Heterosis and outbreeding depression in interpopulation crosses spanning a wide range of divergence. Evolution, 53, 1757–1768.CrossRefGoogle Scholar
Edwards, T. C. Jr., Cutler, D. R., Geiser, L., Alegria, J. and McKenzie, D. (2004). Assessing rarity of species with low detectability: lichens in Pacific Northwest forests. Ecological Applications, 14, 414–424.CrossRefGoogle Scholar
Egea, J. M. (1996). Catalogue of lichenized and lichenicolous fungi of Morocco. Bocconea, 6, 19–114.Google Scholar
Egea, J. M. and Torrente, P. (1993). The lichen genus Bactrospora. Lichenologist, 25, 211–255.CrossRefGoogle Scholar
Egea, J. M. and Torrente, P. (1994). El género de hongos liquenizados Lecanactis (Ascomycotina). Bibliotheca Lichenologica 54, 1–205.Google Scholar
Ekman, S. (1997). The genus Cliostomum revisited. Symbolae Botanicae Upsalienses, 32, 17–28.Google Scholar
Ekman, S. (2001). Molecular phylogeny of the Bacidiaceae (Lecanorales, lichenized Ascomycota). Mycological Research, 105, 783–797.CrossRefGoogle Scholar
Ekman, S. and Fröberg, S. (1988). Taxonomical problems in Aspicilia contorta and A. hoffmannii: an effect of hybridization?International Journal of Mycology and Lichenology, 3, 215–225.Google Scholar
Ekman, S. and Jørgensen, P. M. (2002). Towards a molecular phylogeny for the lichen family Pannariaceae (Lecanorales, Ascomycota). Canadian Journal of Botany, 80, 625–634.CrossRefGoogle Scholar
Ekman, S. and Tønsberg, T. (2002). Most species of Lepraria and Leproloma form a monophyletic group closely related to Stereocaulon. Mycological Research, 106, 1262–1276.CrossRefGoogle Scholar
Ekman, S. and Wedin, M. (2000). The phylogeny of the families Lecanoraceae and Bacidiaceae (lichenized Ascomycota) inferred from nuclear SSU rDNA sequences. Plant Biology, 2, 350–360.CrossRefGoogle Scholar
Elix, J. A. (1991). The lichen genus Relicina in Australasia. In Tropical Lichens: Their Systematics, Conservation and Ecology, ed. Galloway, D. J.. Systematics Association Special Volume 43, pp. 17–34. Oxford: Clarendon Press.Google Scholar
Elix, J. A. (1993). Progress in the generic delimitation of Parmelia sensu lato lichens (Ascomycotina: Parmeliaceae) and a synoptic key to the Parmeliaceae. Bryologist, 96, 359–383.CrossRefGoogle Scholar
Elix, J. A. (1994). Xanthoparmelia. Flora of Australia, 55, 201–308.
Elix, J. A. (1999). Detection and identification of secondary lichen substances: contributions by the Uppsala school of lichen chemistry. Symbolae Botanicae Upsalienses, 32, 103–121.Google Scholar
Elix, J. A. and Ernst-Russell, K. D. (1993). A Catalogue of Standardized Thin Layer Chromatographic Data and Biosynthetic Relationships for Lichen Substances, 2nd edn. Canberra: Australian National University.Google Scholar
Elix, J. A, Giralt, M. and Wardlaw, J. H. (2003). New chloro-depsides from the lichen Dimelaena radiata. Bibliotheca Lichenologica, 86, 1–7.Google Scholar
Elix, J. A., Johnston, J. and Parker, J. L. (1988). A computer program for the rapid identification of lichen substances. Mycotaxon, 31, 89–99.Google Scholar
Ellis, C. J. and Coppins, B. J. (2007). 19th century woodland structure controls stand-scale epiphyte diversity in present-day Scotland. Diversity and Distributions, 13, 84–91.CrossRefGoogle Scholar
Ellis, C. J., Crittenden, P. D., Scrimgeour, C. M. and Ashcroft, C. J. (2005). Translocation of 15N indicates nitrogen recycling in the mat-forming lichen Cladonia portentosa. New Phytologist, 168, 423–434.CrossRefGoogle ScholarPubMed
Elstner, E. F. and Oßwald, W. (1994). Mechanisms of oxygen activation during plant stress. Proceeding of the Royal Society of Edinburgh, 102, 131–154.Google Scholar
Elvebakk, A. and Bjerke, J. W. (2006 a). The Skibotn area in North Norway – an example of very high lichen species richness far to the north. Mycotaxon, 96, 141–146.Google Scholar
Elvebakk, A. and Bjerke, J. W. (2006b). The Skibotn area in North Norway – an example of very high lichen species richness far to the north: a supplement with an annotated list of species. Online: www.mycotaxon.com/resources/weblists.html.
Engelbert, R. and Vonarburg, C. (1995). Lichen diversity and ozone impact in rural areas of central Switzerland. Cryptogamic Botany, 5, 252–263.Google Scholar
Engelmann, M. D. (1966). Energetics, terrestrial field studies, and animal productivity. Advances in Ecological Research, 3, 73–115.CrossRefGoogle Scholar
Englund, B. (1977). The physiology of the lichen Peltigera aphthosa, with special reference to the blue-green phycobiont (Nostoc sp.). Physiologia Plantarum, 41, 298–304.CrossRefGoogle Scholar
Englund, B. (1978). Effects of environmental factors on acetylene reduction by intact thallus and excised cephalodia of Peltigera aphthosa Willd. Ecological Bulletin (Stockholm), 26, 234–246.Google Scholar
Englund, B. and Myerson, H. (1974). In situ measurement of nitrogen fixation at low temperatures. Oikos, 25, 183–187.CrossRefGoogle Scholar
Enriquez, S., Duarte, C. M., Sand-Jensen, K. and Nielsen, S. L. (1996). Broad-scale comparison of photosynthetic rates across phototrophic organisms. Oecologia, 108, 197–206.CrossRefGoogle ScholarPubMed
Eriksson, O. E. (2005). Ascomyceternas ursprung och evolution – Protolichenes-hypotesen. Svensk Mykologisk Tidskrift, 26, 22–29.Google Scholar
Eriksson, O. E. (ed.) (2006 a). Notes on ascomycete systematics Nr. 4324. Myconet, 12, 83–101. Online: www.fieldmuseum.org/myconet/Google Scholar
Eriksson, O. E. (ed.) (2006 b). Outline of Ascomycota, Myconet, 12, 1–82. Online: www.fieldmuseum.org/myconet/.Google Scholar
Ernst, A., Kirschenlohr, H., Diez, J. and Boger, P. (1984). Glycogen content and nitrogenase activity in Anabaena variabilis. Archiv für Microbiologie, 140, 120–125.CrossRefGoogle Scholar
Ertl, L. (1951). Über die Lichtverhältnisse in Laubflecten. Planta, 39, 245–270.CrossRefGoogle Scholar
Ertz, D., Christnach, C., Wedin, M. and Diederich, P. (2005). A world monograph of the genus Plectocarpon (Roccellaceae, Arthoniales). Bibliotheca Lichenologica, 91, 1–155.Google Scholar
Esseen, P.-A. and Renhorn, K. E. (1998). Mass loss of epiphytic lichen litter in a boreal forest. Annals Botanica Fennica, 35, 211–217.Google Scholar
Esseen, P.-A., Ehnström, B., Ericson, L. and Sjöberg, K. (1997). Boreal forests. Ecological Bulletins, 45, 16–47.Google Scholar
Esser, K. (1976). Kryptogamen: Blaualgen, Algen, Pilze, Flechten. Berlin: Springer.CrossRefGoogle Scholar
Esslinger, T. L. (1977). A chemosystematic revision of the brown Parmeliae. Journal of the Hattori Botanical Laboratory, 42, 1–211.Google Scholar
Esslinger, T. L. (1989). Systematics of Oropogon (Alectoriaceae) in the New World. Systematic Botany Monographs, 28, 1–111.CrossRefGoogle Scholar
Ettl, H. and Gärtner, G. (1984). Über die Bedeutung der Cytologie für die Algentaxonomie, dargestellt an Trebouxia (Chlorellales, Chlorophyceae). Plant Systematics and Evolution, 148, 135–147.CrossRefGoogle Scholar
Evans, C. A. and Hutchinson, T. C. (1996). Mercury accumulation in transplanted moss and lichens at high elevation sites in Quebec. Water, Air, and Soil Pollution, 90, 475–488.CrossRefGoogle Scholar
Evans, D. J. A., Archer, S. and Wilson, D. J. H. (1999). A comparison of the lichenometric and Schmidt hammer dating techniques based on data from the proglacial areas of some Icelandic glaciers. Quaternary Science Reviews, 18, 13–41.CrossRefGoogle Scholar
Evans, J. R. (1983). Nitrogen and photosynthesis in the flag leaf of wheat (Triticum aestivum L.). Plant Physiology, 72, 297–302.CrossRefGoogle Scholar
Evans, J. R. (1989). Photosynthesis and nitrogen relationships of C-3 plants. Oecologia, 78, 9–19.CrossRefGoogle Scholar
Evans, R. D. and Johansen, J. R. (1999). Microbiotic crusts and ecosystem processes. Critical Reviews in Plant Sciences, 18, 183–225.CrossRefGoogle Scholar
Evans, R. D., Belnap, J. Garcia-Pichel, F. and Phillips, S. L. (2001). Global change and the future of biological soil crusts. In Biological Soil Crusts: Structure, Function and Management, ed. Belnap, J. and Lange, O. L., pp. 417–429. Berlin: Springer.Google Scholar
Eversman, S. (1978). Effects of low-level SO2 on Usnea hirta and Parmelia chlorochroa. Bryologist, 81, 368–377.CrossRefGoogle Scholar
Eversman, S. and Sigal, L. L. (1987). Effects of SO2, O3, and SO2 and O3 in combination on photosynthesis and ultrastructure of two lichen species. Canadian Journal of Botany, 65, 1806–1818.CrossRefGoogle Scholar
Eversman, S., Johnson, C. and Gustafson, D. (1987). Vertical distribution of epiphytic lichens on three tree species in Yellowstone National Park. Bryologist, 90, 212–216.CrossRefGoogle Scholar
Fahselt, D. (1984). Interthalline variability in levels of lichen products within stands of Cladina stellaris. Bryologist, 87, 50–56.CrossRefGoogle Scholar
Fahselt, D. (1985). Multiple enzyme forms in lichens. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 129–143. New York: Plenum Publishing.CrossRefGoogle Scholar
Fahselt, D. (1987). Electrophoretic analysis of esterase and alkaline phosphatase enzyme forms in single spore cultures of Cladonia cristatella. Lichenologist, 19, 71–75.CrossRefGoogle Scholar
Fahselt, D. (1989). Enzyme polymorphism in sexual and asexual umbilicate lichens from Sverdrup Pass, Ellesmere Island, Canada. Lichenologist, 21, 279–285.CrossRefGoogle Scholar
Fahselt, D. (1991). Enzyme similarity as an indicator of evolutionary divergence: Stereocaulon saxatile H. Magn. Symbiosis, 11, 119–130.Google Scholar
Fahselt, D. (1992). Geothermal effects on multiple enzyme forms in the lichen Cladonia mitis. Lichenologist, 24, 181–192.Google Scholar
Fahselt, D. (1994 a). Secondary biochemistry of lichens. Symbiosis, 16, 117–165.Google Scholar
Fahselt, D. (1994 b). Carbon metabolism in lichens. Symbiosis, 17, 127–182.Google Scholar
Fahselt, D. (1995). Lichen sexuality from the perspective of multiple enzyme forms. Cryptogamic Botany, 5, 137–143.Google Scholar
Fahselt, D. (2001). Analysing lichen enzymes by isoelectricfocussing. In Protocols in Lichenology, ed. Kranner, I., Varma, A. and Beckett, P., pp. 307–331. Berlin: Springer.Google Scholar
Fahselt, D. and Alstrup, V. (1997). High performance liquid chromatography of phenolics in recent and subfossil lichens. Canadian Journal Botany, 75, 1148–1154.CrossRefGoogle Scholar
Fahselt, D. and Hageman, C. (1994). Rhizine and upper thallus isozymes in umbilicate lichens. Symbiosis, 16, 95–103.Google Scholar
Fahselt, D. and Krol, M. (1989). Biochemical comparison of two ecologically distinctive forms of Xanthoria elegans in the Canadian High Arctic. Lichenologist, 21, 135–145.CrossRefGoogle Scholar
Fahselt, D., Krol, M., Hüner, N. and Tønsberg, T. (2000). Pigmentation of Cladonia infected by the lichenicolous fungus Arthrorhaphis aeroguinosa. Lichenologist, 32, 300–303.CrossRefGoogle Scholar
Fahselt, D., Madzia, S. and Alstrup, V. (2001). Scanning electron microscopy of invasive fungi in lichens. Bryologist, 104, 24–39.CrossRefGoogle Scholar
Fahselt, D., Tavares, S. and Mazdia, S. (1997). Isozyme variation in lichens in relation to mine dust exposure. In Progress and Problems in Lichenology in the Nineties, ed. Türk, R. and Zorer, R.. Bibliotheca Lichenologica, 68, 111–127.Google Scholar
Farkas, E. E. and Sipman, H. J. M. (1993). Bibliography and checklist of foliicolous lichenized fungi up to 1992. Tropical Bryology, 7, 93–148.Google Scholar
Farkas, E. and Sipman, H. J. M. (1997). Checklist of foliicolous lichenized fungi – after Farkas and Sipman (1993), with additions to 1996. Abstracta Botanica, 21, 173–206.Google Scholar
Farrar, J. F. (1976 a) Ecological physiology of the lichen Hypogymnia physodes. I. Some effects of constant water saturation. New Phytologist, 77, 93–103.CrossRefGoogle Scholar
Farrar, J. F. (1976 b). Ecological physiology of the lichen Hypogymnia physodes. II. Effects of wetting and drying cycles and the concept of physiological buffering. New Phytologist, 77, 105–113.CrossRefGoogle Scholar
Farrar, J. F. (1976 c). The lichen as an ecosystem: observation and experiment. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 385–406. London: Academic Press.Google Scholar
Farrar, J. F. (1976 d). The uptake and metabolism of phosphate by the lichen Hypogymnia physodes. New Phytologist, 77, 127–134.CrossRefGoogle Scholar
Farrar, J. F. (1988). Physiological buffering. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 101–105. Boca Raton: CRC Press.Google Scholar
Farrar, J. F. and Smith, D. C. (1976). Ecological physiology of the lichen Hypogymnia physodes. III. The importance of the rewetting phase. New Phytologist, 77, 115–125.CrossRefGoogle Scholar
Feige, G. B. (1973). Untersuchungen zur Ökologie und Physiologie der marinen Blaualgenflechte Lichina pygmaea Ag. II. Die Reversibilität der Osmoregulation. Zeitschrift für Pflanzenphyiologie, 68, 415–421.CrossRefGoogle Scholar
Feige, G. B. and Jensen, M. (1992). Basic carbon and nitrogen metabolism of lichens. In Algae and Symbioses, ed. Reisser, W., pp. 277–299. Bristol: Biopress Limited.Google Scholar
Feige, G. B., Lumbsch, H. T., Huneck, S. and Elix, J. A. (1993). The identification of lichen substances by a standardized high-performance liquid chromatographic method. Journal of Chromatography, 646, 417–427.CrossRefGoogle Scholar
Fenn, M. E., Baron, J. S., Allen, E. B., et al. (2003). Ecological effects of nitrogen deposition in the western United States. BioScience, 53, 404–420.CrossRefGoogle Scholar
Ferber, T. (2002). The age and origin of talus cones in the light of lichenometric research. The Skalnisty and Zielony talus cones, High Tatra Mountains, Poland. Studia Geomorphologica Carpatho-Balcanica, 36, 77–90.Google Scholar
Ferraro, L. I., Lücking, R. and Sérusiaux, E. (2001). A world monograph of the lichen genus Gyalectidium (Gomphillaceae). Botanical Journal of the Linnean Society, 137, 311–345.CrossRefGoogle Scholar
Ferry, B. W. and Baddeley, M. S. (1976). Sulphur dioxide uptake in lichens. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 407–418. London: Academic Press.Google Scholar
Feuerer, T. (ed.) (2006). Checklists of lichens and lichenicolous fungi. Version 1 June 2006. Online: www.checklists.de.
Fewer, D., Friedl, T. and Büdel, B. (2002). Chroococcidiopsis and heterocyst-differentiating cyanobacteria are each other's closest living relatives. Molecular Phylogenetics and Evolution, 41, 498–506.Google Scholar
Fiechter, E. (1990). Thallusdifferenzierung und intrathalline Sekundärstoffverteilung bei Parmeliaceae (Lecanorales, lichenisierte Ascomyceten). Inauguraldissertation. Zürich: Universität Zürich.
Fields, R. D. (1988). Physiological responses of lichens to air pollutant fumigations. In Lichens, Bryophytes and Air Quality, ed. Nash, T. H. III and Wirth, V., pp. 175–200. Bibliotheca Lichenologica 30. Berlin-Stuttgart: J. Cramer.Google Scholar
Fields, R. D. and St. Clair, L. L. (1984). The effects of SO2 on photosynthesis and carbohydrate transfer in the two lichens: Collema polycarpon and Parmelia chlorochroa. American Journal of Botany, 71, 986–998.CrossRefGoogle Scholar
Fletcher, A. (1980). Marine and maritime lichens of rocky shores: their ecology, physiology and biological interactions. In The Shore Environment, Vol. 2: Ecosystems, ed. Price, J. H., Irvine, D. E. G. and Farnham, W. F., pp. 789–842. London: Academic Press.Google Scholar
Fletcher, J. (2002). Coordination of cell proliferation and cell fate decisions in the angiosperm shoot apical meristem. BioEssays, 24, 27–37.CrossRefGoogle ScholarPubMed
Fogg, G. E., Fay, P. and Walsby, A. E. (1973). The Bluegreen Algae. London: Academic Press.Google Scholar
Follmann, G. (2002). South America as diversity centre of the lichen family Roccellaceae (Arthoniales). Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg, 30–32, 61–77.Google Scholar
Forman, R. T. T. (1975). Canopy lichens with blue-green algae: a nitrogen source in a Columbian rain forest. Ecology, 56, 1176–1184.CrossRefGoogle Scholar
Forman, R. T. T. and Dowden, D. L. (1977). Nitrogen fixing lichen roles from desert to alpine in the Sangre de Cristo Mountains, New Mexico. Bryologist, 80, 561–70.CrossRefGoogle Scholar
Foyer, C. and Halliwell, B. (1976). The presence of glutathione and glutathione reductase in chloroplasts: a proposed role in ascorbic acid metabolism. Planta, 133, 21–25.CrossRefGoogle ScholarPubMed
Frank, H. A., Young, A., Britton, G. and Cogdell, R. J. (1999). The Photochemistry of Carotenoids. Berlin: Springer.Google Scholar
Freedman, B., Zobens, V., Hutchinson, T. C. and Gizyn, W. I. (1990). Intense, natural pollution affects Arctic tundra vegetation at the Smoking Hills, Canada. Ecology, 71, 492–503.CrossRefGoogle Scholar
Fremstad, E., Paal, J. and Möls, T. (2005). Impacts of increased nitrogen supply on Norwegian lichen-rich alpine communities: a 10-year experiment. Journal of Ecology, 93, 471–481.CrossRefGoogle Scholar
Friedl, T. (1987). Thallus development and phycobionts of the parasitic lichen Diploschistes muscorum. Lichenologist, 19, 183–191.CrossRefGoogle Scholar
Friedl, T. (1989 a). Comparative ultrastructure of pyrenoids in Trebouxia (Microthamniales, Chlorophyta). Plant Systematics and Evolution, 164, 145–159.CrossRefGoogle Scholar
Friedl, T. (1989b). Systematik und Biologie von Trebouxia (Microthamniales, Chlorophyta) als Phycobiont der Parmeliaceae (lichenisierte Ascomyceten). Ph.D. thesis. Bayreuth: Universtät Bayreuth.
Friedl, T. (1993). New aspects of the reproduction by autospores in the lichen alga Trebouxia (Microthamniales, Chlorophyta). Archiv für Protistenkunde, 143, 153–161.CrossRefGoogle Scholar
Friedl, T. and Gärtner, G. (1988). Trebouxia (Pleurastrales, Chlorophyta) as a phycobiont in the lichen genus Diploschistes. Archiv für Protistenkunde, 135, 147–158.CrossRefGoogle Scholar
Friedl, T. and Zeltner, C. (1994). Assessing the relationships of some coccoid green lichen algae and the Microthamniales (Chlorophyta) with 18S rRNA gene sequence comparisons. Journal of Phycology, 30, 500–506.CrossRefGoogle Scholar
Friedmann, E. I. (1982). Endolithic microorganisms in the Antarctic cold desert. Science, 215, 1045–1053.CrossRefGoogle ScholarPubMed
Friedmann, E. I. and Sun, H. J. (2005). Communities adjust their temperature optima by shifting producer-to-consumer ratio, shown in lichens as models. I. Hypothesis. Microbial Ecology, 49, 523–527.CrossRefGoogle ScholarPubMed
Fritz-Sheridan, R. P. (1985). Impact of simulated acid rain on nitrogenase activity in Peltigera aphthosa and P. polydactyla. Lichenologist, 17, 27–31.CrossRefGoogle Scholar
Fritz-Sheridan, R. P. and Coxson, D. S. (1988 a). Nitrogen fixation on the tropical volcano, La Soufrière (Guadeloupe): nitrogen fixation, photosynthesis and respiration during the prevailing cloud/shroud climate by Stereocaulon virgatum. Lichenologist, 20, 41–61.CrossRefGoogle Scholar
Fritz-Sheridan, R. P. and Coxson, D. S. (1988 b). Nitrogen fixation on the tropical volcano, La Soufrière (Guadeloupe): the interaction of temperature, moisture, and light with net photosynthesis and nitrogenase activity in Stereocaulon virgatum and response to periods of insolation shock. Lichenologist, 20, 63–81.CrossRefGoogle Scholar
Fröberg, L., Berg, C. O., Baur, A. and Baur, B. (2001). Viability of lichen photobionts after passing through the digestive tract of a land snail. Lichenologist, 33, 543–545.CrossRefGoogle Scholar
Fujii, I., Watanabe, A., Sankawa, U. and Ebizuka, Y. (2001). Identification of Claisen cyclase domain in fungal polyketide synthase WA, a naphthapyrone synthase of Aspergillus nidulans. Chemical Biology, 8, 187–197.CrossRefGoogle Scholar
Gagunashvili, A. and Andrésson, O. (2006). Heterologous expression of a lichen polyketide synthase in filamentous fungi. EUKETIDES Meeting, Turku, Finland.
Gaio-Oliveira, G., Dahlman, L., Máguas, C. and Palmqvist, K. (2004 a). Growth in relation to microclimatic conditions and physiological characteristics of four Lobaria pulmonaria populations in two contrasting habitats. Ecography, 27, 13–28.CrossRefGoogle Scholar
Gaio-Oliveira, G., Dahlman, L., Martins-Loução, M. A., Máguas, C. and Palmqvist, K. (2005 a). Nitrogen uptake in relation to excess supply and its effects on the lichens Evernia prunastri (L.) Ach and Xanthoria parietina (L.) Th. Fr. Planta, 220, 794–803.CrossRefGoogle Scholar
Gaio-Oliveira, G., Dahlman, L., Palmqvist, K. and Máguas, C. (2004 b). Ammonium uptake in the nitrophytic lichen Xanthoria parietina and its effects on vitality and balance between symbionts. Lichenologist, 36, 75–86.CrossRefGoogle Scholar
Gaio-Oliveira, G., Dahlman, L., Palmqvist, K. and Máguas, C. (2005 b). Responses of the lichen Xanthoria parietina (L.) Th. Fr. to varying thallus nitrogen concentrations. Lichenologist, 37, 171–179.CrossRefGoogle Scholar
Gaio-Oliveira, G., Moen, J., Danell, O. and Palmqvist, K. (2006). Effect of simulated reindeer grazing on the re-growth capacity of mat-forming lichens. Basic and Applied Ecology, 7, 109–121.CrossRefGoogle Scholar
Galley, C. and Linder, H. P. (2006). Geographical affinities of the Cape flora, South Africa. Journal of Biogeography, 33, 236–250.CrossRefGoogle Scholar
Galloway, D. J. (1991 a). Chemical evolution in the order Peltigerales: triterpenoids. Symbiosis, 11, 327–344.Google Scholar
Galloway, D. J. (1991 b). Phytogeography of Southern Hemisphere lichens. In Quantitative Approaches to Phytogeography, ed. Nimis, P. L. and Crovello, T. J., pp. 233–262. Dordrecht: Kluwer Academic.Google Scholar
Galloway, D. J. (1993). Global environmental change: lichens and chemistry. Bibliotheca Lichenologica, 53, 87–95.Google Scholar
Galloway, D. J. (1994 a). Pseudocyphellaria lacerata new to the Faeroe Islands. Lichenologist, 26, 391–393.CrossRefGoogle Scholar
Galloway, D. J. (1994 b). Studies on the lichen genus Sticta (Schreber) Ach. I. Southern South American species. Lichenologist, 26, 223–282.CrossRefGoogle Scholar
Galloway, D. J. (1995 a). The extra-European lichen collections of Archibald Menzies MD, FLS (1754–1842). Edinburgh Journal of Botany, 52, 95–139.CrossRefGoogle Scholar
Galloway, D. J. (1995 b). Studies on the lichen genus Sticta (Schreber) Ach. III. Notes on species described by Bory de St-Vincent, William Hooker, and Delise, between 1804 and 1825. Nova Hedwigia, 61, 147–188.Google Scholar
Galloway, D. J. (1996). Lichen biogeography. In Lichen Biology, ed. Nash, T. H. III, pp. 199–216. Cambridge: Cambridge University Press.Google Scholar
Galloway, D. J. (1997). Studies on the lichen genus Sticta (Schreber) Ach. IV. New Zealand species. Lichenologist, 29, 105–168.CrossRefGoogle Scholar
Galloway, D. J. (1999). Notes on the lichen genus Leptogium (Collemataceae, Ascomycota) in New Zealand. Nova Hedwigia, 69, 317–355.Google Scholar
Galloway, D. J. (2001). Sticta. Flora of Australia, 58A, 78–97.
Galloway, D. J. (2002). Taxonomic notes on the lichen genus Placopsis (Agyriaceae: Ascomycotina) on southern South America, with a key to species. Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg 30–32, 70–107.Google Scholar
Galloway, D. J. (2003). Additional lichen records from New Zealand 40. Buellia aethalea (Ach.) Th. Fr., Catillaria contristans (Nyl.) Zahlbr., Frutidella caesioatra (Schaer.) Kalb, Placynthium rosulans (Th.Fr.) Zahlbr. and Pseudocyphellaria mallota. Australasian Lichenology, 53, 20–29.Google Scholar
Galloway, D. J. (2004 a). Placopsis hertelii (Agyriaceae, Ascomycota) endemic to New Zealand, with descriptions of four additional new species of Placopsis (Nyl.) Linds, from New Zealand. Bibliotheca Lichenologica, 88, 147–161.Google Scholar
Galloway, D. J. (2004 b). New lichen taxa and names in the New Zealand mycobiota. New Zealand Journal of Botany, 42, 105–120.CrossRefGoogle Scholar
Galloway, D. J. (2007). Flora of New Zealand Lichens, Including Lichen-forming and Lichenicolous Fungi, 2nd edn. Lincoln: Manaaki Whenua.Google Scholar
Galloway, D. J. (2008). Southern Hemisphere lichens. New Zealand Journal of Botany (In review,)Google Scholar
Galloway, D. J. and Aptroot, A. (1995). Bipolar lichens: a review. Cryptogamic Botany, 5, 184–191.Google Scholar
Galloway, D. J. and Quilhot, W. (1998) [“1999”]. Checklist of Chilean lichen-forming and lichenicolous fungi. Gayana (Botanica), 55, 111–185.Google Scholar
Galloway, D. J. and Thomas, M. A. (2004). Sticta. In Lichen Flora of the Greater Sonoran Desert Region, Vol. 2, ed. Nash, T. H. III, Ryan, B. D., Diederich, P., Gries, C. and Bungartz, F., pp. 513–524. Tempe: Lichens Unlimited.Google Scholar
Galloway, D. J., Hafellner, J. and Elix, J. A. (2005). Stirtoniella, a new genus for Catillaria kelica (Lecanorales: Ramalinaceae). Lichenologist, 37, 261–271.CrossRefGoogle Scholar
Galloway, D. J., Kantvilas, G. and Elix, J. A. (2001). Pseudocyphellaria. Flora of Australia, 58A, 47–77.
Galun, M. (1988 a). Handbook of Lichenology, Vols. 1, 2 and 3, Boca Raton: CRC Press.Google Scholar
Galun, M. (1988 b). Carbon metabolism. In Handbook of Lichenology, Vol. I, ed. Galun, M., pp. 147–156. Boca Raton: CRC Press.Google Scholar
Galun, M. and Bubrick, P. (1984). Physiological interactions between the partners of the lichen symbiosis. In Cellular Interactions: Encyclopedia of Plant Physiology, ed. Linskens, H. F. and Heslop-Harrison, J., pp. 362–401. Berlin: Springer.CrossRefGoogle Scholar
Galun, M. and Mukhtar, A. (1996). Checklist of the lichens of Israel. Bocconea, 6, 149–171.Google Scholar
Galun, M. and Ronen, R. (1988). Interactions of lichens and pollutants. In CRC Handbook of Lichenology, Vol. III, ed. Galun, M., pp. 55–72. Boca Raton: CRC Press.Google Scholar
Galun, M. and Shomer-Ilan, A. (1988). Secondary metabolic products. In CRC Handbook of Lichenology, Vol. III, ed. Galun, M., pp. 3–8. Boca Raton: CRC Press.Google Scholar
Ganderton, P. and Coker, P. (2005). Environmental Biogeography. Harlow, Essex: Pearson Education Ltd.Google Scholar
Garbaye, J. (1994). Helper bacteria: a new dimension to the mycorrhizal symbiosis. New Phytologist, 128, 197–210.CrossRefGoogle Scholar
Garcia-Molina, F., Hiner, A. N., Fenoll, L. G., et al. (2005). Mushroom tyrosinase: catalase activity, inhibition, and suicide inactivation. Journal of Agricultural and Food Chemistry, 53, 3702–3709.CrossRefGoogle ScholarPubMed
Gargas, A., DePriest, P. T., Grube, M. and Tehler, A. (1995). Multiple origins of lichen symbiosis in fungi suggested by SSU rDNA phylogeny. Science, 268, 1492–1495.CrossRefGoogle ScholarPubMed
Garrett, R. M. (1972). Electrostatic charges on freshly discharged lichen ascospores. Lichenologist, 5, 311–313.CrossRefGoogle Scholar
Gärtner, G. (1985). Die Gattung Trebouxia PUYMALY (Chlorellales, Chlorophyceae). Archiv für Hydrobiologie, Supplementband, 74, (Algological Studies, 41), 495–548.Google Scholar
Gärtner, G. (1992). Taxonomy of symbiotic eukaryotic algae. In Algae and Symbioses, ed. Reisser, W., pp. 325–338. Bristol: Biopress Limited.Google Scholar
Garty, J. (2001). Biomonitoring atmospheric heavy metals with lichens: theory and application. Critical Review in Plant Sciences, 20, 309–371.CrossRefGoogle Scholar
Garty, J., Cohen, Y., Kloog, N. and Karnieli, A. (1997). Effect of air pollution on cell membrane integrity, spectral reflectance and metal and sulfur concentrations in lichens. Environmental Toxicology and Chemistry, 16, 1396–1402.CrossRefGoogle Scholar
Garty, J., Galun, M. and Kessel, M. (1979). Localization of heavy metals and other elements accumulated in the lichen thallus. New Phytologist, 82, 159–168.CrossRefGoogle Scholar
Garty, J.Karary, Y., Harel, J. and Lurie, S. (1993). The impact of air pollution on the integrity of cell membranes and chlorophyll in the lichen Ramalina duriaei (De Not.) Bagl. transplanted to industrial sites in Israel. Archives of Environmental Contamination and Toxicology, 24, 455–460.CrossRefGoogle Scholar
Garty, J., Perry, A. S. and Mozel, J. (1982). Accumulation of polychlorinated biphenyls (PCBs) in the transplanted lichen Ramalina duriaei in air quality biomonitoring experiments. Nordic Journal of Botany, 2, 583–586.CrossRefGoogle Scholar
Gaugh, H. G. Jr. (1982). Multivariate Analysis in Community Ecology. Cambridge: Cambridge University Press.Google Scholar
Gauslaa, Y. (2006). Trade-off between reproduction and growth in the foliose old forest lichen Lobaria pulmonaria. Basic and Applied Ecology, 7, 455–460.CrossRefGoogle Scholar
Gauslaa, Y. and McEvoy, M. (2005). Seasonal changes in solar radiation drive acclimation of the sun-screening compound parietin in the lichen Xanthoria parietina. Basic and Applied Ecology, 6, 75–82.CrossRefGoogle Scholar
Gauslaa, Y. and Solhaug, K. A. (1998). The significance of thallus size for the water economy of the cyanobacterial old-forest lichen Degelia plumbea. Oecologia, 116, 76–84.CrossRefGoogle ScholarPubMed
Gauslaa, Y. and Solhaug, K. A. (1999). High-light damage in air-dry thalli of the old forest lichen Lobaria pulmonaria – interactions of irradiance, exposure duration and high temperature. Journal of Experimental Botany, 50, 697–705.Google Scholar
Gauslaa, Y. and Solhaug, K. A. (2001). Fungal melanins as a sun screen for symbiotic green algae in the lichen Lobaria pulmonaria. Oecologia, 126, 462–471.CrossRefGoogle ScholarPubMed
Gauslaa, Y. and Ustvedt, E. M. (2003). Is parietin a UV-B or a blue-light screening pigment in the lichen Xanthoria parietina?Photochemical and Photobiological Sciences, 2, 424–432.CrossRefGoogle Scholar
Gauslaa, Y., Holien, H., Ohlson, M. and Solhøy, T. (2006 a). Does snail grazing affect growth of the old forest lichen Lobaria pulmonaria?Lichenologist, 38, 587–593.CrossRefGoogle Scholar
Gauslaa, Y., Lie, M., Solhaug, K. and Ohlson, M. (2006 b). Growth and ecophysiological acclimation of the foliose lichen Lobaria pulmonaria in forests with contrasting light climates. Oecologia, 147, 406–416.CrossRefGoogle ScholarPubMed
Gauslaa, Y., Ohlson, M., Solhaug, K. A., Bilger, W. and Nybakken, L. (2001). Aspect dependent high-irradiance damage to two transplanted foliose forest lichens Lobaria pulmonaria and Parmelia sulcata. Canadian Journal of Forest Research, 31, 1639–1649.CrossRefGoogle Scholar
Gaya, E., Lutzoni, F., Zoller, S. and Navarro-Rosinés, P. (2003). Phylogenetic study of Fulgensia and allied Caloplaca and Xanthoria species (Teloschistaceae, lichen-forming Ascomycota). American Journal of Botany, 90, 1095–1103.CrossRefGoogle Scholar
Geebelen, W. and Hoffmann, M. (2001). Evaluation of bio-indication methods using epiphytes by correlating with SO2-pollution parameters. Lichenologist, 33, 249–260.CrossRefGoogle Scholar
Gehrig, H., Schüssler, A. and Kluge, M. (1996). Geosiphon pyriforme, a fungus forming endocytobiosis with Nostoc (cyanobacteria), is an ancestral member of the Glomales: evidence by SSU rRNA analysis. Journal of Molecular Evolution, 43, 71–81.CrossRefGoogle ScholarPubMed
Geitler, L. (1932). Cyanophyceae von Europa unter Berücksichtigung der anderen Kontinente. In Rabenhorst's Kryptogamenflora von Deutschland, Österreich und der Schweiz, 2nd edn., Vol. 14, ed. Kolkwitz, R., pp. 1–1196. Leipzig: Akademische Verlagsgesellschaft.Google Scholar
Geitler, L. (1934). Beiträge zur Kenntnis der Flechtensymbiose. IV, V. Archiv für Protistenkunde 82, 51–85.Google Scholar
Geitler, L. (1937). Beiträge zur Kenntnis der Flechtensymbiose. VI. Die Verbindung von Pilz und Alge bei den Pyrenopsidaceen Synalissa, Thyrea, Peccania und Psorotichia. Archiv für Protistenkunde, 88, 161–179.Google Scholar
Gerson, U. and Seaward, M. R. D. (1977). Lichen-invertebrate associations. In Lichen Ecology, ed. Seaward, M. R. D., pp. 69–119. London: Academic Press.Google Scholar
Geyer, M., Feuerer, T. and Feige, G. B. (1984). Chemie und Systematik in der Flechtengattung Rhizocarpon: Hochdruckflüssigkeitschromatographie (HPLC) der Flechten-Sekundärstoffe der Rhizocarpon superficiale-Gruppe. Plant Systematics and Evolution, 145, 41–54.CrossRefGoogle Scholar
Gignac, D. and Dale, M. R. T. (2005). Effects of fragment size and habitat heterogeneity on cryptogam diversity in the low-boreal forest of western Canada. Bryologist, 108, 50–66.CrossRefGoogle Scholar
Gilbert, O. L. (1970). Further studies on the effect of sulphur dioxide on lichens and bryophytes. New Phytologist, 69, 605–627.CrossRefGoogle Scholar
Gilbert, O. L. (1971). The effect of airborne fluorides on lichens. Lichenologist, 5, 26–32.CrossRefGoogle Scholar
Gilbert, O. L. (1985). Environmental effects of airborne fluorides from aluminium smelting at Invergordon, Scotland 1971–1983. Environmental Pollution, Series A, 39, 293–302.CrossRefGoogle Scholar
Gilbert, O. L. (1992). Lichen reinvasion with declining air pollution. In Bryophytes and Lichens in a Changing Environment, ed. Bates, J. W. and Farmer, A. M., pp. 159–177. Oxford: Clarendon Press.Google Scholar
Gillespie, J. H. (1991). The Causes of Molecular Evolution. Oxford: Oxford University Press.Google Scholar
Gjerde, I., Sætersdal, M., Rolstad, J., et al. (2005). Productivity–diversity relationships for plants, bryophytes, lichens and polypore fungi in six northern forest landscapes. Ecography, 28, 705–720.CrossRefGoogle Scholar
Gob, F., Oetit, F., Bravard, J. P., Ozer, A. and Gob, A. (2003). Lichenometric application to historical and subrecent dynamics and sediment transport of a Corsican stream (Figarella River, France). Quaternary Science Reviews, 22, 2111–2124.CrossRefGoogle Scholar
Goebel, K. (1926). Die Wasseraufnahme der Flechten. Berichte der deutschen botanischen Gesellschaft, 44, 158–161.Google Scholar
Goffinet, B. and Bayer, R. J. (1997). Characterization of mycobionts of photomorph pairs in the Peltigerineae (lichenized ascomycetes) based on internal transcribed spacer sequences of the nuclear ribosomal DNA. Fungal Genetics and Biology, 21, 228–237.CrossRefGoogle ScholarPubMed
Goldner, W. R., Hoffman, F. M. and Medve, R. J. (1986). Allelopathic effects of Cladonia cristatella on ectomycorrhizal fungi common to bituminous strip-mine spoils. Canadian Journal of Botany, 64, 1586–1590.CrossRefGoogle Scholar
Golm, G. T., Hill, P. S. and Wells, H. (1993). Life expectancy in a Tulsa cemetery: growth and population structure of the lichen Xanthoparmelia cumberlandia. American Midland Naturalist, 129, 373–383.CrossRefGoogle Scholar
Gombert, S., Asta, J. and Seaward, M. R. D. (2003). Correlation between the nitrogen concentration of two epiphytic lichens and the traffic density in an urban area. Environmental Pollution, 123, 281–290.CrossRefGoogle Scholar
Gordy, V. R. and Hendrix, D. L. (1982). Respiratory response of the lichens Ramalina stenospora Mull. Arg. and Ramalina complanata (Sw.) Ach. to azide, cyanide, salicylhydroxamic acid and bisulfate during thallus hydration. Bryologist, 85, 361–374.CrossRefGoogle Scholar
Gorin, P. A. J., Baron, M. and Iacomini, M. (1988). Storage products of lichens. In CRC Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 9–24. Boca Raton: CRC Press.Google Scholar
Gough, L. P., Severson, R. C. and Jackson, L. L. (1988). Determining baseline element composition of lichens. I. Parmelia sulcata at Theodore Roosevelt National Park, North Dakota. Water, Air, and Soil Pollution, 38, 157–167.Google Scholar
Goward, T. (1994). Notes on old growth-dependent epiphytic macrolichens in inland British Columbia, Canada. Acta Botanica Fennica, 150, 31–38.Google Scholar
Goward, T. (1995). Nephroma occultum and the maintenance of lichen diversity in British Columbia. Mitteilungen der Eidgenössischen Forschungsanstalt für Wald, Schnee und Landschaft, 70, 93–101.Google Scholar
Goward, T. (1996). Lichens of British Columbia: Rare Species and Priorities for Inventory. Victoria: Province of British Columbia, Ministry of Forests Research Program. Working Paper 08/1996, pp. i–viii + 1–34.
Goward, T. and Ahti, T. (1997). Notes on the distributional ecology of the Cladoniaceae (lichenized ascomycetes) in temperate and boreal western North America. Journal of the Hattori Botanical Laboratory, 82, 143–155.Google Scholar
Goyal, R. and Seaward, M. R. D. (1982). Metal uptake in terricolous lichens. III. Translocation in the thallus of Peltigera canina. New Phytologist, 90, 85–98.CrossRefGoogle Scholar
Grace, B., Gillespie, T. J. and Puckett, K. J. (1985 a). Sulphur dioxide threshold concentration values for Cladina rangiferina in the Mackenzie Valley, N. W. T. Canadian Journal of Botany, 63, 806–812.CrossRefGoogle Scholar
Grace, B., Gillespie, T. J. and Puckett, K. J. (1985 b). Uptake of gaseous sulphur dioxide by the lichen Cladina rangiferina. Canadian Journal of Botany, 63, 797–805.CrossRefGoogle Scholar
Grace, J. (1997). Toward models of resource allocation by plants. In Plant Resource Allocation, ed. Bazzaz, F. A. and Grace, J., pp. 279–291. San Diego: Academic Press.Google Scholar
Gradstein, S. and Lücking, R. (1997). Synthesis of the Symposium (on Foliicolous Cryptogams) and priorities for future research. Abstracta Botanica, 21, 207–214.Google Scholar
Grant, B. S., Owen, D. F. and Clarke, C. A. (1996). Parallel rise and fall of melanic peppered moths in America and Britain. Journal of Heredity, 87, 351–357.CrossRefGoogle Scholar
Green, T. G. A. and Lange, O. L. (1995). Photosynthesis in poikilohydric plants: a comparison of lichens and bryophytes. In Ecophysiology of Photosynthesis, ed. Schulze, E. D. and Caldwell, M. M., pp. 71–101. Berlin: Springer.Google Scholar
Green, T. G. A., Büdel, B., Heber, U., et al. (1993). Differences in photosynthetic performance between cyanobacterial and green algal components of lichen photosymbiodemes measured in the field. New Phytologist, 125, 723–731.CrossRefGoogle Scholar
Green, T. G. A, Büdel, B., Meyer, A., Zellner, H. and Lange, O. L. (1997). Temperate rainforest lichens in New Zealand: light response of photosynthesis. New Zealand Journal of Botany, 35, 493–504.CrossRefGoogle Scholar
Green, T. G. A., Horstmann, J., Bonnett, H., Wilkins, A. and Silvester, W. B. (1980). Nitrogen fixation by members of the Stictaceae (Lichens) of New Zealand. New Phytologist, 84, 339–348.CrossRefGoogle Scholar
Green, T. G. A., Meyer, A., Büdel, B., Zellner, H. and Lange, O. L. (1995). Diel patterns of CO2-exchange for six lichens from a temperate rain forest in New Zealand. Symbiosis, 18, 251–273.Google Scholar
Green, T. G. A., Schlensog, M., Sancho, L. G., et al. (2002). The photobiont (cyanobacterial or green algal) determines the pattern of photosynthetic activity within a lichen photosymbiodeme: evidence obtained from in situ measurements of chlorophyll a fluorescence. Oecologia, 130, 191–198.CrossRefGoogle Scholar
Green, T. G. A., Schroeter, B., Kappen, L., Seppelt, R. D. and Maseyk, K. (1998). An assessment of the relationship between chlorophyll a fluorescence and CO2 gas exchange from field measurements on a moss and lichen. Planta, 206, 611–618.CrossRefGoogle Scholar
Green, T. G. A., Schroeter, B. and Sancho, L. G. (1999). Plant life in Antarctica. In Handbook of Functional Plant Ecology, ed. Pugnaire, F. I. and Valladares, F., pp. 495–543. New York: Marcel Dekker, Inc.Google Scholar
Green, T. G. A., Schroeter, B. and Sancho, L. G. (2007). Plant life in Antarctica. In Functional Plant Ecology, 2nd edn., ed. Pugnaire, F. I. and Valladares, F., pp. 389–433. New York: Marcel Dekker.Google Scholar
Green, T. G. A., Snelgar, W. P. and Wilkins, A. L. (1985). Photosynthesis, water relations and thallus structure of Stictaceae lichens. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 57–75. New York: Plenum Press.CrossRefGoogle Scholar
Gregory, K. J. (1975). Lichens and the determination of river channel capacity. Earth Surface Processes, 1, 273–285.CrossRefGoogle Scholar
Grehan, J. R. (2001). Panbiogeography from tracks to ocean basins: evolving perspectives. Journal of Biogeography, 28, 413–429.CrossRefGoogle Scholar
Gries, C., Nash, T. H. III and Kesselmeier, J. (1994). Exchange of reduced sulfur gases between lichens and the atmosphere. Biogeochemistry, 23, 25–39.CrossRefGoogle Scholar
Gries, C., Romagni, J. G., Nash, T. H. III, Kuhn, U. and Kesselmeier, J. (1997 a). The relation of H2S release to SO2. New Phytologist, 136, 703–711.CrossRefGoogle Scholar
Gries, C., Sanz, M.-J. and Nash, T. H. III (1995). The effect of SO2 fumigation on CO2 gas exchange, chlorophyll fluorescence and chlorophyll degradation in different lichen species from western North America. Cryptogamic Botany, 5, 239–246.Google Scholar
Gries, C., Sanz, M.-J., Romagni, J. G., et al. (1997 b). The uptake of gaseous sulphur dioxide by non-gelatinous lichens. New Phytologist, 135, 595–602.CrossRefGoogle Scholar
Griffith, M. and Yaish, M. W. F. (2004). Antifreeze proteins in overwintering plants: a tale of two activities. Trends in Plant Science, 9, 399–405.CrossRefGoogle ScholarPubMed
Grime, J. P. (1979). Plant Strategies and Vegetation Processes. Chichester: Wiley.Google Scholar
Grime, J. P., Hodgson, J. P. and Hunt, R. (1988). Comparative Plant Ecology: a Functional Approach to Common British Species. London: Unwin Hyman.CrossRefGoogle Scholar
Grodzinska, K., Godzik, B. and Bienkowski, P. (1999). Cladina stellaris (Opiz) Brodo as a bioindicator of atmospheric deposition on the Kola Peninsula, Russia. Polar Research, 18, 105–110.CrossRefGoogle Scholar
Groombridge, B., ed. (1992). Global Biodiversity: Status of the Earth's Living Resources. London: Chapman and Hall.CrossRefGoogle Scholar
Grube, M. (1998). Classification and phylogeny in the Arthoniales (lichenized Ascomycetes). Bryologist, 101, 377–391.CrossRefGoogle Scholar
Grube, M. and Blaha, J. (2003). On the pylogeny of some polyketide synthase genes in the lichenized genus Lecanora. Mycological Research, 107, 1419–1426.CrossRefGoogle Scholar
Grube, M. and Los Ríos, A. (2001). Observations on Biatoropsis usnearum, a lichenicolous heterobasidiomycete, and other gall-forming fungi, using different microscopical techniques. Mycological Research, 105, 1116–1122.CrossRefGoogle Scholar
Grube, M. and Hafellner, F. (1990). Studien an flechtenbewohnenden Pilzen der Sammelgattung Didymella (Ascomycetes, Dothideales). Nova Hedwigia, 51, 283–360.Google Scholar
Grube, M. and Kantvilas, G. (2006). Siphula represents a remarkable case of morphological congruence in sterile lichens. Lichenologist, 38, 241–249.CrossRefGoogle Scholar
Grube, M. and Kroken, S. (2000). Molecular approaches and the concept of species and species complexes in lichenized fungi. Mycological Research, 104, 1284–1294.CrossRefGoogle Scholar
Grube, M., Baloch, E. and Lumbsch, H. T. (2004). The phylogeny of Porinaceae (Ostropomycetidae) suggests a neotenic origin of perithecia in Lecanoromycetes. Mycological Research, 108, 1111–1118.CrossRefGoogle ScholarPubMed
Guenther, J. E. and Melis, A. (1990). Dynamics of photosystem II heterogeneity in Dunaliella salina (green alga). Photosynthesis Research, 23, 195–203.CrossRefGoogle Scholar
Gugger, M. F. and Hoffmann, L. (2004). Polyphyly of true branching cyanobacteria (Stigonematales). International Journal of Systematic and Evolutionary Microbiology, 54, 349–357.CrossRefGoogle Scholar
Gunn, J., Keller, W., Negusanti, J., et al. (1995). Ecosystem recovery after emission reductions: Sudbury, Canada. Water, Air, and Soil Pollution, 85, 1783–1788.CrossRefGoogle Scholar
Gunther, A. J. (1988). Effects of simulated acid rain on nitrogenase activity in the lichen genus Peltigera under field and laboratory conditions. Water, Air, and Soil Pollution, 38, 379–385.Google Scholar
Gunther, A. J. (1989). Nitrogen fixation by lichens in a subarctic Alaskan watershed. Bryologist, 92, 202–208.CrossRefGoogle Scholar
Gupta, V. (2005). Application of lichenometry to slided materials in the Higher Himalayan landslide zone. Current Science, 89, 1032–1036.Google Scholar
Haas, D. and Keel, C. (2003). Regulation of antibiotic production in root-colonizing Pseudomonas spp. and relevance for biological control of plant disease. Annual Review of Phytopathology, 41, 117–153.CrossRefGoogle ScholarPubMed
Hafellner, J. (1984). Studien in Richtung einer naturlicheren Gliederung der Sammelfamilien Lecanoraceae und Lecideaceae. Beitrage zur Lichenologie. Festschrift J. Poelt. Beihefte zur Nova Hedwigia, 79, 241–371.Google Scholar
Hafellner, J. (1995) A new checklist of lichens and lichenicolous fungi of insular Laurimacaronesia including a lichenological bibliography for the area. Fristchiana 5, 3–132.Google Scholar
Hafellner, J., Hertel, H., Rambold, G. and Timdal, E. (1993). A New Outline of the Lecanorales. Graz: Privately published by the authors.Google Scholar
Häffner, E., Lomský, B., Hynek, V.,et al. (2001). Air pollution and lichen physiology. Physiological responses of different lichens in a transplant experiment following an SO2-gradient. Water, Air, and Soil Pollution, 131, 185–201.CrossRefGoogle Scholar
Hageman, C. M. (1989). Enzyme electromorph variation in the lichen family Umbilicariaceae. Ph.D. thesis. London, Ontario: University of Western Ontario.
Hageman, C. M. and Fahselt, D. (1986). A comparison of isozyme patterns of morphological variants in the lichen Umbilicaria muhlenbergii (Ach.) Tuck. Bryologist, 89, 285–290.CrossRefGoogle Scholar
Hageman, C. M. and Fahselt, D. (1990). Enzyme electromorph variation in the lichen family Umbilicariaceae: within-stand polymorphism in umbilicate lichens of eastern Canada. Canadian Journal of Botany, 68, 2636–2643.CrossRefGoogle Scholar
Hageman, C. M. and Fahselt, D. (1992). Geographical distance and enzyme polymorphisms in the lichen Umbilicaria mammulata. Bryologist, 93, 316–323.CrossRefGoogle Scholar
Hahn, S. C., Tenhunen, J. D., Popp, P. W., Meyer, A. and Lange, O. L. (1993). Upland tundra in the foothills of the Brooks Range, Alaska: diurnal CO2 exchange patterns of characteristic lichen species. Flora, 188, 125–143.CrossRefGoogle Scholar
Hale, M. E. (1973). Growth. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 473–492. New York: Academic Press.Google Scholar
Hale, M. E. (1974). The Biology of Lichens. 2nd edn. London: Edward Arnold.Google Scholar
Hale, M. E. (1983). The Biology of Lichens. 3rd edn. London: Edward Arnold.Google Scholar
Hale, M. E. (1984). The lichen line and high water levels in a freshwater stream in Florida. Bryologist, 87, 261–265.CrossRefGoogle Scholar
Hale, M. E. (1990). A synopsis of the lichen genus Xanthoparmelia (Vainio) Hale (Ascomycotina, Parmeliaceae). Smithsonian Contributions to Botany, 74, 1–250.CrossRefGoogle Scholar
Hallbom, L. and Bergman, B. (1979). Influence of certain herbicides and a forest fertilizer on the nitrogen fixation by the lichen Peltigera praetextata. Oecologia, 40, 19–27.CrossRefGoogle Scholar
Hällgren, J.-E. and Huss, K. (1975). Effects of SO2 on photosynthesis and nitrogen fixation. Physiologia Plantarum, 34, 171–176.CrossRefGoogle Scholar
Hallingbäck, T. (1991). Blue-green algae and cyanophilic lichens are threatened by air pollution and fertilization. Svensk Botanisk Tidskrift, 85, 87–104.Google Scholar
Hallingbäck, T. and Kellner, O. (1992). Effects of simulated nitrogen rich and acid rain on the nitrogen-fixing lichen Peltigera aphthosa (L.) Willd. New Phytologist, 120, 99–103.CrossRefGoogle Scholar
Halliwell, B. (2006). Reactive species and antioxidants. Redox biology is a fundamental theme of aerobic life. Plant Physiology, 141, 312–322.CrossRefGoogle ScholarPubMed
Halliwell, B. and Gutteridge, J. M. C. (1999). Free Radicals in Biology and Medicine. Oxford: Oxford University Press.Google Scholar
Hamada, N., Miyawaki, H. and Yamada, A. (1995). Distribution pattern of air pollution and epiphytic lichens in the Osaka Plain (Japan). Journal of Plant Research, 108, 483–491.CrossRefGoogle Scholar
Hamada, N., Tanahashi, T., Miyagawa, H. and Miyawaki, H. (2001). Characteristics of secondary metabolites from isolated lichen mycobionts. Symbiosis, 31, 23–33.Google Scholar
Hammer, S. (2003). Notocladonia, a new genus in the Cladoniaceae. Bryologist, 106, 162–167.CrossRefGoogle Scholar
Hardy, R. W. F., Burns, R. C. and Holsten, R. D. (1973). Applications of the C2H2 reduction assay for measurement of N2 fixation. Soil Biology and Biochemistry, 5, 47–81.CrossRefGoogle Scholar
Harper, K. T. and Marble, J. R. (1988). A role of nonvascular plants in management of semiarid rangelands. In Vegetation Science Applications for Rangeland Analysis and Management, ed. Tueller, P. T., pp. 135–169. London: Kluwer Academic.CrossRefGoogle Scholar
Harris, G. B. (1971). The ecology of corticolous lichens. I. The zonation on oak and birch in South Devon. Journal of Ecology, 59, 431–439.CrossRefGoogle Scholar
Harrison, S. and Winchester, V. (2000). Nineteenth- and twentieth-century glacier fluctuations and climatic implications in the Arco and Colonia valleys, Hielo Patagónico Norte, Chile. Arctic, Antarctic, and Alpine Research, 32, 55–63.CrossRefGoogle Scholar
Hasegawa, M., Kishino, H. and Yano, K. (1985). Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. Journal of Molecular Evolution, 22, 160–174.CrossRefGoogle ScholarPubMed
Hasenhüttl, G. and Poelt, J. (1978). Über die Brutkörner bei der Flechtengattung Umbilicaria. Berichte der deutschen botanischen Gesellschaft, 91, 275–296.Google Scholar
Hasse, H. E. (1913). The lichen flora of southern California. Contributions from the United States National Herbarium, 17, 1–132.Google Scholar
Hauck, M. and Paul, A. (2005). Manganese as a site factor for epiphytic lichens. Lichenologist, 37, 409–423.CrossRefGoogle Scholar
Hauck, M. and Spribille, T. (2005). The significance of precipitation and substrate chemistry for epiphytic lichen diversity in spruce-fir forests of the Salish Mountains, northwestern Montana. Flora, 200, 547–562.CrossRefGoogle Scholar
Hauck, M. and Zoller, T. (2003). Copper sensitivity of soredia of the epiphytic lichen Hypogymnia physodes. Lichenologist, 35, 271–274.CrossRefGoogle Scholar
Hauck, M., Hesse, V., Jung, R., Zöller, T. and Runge, M. (2001). Long-distance transported sulphur as a limiting factor for the abundance of Lecanora conizaeoides in montane spruce forests. Lichenologist, 33, 267–269.CrossRefGoogle Scholar
Hauck, M., Hesse, V. and Runge, M. (2002 a). Correlations between the Mn/Ca ratio in stemflow and epiphytic lichen abundance in a dieback-affected spruce forest of the Harz Mountains, Germany. Flora, 197, 361–369.CrossRefGoogle Scholar
Hauck, M., Mulack, C. and Paul, A. (2002 b). Manganese uptake in the epiphytic lichens Hypogymnia physodes and Lecanora conizaeoides. Environmental and Experimental Botany, 48, 107–117.CrossRefGoogle Scholar
Hauck, M., Paul, A., Mulack, C., Fritz, E. and Runge, M. (2002 c). Effects of manganese on the viability of vegetative diaspores of the epiphytic lichen Hypogymnia physodes. Environmental and Experimental Botany, 47, 127–142.CrossRefGoogle Scholar
Hauck, M., Paul, A. and Spribille, T. (2006). Uptake and toxicity of manganese in epiphytic cyanolichens. Environmental and Experimental Botany, 56, 216–224.CrossRefGoogle Scholar
Hawksworth, D. L. (1971). Lichens as a litmus for air pollution: a historical review. International Journal of Environmental Studies, 1, 281–296.CrossRefGoogle Scholar
Hawksworth, D. L. (1982). Secondary fungi in lichen symbioses: parasites, saprophytes and parasymbionts. Journal of the Hattori Botanical Laboratory, 52, 357–366.Google Scholar
Hawksworth, D. L. (1983). A key to the lichen-forming, parasitic, parasymbiotic and saprophytic fungi occurring on lichens in the British Isles. Lichenologist, 15, 1–44.CrossRefGoogle Scholar
Hawksworth, D. L. (1985). Problems and prospects in the systematics of the Ascomycotina. Proceedings of the Indian Academy of Sciences (Plant Sciences), 94, 319–39.Google Scholar
Hawksworth, D. L. (1988 a). The fungal partner. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 35–38. Boca Raton: CRC Press.Google Scholar
Hawksworth, D. L. (1988 b). The variety of fungal-algal symbioses, their evolutionary significance, and the nature of lichens. Botanical Journal of the Linnean Society, 96, 3–20.CrossRefGoogle Scholar
Hawksworth, D. L. (1988 c). Effects of algae and lichen-forming fungi on tropical crops. In Perspectives of Mycopathology, ed. Agnihotri, V. P., Sarbhoy, K. A. and Kumar, D., pp. 76–83. New Delhi: Malhorta Publishing House.Google Scholar
Hawksworth, D. L. (1988 d). Conidiomata, conidiogenesis, and conidia. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 181–193. Boca Raton: CRC Press.Google Scholar
Hawksworth, D. L. (1991). The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycological Research, 95, 641–655.CrossRefGoogle Scholar
Hawksworth, D. L. (2002). Bioindication: calibrated scales and their utility. In Monitoring with Lichens – Monitoring Lichens, Nato Science Series IV: Earth and Environmental Sciences, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 11–20. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Hawksworth, D. L. (2003). The lichenicolous fungi of Great Britain and Ireland: an overview and annotated checklist. Lichenologist, 35, 191–232.CrossRefGoogle Scholar
Hawksworth, D. L. and Hill, D. J. (1984). The Lichen-forming Fungi. Glasgow: Blackie.CrossRefGoogle Scholar
Hawksworth, D. L. and Honegger, R. (1994). The lichen thallus: a symbiotic phenotype of nutritionally specialized fungi and its response to gall producers. In Plant Galls: Organisms, Interactions, Populations, ed. Williams, M. A. J., pp. 77–98. Oxford: Clarendon Press.Google Scholar
Hawksworth, D. L. and McManus, P. M. (1989). Lichen recolonization in London under conditions of rapidly falling sulphur dioxide levels, and the concept of zone skipping. Botanical Journal of the Linnean Society, 100, 99–109.CrossRefGoogle Scholar
Hawksworth, D. L. and McManus, P. M. (1992). Changes in the lichen flora on trees in Epping Forest through periods of increasing and then ameliorating sulphur dioxide air pollution. Essex Naturalist, 11, 92–101.Google Scholar
Hawksworth, D. L. and Rose, F. (1970). Qualitative scale for estimating sulphur dioxide air pollution in England and Wales using epiphytic lichens. Nature, 227, 145–148.CrossRefGoogle ScholarPubMed
Hawksworth, D. L., Kirk, P. M., Sutton, B. C. and Pegler, D. N. (1995). Ainsworth and Bisby's Dictionary of the Fungi. 8th edn. Wallingford: CAB International.Google Scholar
Hawksworth, D. L., Lawton, R. M., Martin, P. G. and Stanley-Price, K. (1984). Nutritive value of Ramalina duriaei grazed by gazelles in Oman. Lichenologist, 16, 93–94.CrossRefGoogle Scholar
Hawksworth, D. L., Sutton, B. C. and Ainsworth, D. C. (1983). Ainsworth and Bisby's Dictionary of the Fungi. 7th edn. Kew: Commonwealth Mycological Institute.Google Scholar
Hayward, G. D. and Rosentreter, R. (1994). Lichens as nesting material for northern flying squirrels in the northern Rocky Mountains. Journal of Mammalogy, 75, 663–673.CrossRefGoogle Scholar
Heads, M. (1997). Regional patterns of biodiversity in New Zealand: one degree grid analysis of plant and animal distributions. Journal of the Royal Society of New Zealand, 27, 337–354.CrossRefGoogle Scholar
Heads, M. (1998). Biogeographic disjunction along the Alpine Fault, New Zealand. Biological Journal of the Linnean Society, 63, 161–176.CrossRefGoogle Scholar
Heads, M. (2001). Birds of paradise, biogeography and ecology in New Guinea: a review. Journal of Biogeography, 28, 893–927.CrossRefGoogle Scholar
Heads, M. (2004). What is a node?Journal of Biogeography, 31, 1883–1891.CrossRefGoogle Scholar
Heads, M. (2005). Towards a panbiogeography of the seas. Biological Journal of the Linnean Society, 84, 675–723.CrossRefGoogle Scholar
Heads, M. and Craw, R. (2004). The alpine fault biogeographic hypothesis revisited. Cladistics, 20, 184–190.CrossRefGoogle Scholar
Heber, U., Bilger, W. and Shuvalov, V. A. (2006 a). Thermal energy dissipation in reaction centers and in the antenna of photosystem II protects desiccated poikilohydric mosses against photo-oxidation. Journal of Experimental Botany, 57, 2993–3006.CrossRefGoogle Scholar
Heber, U., Lange, O. L. and Shuvalov, V. A. (2006 b). Conservation and dissipation of light energy as complementary processes: homoiohydric and poikilohydric autotrophs. Journal of Experimental Botany, 57, 1211–1223.CrossRefGoogle ScholarPubMed
Heibel, E., Lumbsch, H. T. and Schmitt, I. (1999). Genetic variation of Usnea filipendula (Parmeliaceae) populations in western Germany investigated by RAPDs suggest reinvasion from various sources. American Journal of Botany, 86, 753–757.CrossRefGoogle ScholarPubMed
Heiđmarsson, S., Mattson, J. E., Moberg, R., et al. (1997). Classification of lichen photomorphs. Taxon, 46, 519–520.CrossRefGoogle Scholar
Helms, G. (2003). Taxonomy and symbiosis in associations of Physciaceae and Trebouxia. Ph.D. thesis. Göttingen: Albrecht-von-Haller Institute, University of Göttingen.
Helms, G., Friedl, T., Rambold, G. and Mayrhofer, H. (2001). Identification of photobionts from the lichen family Physciaceae using algal-specific ITS rDNA sequencing. Lichenologist, 33, 73–86.CrossRefGoogle Scholar
Henderson, A. (1999). Lichen dyes. An historical perspective. Lees Museums and Galleries Review, 2, 30–34.Google Scholar
Henderson-Sellers, A. and Seaward, M. R. D. (1979). Monitoring lichen reinvasion of ameliorating environments. Environmental Pollution, 19, 207–213.CrossRefGoogle Scholar
Henriksson, E. and Pearson, L. C. (1981). Nitrogen fixation rate and chlorophyll content of the lichen Peltigera canina exposed to sulfur dioxide. American Journal of Botany, 68, 680–684.CrossRefGoogle Scholar
Henson, B. J., Hesselbrock, S. M., Watson, L. E., and Barnum, S. R. (2004). Molecular phylogeny of the heterocystous cyanobacteria (subsections IV and V) based on nifD. International Journal of Systematic and Evolutionary Microbiology, 54, 493–497.CrossRefGoogle ScholarPubMed
Henssen, A. (1963). Eine Revision der Flechtenfamilien Lichinaceae und Ephebaceae. Symbolae Bototanicae Upsala, 18, 1–123.Google Scholar
Henssen, A., in cooperation with Keuck, G., Renner, B. and Vobis, G. (1981). The lecanoralean centrum. In Ascomycete Systematics: The Lutrellian Concept, ed. Reynolds, D. R., pp. 138–234. New York: Springer.CrossRefGoogle Scholar
Henssen, A. (1986). The genus Paulia (Lichinaceae). Lichenologist, 18, 201–229.CrossRefGoogle Scholar
Henssen, A. (1995). The new lichen family Gloeoheppiaceae and its genera Gloeoheppia, Pseudopeltula and Gudella (Lichinales). Lichenologist, 27, 261–290.CrossRef
Henssen, A. and Jahns, H. M. (1973[1974]). Lichenes. Eine Einführung in die Flechtenkunde. Stuttgart: Thieme.Google Scholar
Henssen, A. and Tretiach, M. (1995). Paulia glomerata, a new epilithic species from Europe, and additional notes on some other Paulia species. Nova Hedwigia, 60, 297–309.Google Scholar
Henssen, A., Büdel, B. and Titze, A. (1987). Euopsis and Harpidium, genera of the Lichinaceae (Lichenes) with rostrate asci. Botanica Acta, 101, 49–55.Google Scholar
Henzler, T. and Steudle, E. (2000). Transport and metabolic degradation of hydrogen peroxide in Chara coralline: model calculations and measurements with the pressure probe suggest transport of H2O2 across water channels. Journal of Experimental Botany, 51, 2053–2066.CrossRefGoogle Scholar
Herrera-Campos, M. A., Lücking, R., Perez, R. E., et al. (2004). The foliicolous lichen flora of Mexico. V. Biogeographical affinities, altitudinal preferences, and an updated checklist of 293 species. Lichenologist, 36, 309–327.CrossRefGoogle Scholar
Hersoug, L. G. (1983). Lichen protein affinity towards walls of cultured and freshly isolated phycobionts and its relationship to cell wall cytochemistry. FEMS Microbiology Letters, 20, 417–420.CrossRefGoogle Scholar
Herzig, R. and Urech, M. (1991). Flechten als Bioindikatoren: Intergriertes biologisches Messsystem der Luftverschmutzung für das Schweizer Mittelland. Bibliotheca Lichenologica, 43, 1–283.Google Scholar
Hesbacher, S., Fröbery, L., Baur, A., Baur, B. and Proksch, P. (1996). Chemical variation within and between individuals of the lichenized ascomycete Tephromela atra. Biochemical Systematics and Ecology, 24, 603–609.CrossRefGoogle Scholar
Hestmark, G. (1990). Thalloconidia in the genus Umbilicaria. Nordic Journal of Botany, 9, 547–574.CrossRefGoogle Scholar
Hestmark, G. (1992). Sex, size competition and escape – strategies of reproduction and dispersal in Lasallia pustulata (Umbilicariaceae, Ascomycetes). Oecologia, 92, 305–312.CrossRefGoogle Scholar
Hestmark, G. (1997). Species diversity and reproductive strategies in the family Umbilicariaceae on high equatorial mountains – with remarks on global patterns. Bibliotheca Lichenologica, 68, 195–202.Google Scholar
Hestmark, G. (2000). The ecophysiology of lichen population biology. Bibliotheca Lichenologica, 75, 397–403.Google Scholar
Hestmark, G., Schroeter, B. and Kappen, L. (1997). Intrathalline and size-dependent patterns of activity in Lasallia pustulata and their possible consequences for competitive interactions. Functional Ecology, 11, 318–322.CrossRefGoogle Scholar
Heywood, V. H. (ed.) (1995). Global Biodiversity Assessment. Cambridge: Cambridge University Press.Google Scholar
Hildreth, K. C. and Ahmadjian, V. (1981). A study of Trebouxia and Pseudotrebouxia isolated from different lichens. Lichenologist, 13, 65–86.CrossRefGoogle Scholar
Hill, D. J. (1971). Experimental study of the effect of sulfite on lichens with reference to atmospheric pollution. New Phytologist, 70, 831–836.CrossRefGoogle Scholar
Hill, D. J. (1976). The physiology of lichen symbiosis. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 457–496. London: Academic Press.Google Scholar
Hill, D. J. (1985). Changes in photobiont dimensions and numbers during co-development of lichen symbionts. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 303–317. New York: Plenum Press.CrossRefGoogle Scholar
Hill, D. J. (1989). The control of the cell cycle in microbial symbionts. New Phytologist, 112, 175–184.CrossRefGoogle Scholar
Hill, D. J. and Smith, D. C. (1972). Lichen physiology XII. The “inhibition technique”. New Phytologist, 71, 15–30.CrossRefGoogle Scholar
Hilmo, O. (1994). Distribution and succession of epiphytic lichens on Picea abies in a boreal forest, central Norway. Lichenologist, 26, 149–169.CrossRefGoogle Scholar
Hilmo, O. and Holien, H. (2002). Epiphytic lichen response to the edge environment in a boreal Picea abies forest in central Norway. Bryologist, 105, 48–56.CrossRefGoogle Scholar
Hilmo, O. and Ott, S. (2002). Juvenile development of the cyanolichen Lobaria scrobiculata and the green algal lichen Platismatia glauca and Platismatia norvegica in a boreal Picea abies forest. Plant Biology, 4, 273–280.CrossRefGoogle Scholar
Hilmo, O. and Sastad, S. M. (2001). Colonization of old-forest lichens in a young and an old boreal Picea abies forest: an experimental approach. Biological Conservation, 102, 251–259.CrossRefGoogle Scholar
Hilmo, O., Holien, H. and Hytteborn, H. (2005). Logging strategy influences colonization of common chlorolichens on branches of Picea abies. Ecological Applications, 15, 983–996.CrossRefGoogle Scholar
Hirt, R. P., Logsdon, J., Doolittle, W. F. and Embley, T. M. (1999). Microsporidia are related to Fungi: evidence from the largest subunit of RNA polymerase II and other proteins. Proceedings of the National Academy of Sciences, USA, 96, 580–585.CrossRefGoogle ScholarPubMed
Hiserodt, R. D., Swijter, D. F. H. and Mussinan, C. J. (2000). Identification of atranorin and related potential allergens in oakmoss absolute by high-performance liquid chromatography-tandem mass spectrometry using negative ion atmospheric pressure chemical ionization. Journal of Chromatography A, 888, 103–111.CrossRefGoogle ScholarPubMed
Hitch, C. J. B. and Millbank, J. W. (1975). Nitrogen metabolism in lichens. VI. The blue-green phycobiont content, heterocyst frequency and nitrogenase activity in Peltigera species. New Phytologist, 74, 473–476.CrossRefGoogle Scholar
Hitch, C. J. B. and Millbank, J. W. (1976). Nitrogen metabolism in lichens. VII. Nitrogenase activity and heterocyst frequency in lichens with blue-green phycobionts. New Phytologist, 75, 239–244.CrossRefGoogle Scholar
Hocking, D., Kuchar, P., Plambeck, J. A. and Smith, R. A. (1978). The impact of gold smelter emissions on vegetation and soils of a sub-arctic forest–tundra transition ecosystem. Air Pollution Control Association Journal, 28, 133–137.CrossRefGoogle Scholar
Högnabba, F. (2006). Molecular phylogeny of the genus Stereocaulon (Stereocaulaceae, lichenized Ascomycetes). Mycological Research, 110, 1080–1092.CrossRefGoogle Scholar
Holopainen, T. (1983). Ultrastructural changes in epiphytic lichen Bryoria capillaris and Hypogymnia physodes in central Finland. Annales Botanici Fennici, 19, 39–52.Google Scholar
Holopainen, T. (1984). Types and distribution of ultrastructural symptoms in epiphytic lichens in several urban and industrial environments in Finland. Annales Botanici Fennici, 21, 219–229.Google Scholar
Holopainen, T. and Kärenlampi, L. (1984). Injuries to lichen ultrastructure caused by sulphur dioxide fumigations. New Phytologist, 98, 285–294.CrossRefGoogle Scholar
Holopainen, T. and Kärenlampi, L. (1985). Characteristic ultrastructural symptoms caused in lichens by experimental exposure to nitrogen compounds and fluorides. Annales Botanici Fennici, 22, 333–342.Google Scholar
Holopainen, T. and Kauppi, M. (1989). A comparison of light, fluorescence and electron microscopic observations in assessing the SO2 injury of lichens under different moisture conditions. Lichenologist, 21, 119–134.CrossRefGoogle Scholar
Holt, N. E., Tigmantas, D., Valkunas, L., et al. (2005). Carotenoid cation formation and the regulation of photosynthetic light harvesting. Science, 307, 433–436.CrossRefGoogle ScholarPubMed
Holub, S. M. and Lajtha, K. (2003). Mass loss and nitrogen dynamics during the decomposition of a 15N-labeled N2-fixing epiphytic lichen, Lobaria oregana. Canadian Journal of Botany, 81, 698–705.CrossRefGoogle Scholar
Holub, S. M. and Lajtha, K. (2004). The fate and retention of organic and inorganic 15N-nitrogen in an old-growth forest soil in Western Oregon. Ecosystems, 7, 368–380.CrossRefGoogle Scholar
Honegger, R. (1978 a). Ascocarpontogenie, Ascusstruktur und -funktion bei Vertretern der Gattung Rhizocarpon. Berichte der deutschen botanischen Gesellschaft, 91, 579–594.Google Scholar
Honegger, R. (1978 b). The ascus apex in lichenized fungi. I. The Lecanora-, Peltigera- and Teloschistes-types. Lichenologist, 10, 47–67.CrossRefGoogle Scholar
Honegger, R. (1980). The ascus apex in lichenized fungi. II. The Rhizocarpon-type. Lichenologist, 12, 157–172.CrossRefGoogle Scholar
Honegger, R. (1982 a). The ascus apex in lichenized fungi. III. The Pertusaria–type. Lichenologist, 14, 205–217.CrossRefGoogle Scholar
Honegger, R. (1982 b). Ascus structure and function, ascospore delimitation, and phycobiont cell wall types associated with the Lecanorales (lichenized ascomycetes). Journal of the Hattori Botanical Laboratory, 52, 417–429.Google Scholar
Honegger, R. (1984 a). Scanning electron microscopy of the contact site of conidia and trichogynes in Cladonia furcata. Lichenologist, 16, 11–19.CrossRefGoogle Scholar
Honegger, R. (1984 b). Ultrastructural studies on conidiomata, conidiophores, and conidiogenous cells in six lichen-forming Ascomycetes. Canadian Journal of Botany, 62, 2081–2093.CrossRefGoogle Scholar
Honegger, R. (1985). Ascus structure and ascospore formation in the lichen-forming Chaenotheca chrysocephala (Caliciales). Sydowia, Annales Mycologici Series II, 38, 146–157.Google Scholar
Honegger, R. (1986 a). Ultrastructural studies in lichens. I. Haustorial types and their frequencies in a range of lichens with trebouxioid phycobionts. New Phytologist, 103, 785–795.CrossRefGoogle Scholar
Honegger, R. (1986 b). Ultrastructural studies in lichens. II. Mycobiont and photobiont cell wall surface layers and adhering crystalline lichen products in four Parmeliaceae. New Phytologist, 103, 797–808.CrossRefGoogle Scholar
Honegger, R. (1991 a). Functional aspects of the lichen symbiosis. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 553–578.CrossRefGoogle Scholar
Honegger, R. (1991 b). Fungal evolution: Symbioses and morphogenesis. In Symbiosis, a Source of Evolutionary Innovation, ed. Margulis, L. and Fester, R., pp. 319–340. Cambridge: Massachusetts Institute of Technology Press.Google ScholarPubMed
Honegger, R. (1991 c). Haustoria-like structures and hydrophobic cell wall surface layers in lichens. In Electron Microscopy of Plant Pathogens, ed. Mendgen, K. and Lesemann, D. E., pp. 277–290. Berlin: Springer.CrossRefGoogle Scholar
Honegger, R. (1992). Lichens: mycobiont-photobiont relationships. In Algae and Symbioses. Plants, Animals, Fungi, Viruses, Interactions Explored, ed. Reisser, W., pp. 255–275. Bristol: Biopress.Google Scholar
Honegger, R. (1993). Developmental biology of lichens. New Phytologist, 125, 659–677.CrossRefGoogle Scholar
Honegger, R. (1995). Experimental studies with foliose macrolichens: fungal responses to spatial disturbance at the organismic level and to spatial problems at the cellular level. Canadian Journal of Botany, 73, 569–578.CrossRefGoogle Scholar
Honegger, R. (1997). Metabolic interactions at the mycobiont-photobiont interface in lichens. In Plant Relationships, The Mycota, Vol. V, Part A, ed. Carroll, G. C. and Tudzynski, P., pp. 209–221. Berlin: Springer.Google Scholar
Honegger, R. (1998). The lichen symbiosis – what is so spectacular about it?Lichenologist, 30, 193–212.CrossRefGoogle Scholar
Honegger, R. (2000). Great discoveries in bryology and lichenology – Simon Schwendener (1829–1919) and the Dual Hypothesis of Lichens. Bryologist, 103, 307–313.CrossRefGoogle Scholar
Honegger, R. (2001). The symbiotic phenotype of lichen-forming ascomycetes. In Fungal Associations, Vol. IX: The Mycota, ed. Hock, B., pp. 165–188. Berlin: Springer.CrossRefGoogle Scholar
Honegger, R. (2006). Water relations in lichens. In Fungi in the Environment, ed. Gadd, G. M., Watkinson, S. C. and Dyer, P., pp. 185–200. Cambridge: Cambridge University Press.Google Scholar
Honegger, R. and Bartnicki-Garcia, S. (1991). Cell wall structure and composition of cultured mycobionts from the lichen Cladonia macrophylla, Cladonia caespiticia, and Physcia stellaris (Lecanorales, Ascomycetes). Mycological Research, 95, 905–914.CrossRefGoogle Scholar
Honegger, R. and Haisch, A. (2001). Immunocytochemical location of the (1 → 3) (1 → 4)-beta-glucan lichenin in the lichen-forming ascomycete Cetraria islandica (Icelandic moss). New Phytologist, 150, 739–746.CrossRefGoogle Scholar
Honegger, R. and Zippler, U. (2007). Mating systems in representatives of the Parmeliaceae, Ramalinaceae and Physciaceae (Lecanoromycetes, lichen-forming ascomycetes). Mycological Research, 11, 424–432.CrossRefGoogle Scholar
Honegger, R., Peter, M. and Scherrer, S. (1996). Drought-stress induced structural alterations at the mycobiont-photobiont interface in a range of foliose macrolichens. Protoplasma, 190, 221–232.CrossRefGoogle Scholar
Honegger, R., Zippler, U., Gansner, H. and Scherrer, S. (2004). Mating systems in the genus Xanthoria (lichen-forming ascomycetes). Mycological Research, 108, 480–488.CrossRefGoogle Scholar
Hopwood, D. A. (1997). Genetic contribution to understanding polyketide synthases. Chemical Reviews, 97, 2465–2498.CrossRefGoogle Scholar
Horstmann, J. L., Denison, W. C. and Silvester, W. B. (1982). 15N2 fixation and molybdenum enhancement of acetylene reduction by Lobaria spp. New Phytologist, 92, 235–241.CrossRefGoogle Scholar
Houdijk, A. L. F. M. and Roelofs, J. G. M. (1991). Deposition of acidifying and eutrophicating substances in Dutch forest. Acta Botanica Neerlandica, 40, 245–255.CrossRefGoogle Scholar
Huebert, D. B., L'Hirondelle, S. J. and Addison, P. A. (1985). The effects of sulphur dioxide on net CO2 assimilation in the lichen Evernia mesomorpha Nyl. New Phytologist, 100, 643–651.CrossRefGoogle Scholar
Huelsenbeck, J. P. and Ronquist, F. (2001). Bayesian inference of phylogenetic trees. Bioinformatics, 17, 754–755.CrossRefGoogle ScholarPubMed
Humphries, C. J. and Ebach, M. C. (2004). Biogeography on a dynamic earth. In Frontiers of Biogeography: New Directions in the Geography of Nature, ed. Lomolino, M. V. and Heaney, L. R., pp. 67–86. Sunderland: Sinauer Associates.Google Scholar
Humphries, C. J. and Parenti, L. R. (1999). Cladistic Biogeography: Interpreting Patterns of Plant and Animal Distributions. 2nd edn, Oxford Biogeography Series 12. Oxford: Oxford University Press.Google Scholar
Humphries, C. J., Williams, P. H. and Vane-Wright, R. I. (1995). Measuring biodiversity value for conservation. Annual Reviews of Ecology and Systematics, 26, 93–111.CrossRefGoogle Scholar
Huneck, S. (1999). The significance of lichens and their metabolites. Die Naturwissenschaften, 86, 559–570.CrossRefGoogle ScholarPubMed
Huneck, S. (2001). New results on the chemistry of lichen substances. In Progress in the Chemistry of Organic Products, ed. Herz, W., Falk, H., Kirby, G. W. and Moore, R. E., pp. 1–276. New York: Springer.Google Scholar
Huneck, S. and Schreiber, K. (1972). Wachstumsregulatorische Eigenschaften von Flechten- und Moos-Inhaltstoffen. Phytochemistry, 11, 2429–2434.CrossRefGoogle Scholar
Huneck, S. and Yoshimura, I. (1996). Identification of Lichen Substances. Springer: Berlin.CrossRefGoogle Scholar
Huneck, S., Bothe, H.-K. and Richter, W. (1990). Über den Metallgehalt von Flechten von Kupferschieferhalden der Umgebung von Mansfeld. Herzogia, 8, 295–304.Google Scholar
Huovinen, K., Hiltunen, R. and Schantz, M. (1985). A high performance liquid chromatographic method for the analyses of lichen compounds from the genera Cladina and Cladonia. Acta Pharmaceutica Fennica, 94, 99–112.Google Scholar
Huss, V. A. R. and Sogin, M. L. (1990). Phylogenetic position of some Chlorella species within the Chlorococcales based upon complete small-subunit ribosomal RNA sequences. Journal of Molecular Evolution, 31, 432–442.CrossRefGoogle ScholarPubMed
Huss-Danell, K. (1977). Nitrogen fixation by Stereocaulon paschale under field conditions. Canadian Journal of Botany, 55, 585–592.CrossRefGoogle Scholar
Huss-Danell, K. (1978). Seasonal variation in the capacity for nitrogenase activity in the lichen Stereocaulon paschale. New Phytologist, 81, 89–98.CrossRefGoogle Scholar
Huss-Danell, K. (1979). The cephalodia and their nitrogenase activity in the lichen Stereocaulon paschale. Zeitschrift für Pflanzenphysiologie, 95, 431–440.CrossRefGoogle Scholar
Hyvärinen, M. and Crittenden, P. D. (1998 a). Relationships between atmospheric nitrogen inputs and the vertical nitrogen and phosphorus concentration gradients in the lichen Cladonia portentosa. New Phytologist, 140, 519–530.CrossRefGoogle Scholar
Hyvärinen, M. and Crittenden, P. D. (1998 b). Growth of the cushion-forming lichen, Cladonia portentosa, at nitrogen-polluted and unpolluted heathland sites. Environmental and Experimental Botany, 40, 67–76.CrossRefGoogle Scholar
Hyvärinen, M. and Crittenden, P. D. (2000). 33P translocation in the thallus of the mat-forming lichen Cladonia portentosa. New Phytologist, 145, 281–288.CrossRefGoogle Scholar
Hyvärinen, M., Härdling, R. and Tuomi, J. (2002). Cyanobacterial lichen symbiosis: the fungal partner as an optimal harvester. Oikos, 98, 498–504.CrossRefGoogle Scholar
Hyvärinen, M., Roitto, M., Ohtonen, R. and Markkola, A. (2000). Impact of wet deposited nickel on the cation content of a mat-forming lichen Cladina stellaris. Environmental and Experimental Botany, 43, 211–218.CrossRefGoogle ScholarPubMed
Hyvärinen, M., Walter, B. and Koopmann, R. (2003). Impact of fertilization on phenol content and growth rate of Cladina stellaris: a test of the carbon-nutrient balance hypothesis. Oecologia, 134, 176–181.CrossRefGoogle ScholarPubMed
Ihda, T. A., Nakano, T., Yoshimura, I. and Iwatsuki, Z. (1993). Phycobionts isolated from Japanese species of Anzia (lichenes). Archiv für Protistenkunde, 143, 163–172.CrossRefGoogle Scholar
Ihlen, P. G. and Ekman, S. (2002). Outline of phylogeny and character evolution in Rhizocarpon (Rhizocarpaceae, lichenized Ascomycota) based on nuclear ITS and mitochondrial SSU ribosomal DNA sequences. Biological Journal of the Linnean Society, London, 77, 535–546.CrossRefGoogle Scholar
Ingólfsdóttir, K. (2002). Molecules of interest: usnic acid. Phytochemistry, 61, 729–736.CrossRefGoogle Scholar
Innes, J. L. (1983). Lichenometric dating of debris-flow deposits in the Scottish Highlands. Earth Surface Processes and Landforms, 8, 579–588.CrossRefGoogle Scholar
Innes, J. L. (1985). Lichenometry. Progress in Physical Geography, 9, 187–254.CrossRefGoogle Scholar
Innes, J. L. (1988). The use of lichens in dating. In Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 75–91. Boca Raton: CRC Press.Google Scholar
Insarov, G. and Schroeter, B. (2002). Lichen monitoring and climate change. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 183–201. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Jaag, O. and Thomas, E. (1934). Neue Untersuchungen über die Flechte Epigloea bactrospora Zukal. Berichte der schweierischen botanischen Gesellschaft, 34, 77–89.Google Scholar
Jahns, H. M. (1970). Untersuchungen zur Entwicklungsgeschichte der Cladoniaceen, unter besonderer Berücksichtigung des Podetien-Problems. Nova Hedwigia, 20, 1–177.Google Scholar
Jahns, H. M. (1984). Morphology, reproduction and water relations – a system of morphogenetic interactions in Parmelia saxatilis. Nova Hedwigia, 79, 715–37.Google Scholar
Jahns, H. M. (1987). New trends in developmental morphology of the thallus. Bibliotheca Lichenologica, 25, 17–33.Google Scholar
Jahns, H. M. (1988). The lichen thallus. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 95–143. Boca Raton: CRC Press.Google Scholar
Jahns, H. M., Tuiz-Dubiel, A. and Blank, L. (1976). Hygroskopische Bewegungen der Sorale von Hypogymnia physodes. Herzogia, 4, 15–23.Google Scholar
James, P. W. and Henssen, A. (1976). The morphological and taxonomic significance of cephalodia. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 27–77. London: Academic Press.Google Scholar
James, P. W., Hawksworth, D. L. and Rose, F. (1977). Lichen communities in the British Isles: a preliminary conspectus. In Lichen Ecology, ed. Seaward, M. R. D., pp. 295–413. London: Academic Press.Google Scholar
Jancey, R. (1966). The application of numerical methods of data analysis to the genus Phyllota (Benth.) in New South Wales. Australian Journal of Botany, 14, 131–149.CrossRefGoogle Scholar
Janex-Favre, M. C. and Ghaleb, M. I. (1986). L'ontogenie et la structure des apothecies du Xanthoria parietina (L.) Beltr. (discolichen). Cryptogamie, Bryologie et Lichenologie, 7, 457–478.Google Scholar
Jansova, I. and Soldan, Z. (2006). The habitat factors that affect the composition of bryophyte and lichen communities on fallen logs. Preslia, 78, 67–86.Google Scholar
Jarmuszkiewicz, W. (2001). Uncoupling proteins in mitochondria of plants and some microorganisms. Acta Biochimica Polonica, 48, 145–155.Google ScholarPubMed
Jennings, D. (1995). The Physiology of Fungal Nutrition. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jensen, M. (2002). Measurement of chlorophyll fluorescence in lichens. In Protocols in Lichenology, ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 135–151. Berlin: Springer.CrossRefGoogle Scholar
Jensen, M. and Kricke, R. (2002). Chlorophyll fluorescence measurements in the field: assessment of the vitality of large numbers of lichen thalli. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 327–332. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Jensen, M., Linke, K., Dickhäuser, A. and Feige, G. B. (1999). The effect of agronomic photosystem-II herbicides on lichens. Lichenologist, 31, 95–103.Google Scholar
John, D. M., Whitton, B. A. and Brook, A. J. (2002). The Freshwater Algal Flora of the British Isles. Cambridge: Cambridge University Press.Google Scholar
John, E. and Dale, M. R. T. (1991). Determinants of spatial pattern in saxicolous lichen communities. Lichenologist, 23, 227–236.CrossRefGoogle Scholar
John, V. (1996). Preliminary catalogue of lichenized and lichenicolous fungi of Mediterranean Turkey. Bocconea, 6, 173–216.Google Scholar
Johnson, L. R. and John, D. M. (1990). Observations on Dilabifilium (class Chlorophyta, order Chaetophorales sensu strictu) and allied genera. British Phycological Journal, 25, 53–61.CrossRefGoogle Scholar
Johnson, P. N. and Galloway, D. J. (2002). Lichens and Their Conservation Needs in New Zealand. Landcare Research Contract Report: LC0102/132. Dunedin: Landcare Research.Google Scholar
Johnston, J. (2001). Siphulella. Flora of Australia, 58A, 22–23.
Jones, D. (1988). Lichens and pedogenesis. In CRC Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 109–124. Boca Raton: CRC Press.Google Scholar
Jones, D., Wilson, M. J. and Laundon, J. R. (1982). Observations on the location and form of lead in Stereocaulon vesuvianum. Lichenologist, 14, 281–286.CrossRefGoogle Scholar
Jørgensen, P. M. (1996). The oceanic element in the Scandinavian lichen flora revisited. Symbolae Botanicae Upsalienses, 31, 297–317.Google Scholar
Jørgensen, P. M. (1997). Further notes on hairy Leptogium species. Symbolae Botanicae Upsalienses, 32, 113–130.Google Scholar
Jørgensen, P. M. (1998). What shall we do with the blue-green counterparts?Lichenologist, 30, 351–356.CrossRefGoogle Scholar
Jørgensen, P. M. (2000 a). On the sorediate counterparts of the lichen Fuscopannaria leucosticta. Bryologist, 103, 104–107.CrossRefGoogle Scholar
Jørgensen, P. M. (2000 b). New or interesting Parmeliella species from the Andes and Central America. Lichenologist, 32, 139–147.CrossRefGoogle Scholar
Jørgensen, P. M. (2001). The lichen genus Erioderma (Pannariaceae) in China and Japan. Annales Botanici Fennici, 38, 259–264.Google Scholar
Jørgensen, P. M. (2005). A new Atlantic species in Fuscopannaria, with a key to its European species. Lichenologist, 37, 221–225.CrossRefGoogle Scholar
Jørgensen, P. M. and Arvidsson, L. (2002). The lichen genus Erioderma (Pannariaceae) in Ecuador and neighbouring countries. Nordic Journal of Botany, 22, 87–114.CrossRefGoogle Scholar
Jørgensen, P. M. and Jahns, H. M. (1987). Muhria, a remarkable new lichen genus from Scandinavia. Notes of the Royal Botanical Garden Edinburgh, 44, 581–599.Google Scholar
Jovan, S. and Carlberg, T. (2007). Nitrogen content of Letharia vulpina tissue from forests of the Sierra Nevada, California: geographic patterns and relationships to ammonia estimates and climate. Environmental Monitoring and Assessment, 129, 243–251.CrossRefGoogle ScholarPubMed
Jovan, S. and McCune, B. (2004). Regional variation in epiphytic macrolichen communities in northern and central California forests. Bryologist, 107, 328–339.CrossRefGoogle Scholar
Jovan, S. and McCune, B. (2005). Air-quality bioindication in the greater Central Valley of California, with epiphytic macrolichen communities. Ecological Applications, 15, 1712–1726.CrossRefGoogle Scholar
Jovan, S. and McCune, B. (2006). Using epiphytic macrolichen communities for biomonitoring ammonia in forests of the Greater Sierra Nevada, California. Water, Air, and Soil Pollution, 170, 69–93.CrossRefGoogle Scholar
Kaasalainen, S. and Rautiainen, M. (2005). Hot spot reflectance signatures of common boreal lichens. Journal of Geophysical Research, 110, 15–24.CrossRefGoogle Scholar
Kaiser, M. A. and Debbrecht, F. J. (1977). Qualitative and quantitative analysis of gas chromatography. In Modern Practice of Gas Chromatography, ed. Grob, R., pp. 151–211. New York: John Wiley.Google Scholar
Kalb, K. (1987). Brasilianische Flechten. 1. Die Gattung Pyxine. Bibliotheca Lichenologica, 24, 1–89.Google Scholar
Kallio, P. (1974). Nitrogen fixation in subarctic lichens. Oikos, 25, 194–198.CrossRefGoogle Scholar
Kallio, P., Suhonen, S. and Kallio, S. (1972). The ecology of nitrogen fixation in Nephroma arcticum and Solorina crocea. Reports from the Kevo Subarctic Research Station, 9, 7–14.Google Scholar
Kallio, S. (1978). On the effect of forest fertilizers on nitrogenase activity in two subarctic lichens. In Environmental Role of Nitrogen-fixing Blue-green Algae and Asymbiotic Bacteria, ed. Granhall, U., pp. 217–224. Stockholm: Swedish Natural Science Research Council.Google Scholar
Kandler, O. (1987). Lichen and conifer recolonization in Munich's cleaner air. In Symposium of the Commission of the European Communities on “Effects of Air Pollution on Terrestrial and Aquatic Ecosystems” 18–22 May 1987, ed. Mathy, P., pp. 1–7. Brussels: European Commission.Google Scholar
Kantvilas, G. (2000). Conservation of Tasmanian lichens. Mitteilungen der Eidgenössischen Forschungsanstalt für Wald, Schnee und Landschaft, 75, 357–367.Google Scholar
Kantvilas, G. (2002). Studies on the lichen genus Siphula Fr. Bibliotheca Lichenologica, 82, 37–53.Google Scholar
Kantvilas, G. (2004). Steinia. Flora of Australia 56A, 1–3.
Kantvilas, G. and Jarman, S. J. (1999). Lichens of rainforest in Tasmania and south-eastern Australia. Flora of Australia Supplementary Series, 9, 1–212.Google Scholar
Kantvilas, G. and Jarman, S. J. (2006). Recovery of lichens after logging: preliminary results from Tasmania's wet forests. Lichenologist, 38, 383–394.CrossRefGoogle Scholar
Kantvilas, G. and McCarthy, P. M. (2004). Hueidea. Flora of Australia 56A, 182–183.
Kantvilas, G., Elix, J. A. and Jarman, S. J. (2002). Tasmanian lichens, identification, distribution and conservation status. I. Parmeliaceae. Flora of Australia Supplementary Series 15, 1–274.Google Scholar
Kappen, L. (1974). Response to extreme environments. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 311–380. New York: Academic Press.Google Scholar
Kappen, L. (1985). Vegetation and ecology of ice-free areas of northern Victoria Land, Antarctica. Polar Biology, 4, 213–225.CrossRefGoogle Scholar
Kappen, L. (1988). Ecophysiological relationships in different climatic regions. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 37–99. Boca Raton: CRC Press.Google Scholar
Kappen, L. (1993). Lichens in the antarctic region. In Antarctic Microbiology, ed. Friedmann, E. I., pp. 433–490. New York: Wiley-Liss.Google Scholar
Kappen, L. (2000). Some aspects of the great success of lichens in Antarctica. Antarctic Science, 12, 314–324.CrossRefGoogle Scholar
Kappen, L. (2004). The diversity of lichens in Antarctica, a review and comments. Bibliotheca Lichenologica, 88, 331–343.Google Scholar
Kappen, L. and Valladares, F. (1999). Opportunistic growth and desiccation tolerance: the ecological success of poikilohydrous autotrophs. In Handbook of Functional Plant Ecology, ed. Pugnaire, F. I. and Valldares, F., pp. 121–194. New York: Marcel Dekker.Google Scholar
Kappen, L., Breuer, M. and Bölter, M. (1991). Ecological and physiological investigations in continental Antarctic cryptogams. 3. Photosynthetic production of Usnea sphacelata: diurnal courses, models, and the effect of photoinhibition. Polar Biology, 11, 393–401.CrossRefGoogle Scholar
Kappen, L., Schroeter, B., Green, T. G. A. and Seppelt, R. D. (1998 a). Chlorophyll a fluorescence and CO2 exchange on Umbilicaria aprina under extreme light stress in the cold. Oecologia, 113, 325–331.CrossRefGoogle ScholarPubMed
Kappen, L., Schroeter, B., Green, T. G. A. and Seppelt, R. D. (1998 b). Microclimatic conditions, meltwater moistening, and the distributional pattern of Buellia frigida on rock in a southern continental Antarctic habitat. Polar Biology, 19, 101–106.CrossRefGoogle Scholar
Kappen, L., Sommerkorn, M. and Schroeter, B. (1995). Carbon acquisition and water relations of lichens in polar regions – potentials and limitations. Lichenologist, 27, 531–545.Google Scholar
Kardish, N., Silberstein, L., Fleminger, G. and Galun, M. (1991). Lectin from the lichen Nephroma laevigatum Ach. Localization and function. Symbiosis, 11, 47–62.Google Scholar
Kärnefelt, I. (1989). Morphology and phylogeny in the Teloschistales. Cryptogamic Botany, 1, 147–203.Google Scholar
Kärnefelt, I. (1990). Evidence of a slow evolutionary change in the speciation of lichens. Bibliotheca Lichenologica, 38, 291–306.Google Scholar
Kärnefelt, E. I. and Thell, A. (1995). Genotypical variation and reproduction in natural populations of Thamnolia. Bibliotheca Lichenologica, 58, 213–234.Google Scholar
Karnieli, A., Kokaly, R., West, N. E. and Clark, R. N. (2001). Remote sensing of biological soil crusts. In Biological Soil Crusts: Structure, Function and Management, ed. Belnap, J. and Lange, O. L., pp. 431–455. Berlin: Springer.Google Scholar
Kauff, F. and Büdel, B. (2005). Ascoma ontogeny and apothecial anatomy in the Gyalectaceae (Ostropales, Ascomycota) support the re-establishment of the Coenogoniaceae. Bryologist, 108, 272–281.CrossRefGoogle Scholar
Kauff, F. and Lutzoni, F. (2002). Phylogeny of the Gyalectales and Ostropales (Ascomycota, Fungi): among and within order relationships based on nuclear ribosomal RNA small and large subunits. Molecular Phylogenetics and Evolution, 25, 138–156.CrossRefGoogle ScholarPubMed
Keller, N. P. and Hohn, T. M. (1997). Metabolic pathway gene clusters in filamentous fungi. Fungal Genetics and Biology, 21, 17–21.CrossRefGoogle ScholarPubMed
Kelly, B. B. and Becker, V. E. (1975). Effects of light intensity and temperature on nitrogen fixation by Lobaria pulmonaria, Sticta weigelii, Leptogium cyanescens and Collema subfurvum. Bryologist, 78, 350–355.CrossRefGoogle Scholar
Kelly, B. C. and Gobas, F. A. P. C. (2001). Bioaccumulation of persistent organic pollutants in lichen-caribou-wolf food chains of Canada's central and western Arctic. Environmental Science and Technology, 35, 325–334.CrossRefGoogle ScholarPubMed
Kerfin, W. and Boger, P. (1982). Light-induced hydrogen evolution by blue-green algae (Cyanobacteria). Physiologia Plantarum, 54, 93–98.CrossRefGoogle Scholar
Kershaw, K. A. (1974). Dependence of the level of nitrogenase activity on the water content of the thallus in Peltigera canina, P. evansiana, P. polydactyla, and P. praetextata. Canadian Journal of Botany, 52, 1423–1427.CrossRefGoogle Scholar
Kershaw, K. A. (1977). Physiological-environmental interactions in lichens. II. The pattern of net photosynthetic acclimation of Peltigera canina (L.) Willd. var. praetextata (Floreke in Somm.) Hue, and P. polydactyla (Neck.) Hoff. New Phytologist, 79, 377–390.CrossRefGoogle Scholar
Kershaw, K. A. (1985). Physiological Ecology of Lichens. Cambridge: Cambridge University Press.Google Scholar
Kershaw, K. A. and Larson, D. W. (1974). Studies on lichen-dominated systems. IX. Topographic influences on microclimate and species distribution. Canadian Journal of Botany, 52, 1935–1945.CrossRefGoogle Scholar
Kershaw, K. A. and Looney, J. H. H. (1985). Quantitative and Dynamic Plant Ecology, 3rd edn. London: Edward Arnold.Google Scholar
Kershaw, K. A. and Rouse, W. R. (1971). Studies on lichen dominated ecosystems. I. The water relation of Cladonia alpestris in spruce-lichen woodland in northern Ontario. Canadian Journal of Botany, 49, 1389–1399.CrossRefGoogle Scholar
Kershaw, M. J. and Talbot, N. J. (1998). Hydrophobins and repellents: proteins with fundamental roles in fungal morphogenesis. Fungal Genetics and Biology, 23, 18–33.CrossRefGoogle ScholarPubMed
Kershaw, M. J., Thornton, C., Wakley, G. and Talbot, N. J. (2005). Four conserved intramolecular disulphide linkages are required for secretion and cell wall localization of a hydrophobin during fungal morphogenesis. Molecular Microbiology, 56, 117–125.CrossRefGoogle ScholarPubMed
Kets, E., Galinski, E., Wit, M., Bont, J. and Heipieper, H. (1996). Mannitol, a novel bacterial compatible solute in Pseudomonas putida S12. Journal of Bacteriology, 178, 6665–6670.CrossRefGoogle ScholarPubMed
Kidd, D. M. and Ritchie, M. G. (2006). Phylogeographic information systems: putting the geography into phylogeography. Journal of Biogeography, 33, 1851–1865.CrossRefGoogle Scholar
Kieft, T. L. and Ruscetti, T. (1990). Characterization of biological ice nuclei from a lichen. Journal of Bacteriology, 172, 3519–3523.CrossRefGoogle ScholarPubMed
Kinoshita, Y. (1993). The Production of Lichen Substances for Pharmaceutical Use by Lichen Tissue Culture. Osaka: Nippon Paint Publication.Google Scholar
Kirk, P. M., Cannon, P. F., David, J. C. and Stalpers, J. A. (2001). Ainsworth and Bisby's Dictionary of the Fungi. 9th edn. Wallingford, UK: CAB International.Google Scholar
Kirkpatrick, R. C., Long, Y. C., Zhong, T. and Xia, L. (1998). Social organization and range use in the Yunnan snub-nosed monkey Rhinopithecus bieti. International Journal of Primatology, 19, 13–51.CrossRefGoogle Scholar
Knops, J. M.H, Nash, T. H. III, Boucher, V. L. and Schlesinger, W. L. (1991). Mineral cycling and epiphytic lichens: implications at the ecosystem level. Lichenologist, 23, 309–321.CrossRefGoogle Scholar
Knops, J. M. H., Nash, T. H. III and Schlesinger, W. H. (1996). The influence of epiphytic lichens on the nutrient cycling of an oak woodland. Ecological Monographs, 66, 159–179.CrossRefGoogle Scholar
Knowles, R. D., Pastor, J. and Biesboer, D. D. (2006). Increased soil nitrogen associated with dinitrogen-fixing, terricolous lichens of the genus Peltigera in northern Minnesota. Oikos, 114, 37–48.CrossRefGoogle Scholar
Koch, J. and Kilian, R. (2005). ‘Little Ice Age’ glacier fluctuations, Gran Campo Nevado, southernmost Chile. Holocene, 15, 20–28.CrossRefGoogle Scholar
Köck, M., Schlee, D. and Metzger, U. (1985). Sulfite-induced changes of oxygen metabolism in the action of superoxide dismutase in Euglena gracilis and Trebouxia sp. Biochemie und Physiologie der Pflanzen, 180, 213–224.CrossRefGoogle Scholar
Koga, S., Echigo, A. and Nunomura, K. (1966). Physical properties of cell water in partially dried Saccharomyces cerevisiae. Biophysics Journal, 6, 665–674.CrossRefGoogle ScholarPubMed
Komárek, J. and Anagnostidis, K. (1998). Cyanoprokaryota. 1. Teil: Chroococcales. Jena: Gustav Fischer.Google Scholar
Komárek, J. and Anagnostidis, K. (2005). Cyanoprokaryota. 2. Teil: Oscillatoriales. München: Elsevier.Google Scholar
Kondratyuk, S. and Kärnefelt, I. (1997). Josefpoeltia and Xanthomendoxa, two new genera in the Teloschistaceae (lichenized Ascomycotina). Bibliotheca Lichenologica, 68, 19–44.Google Scholar
Kong, F. X., Hu, W., Chao, S. Y., Sang, W. L. and Wang, L. S. (1999). Physiological responses of the lichen Xanthoparmelia mexicana to oxidative stress of SO2. Environmental and Experimental Botany, 42, 201–209.CrossRefGoogle Scholar
Kopecky, J., Azarkovich, M., Pfündel, E. E., Shuvalov, V. A. and Heber, U. (2005). Thermal dissipation of light energy is regulated differently and by different mechanisms in lichens and higher plants. Plant Biology, 7, 156–167.CrossRefGoogle ScholarPubMed
Koptsik, S. V., Koptsik, G. N. and Meryashkina, L. V. (2004). Ordination of plant communities in forest biogeocenoses under conditions of air pollution in the northern Kola Peninsula. Russian Journal of Ecology, 35, 190–199.CrossRefGoogle Scholar
Korf, R. P. (1973). Discomycetes and Tuberales. In The Fungi. Vol. IVA: A Taxonomic Review with Keys: Ascomycetes and Fungi Imperfecti, ed. Ainsworth, G. C., Sparrow, F. K. and Sussman, A. S., pp. 249–319. New York: Academic Press.Google Scholar
Kostner, B. and Lange, O. L. (1986). Epiphytische Flechten in bayerischen Waldschadensgebieten des nordlichen Alpenraumes: Floristisch-soziologische Untersuchungen und Vitalitätstests durch Photosynthesemessungen. Berichte der Akademie fur Naturschutz und Landschaftspflege, 10, 185–210.Google Scholar
Kranner, I. and Birtić, S. (2005). A modulating role for antioxidants in desiccation tolerance. Integrative and Comparative Biology, 45, 734–740.CrossRefGoogle ScholarPubMed
Kranner, I. and Grill, D. (1994). Rapid changes of the glutathione status and the enzymes involved in the reduction of glutathione-disulfide during the initial stage of wetting of lichens. Cryptogamic Botany, 4, 203–206.Google Scholar
Kranner, I. and Lutzoni, F. (1999). Evolutionary consequences of transition to a lichen symbiotic state and physiological adaptation to oxidative damage associated with poikilohydry. In Plant Response to Environmental Stress: From Phytohormones to Genome Reorganisation, ed. Lerner, H. R., pp. 591–628. New York: Marcel Dekker.Google Scholar
Kranner, I., Beckett, R., Hochman, A. & Nash, T. H. III (2008). Desiccation tolerance in lichens: a review. Bryologist, 111 (in press).CrossRefGoogle Scholar
Kranner, I., Cram, W. J., Zorn, M., et al. (2005). Antioxidants and photoprotection in a lichen as compared with its isolated symbiotic partners. Proceedings of the National Academy of Sciences, USA, 102, 3141–3146.CrossRefGoogle Scholar
Kranner, I., Zorn, M., Turk, B., et al. (2003). Biochemical traits of lichens differing in relative desiccation tolerance. New Phytologist, 160, 167–176.CrossRefGoogle Scholar
Kroken, S. and Taylor, J. W. (2000). Phylogenetic species, reproductive mode, and specificity of the green alga Trebouxia forming lichens with the fungal genus Letharia. Bryologist, 103, 645–660.CrossRefGoogle Scholar
Kroken, S. and Taylor, J. W. (2001). A gene geneology approach to recognize phylogenetic species boundaries in the lichenized fungus Letharia. Mycologia, 93, 38–53.CrossRefGoogle Scholar
Kroken, S., Glass, N. L., Taylor, J. W., Yoder, O. C. and Turgeon, B. G. (2003). Phylogenomic analysis of type I polyketide synthase genes in pathogenic and saprobic ascomycetes. Proceedings of the National Academy of Sciences, USA, 100, 15 670–15 675.CrossRefGoogle ScholarPubMed
Kurina, L. M. and Vitousek, P. M. (1999). Controls over the accumulation and decline of a nitrogen-fixing lichen, Stereocaulon vulcani on young Hawaiian lava flows. Journal of Ecology, 87, 784–799.CrossRefGoogle Scholar
Kurina, L. M. and Vitousek, P. M. (2001). Nitrogen fixation rates of Stereocaulon vulcani on young Hawaiian lava flows. Biogeochemistry, 55, 179–194.CrossRefGoogle Scholar
Kytöviita, M. M. and Crittenden, P. D. (1994). Effects of simulated acid rain on nitrogenase activity (acetylene reduction) in the lichen Stereocaulon paschale (L.) Hoffm., with special reference to nutritional aspects. New Phytologist, 128, 263–271.CrossRefGoogle Scholar
Kytöviita, M. M. and Crittenden, P. D. (2002). Seasonal variation in growth rate in Stereocaulon paschale. Lichenologist, 34, 533–537.CrossRefGoogle Scholar
Laaksovirta, K. and Olkkonen, H. (1979). Effect of air pollution on epiphytic lichen vegetation and element contents of a lichen and pine needles at Valkeakoski, S. Finland. Annales Botanici Fennici, 16, 285–296.Google Scholar
Laaksovirta, K., Olkkonen, H. and Alakijala, P. (1976). Observations on the lead content of lichen and bark adjacent to a highway in southern Finland. Environmental Pollution, 11, 247–255.CrossRefGoogle Scholar
Lakatos, M. (2002). Ökologische Untersuchungen wuchsformbedingter Verbreitungsmuster von Flechten im tropischen Regenwald. Ph.D. thesis. Kaiserslautern: University of Kaiserslautern.
Lakatos, M., Rascher, U. and Büdel, B. (2006). Functional characteristics of corticolous lichens in the understory of a tropical lowland rain forest. New Phytologist, 172, 679–695.CrossRefGoogle ScholarPubMed
Lambers, H. (1985). Respiration in intact plants and tissues: its regulation and dependence on environmental factors, metabolism and invaded organisms. In Higher Plant Respiration, ed. Douce, R. and Day, D. A., pp. 418–465. Berlin: Springer.CrossRefGoogle Scholar
Lambers, H., Chapin, F. S., III and Pons, T. L. (1998). Photosynthesis, respiration, and long distance transport. In Plant Physiological Ecology, ed. Lambers, H., Chapin, F. S. III and Pons, T. L., pp. 10–95. Berlin: Springer.CrossRefGoogle Scholar
Lang, G. E., Reiners, W. A. and Heier, R. K. (1976). Potential alteration of precipitation chemistry by epiphytic lichens. Oecologia, 25, 229–241.CrossRefGoogle ScholarPubMed
Lang, G. E., Reiners, W. A. and Pike, L. H. (1980). Structure and biomass of epiphytic lichen communities of balsam fir forests in New Hampshire. Ecology, 61, 541–550.CrossRefGoogle Scholar
Lange, O. L. (1953). Hitze- und Trockenresistenz der Flechten in Beziehung zu ihrer Verbreitung. Flora, 140, 39–97.Google Scholar
Lange, O. L. (1965). Der CO2-Gaswechsel von Flechten bei tiefen Temperaturen. Planta, 64, 1–19.CrossRefGoogle Scholar
Lange, O. L. (1969). Experimentell-ökologische Untersuchungen an Flechten der Negev-Wüste. I. CO2-Gaswechsel von Ramalina maciformis (Del.) Bory unter kontrollierten Bedingungen im Laboratorium. Flora, 158, 324–359.Google Scholar
Lange, O. L. (1980). Moisture content and CO2 exchange of lichens. I. Influence of temperature on moisture-dependent net photosynthesis and dark respiration in Ramalina maciformis. Oecologia, 45, 82–87.CrossRefGoogle ScholarPubMed
Lange, O. L. (2002). Photosynthetic productivity of the epilithic lichen Lecanora muralis: long-term field monitoring of CO2 exchange and its physiological interpretation. I. Dependence of photosynthesis on water content, light, temperature, and CO2 concentration from laboratory measurements. Flora, 197, 233–249.CrossRefGoogle Scholar
Lange, O. L. (2003 a). Photosynthetic productivity of the epilithic lichen Lecanora muralis: long-term field monitoring of CO2 exchange and its physiological interpretation. II. Diel and seasonal patterns of net photosynthesis and respiration. Flora, 198, 55–70.Google Scholar
Lange, O. L. (2003 b). Photosynthetic productivity of the epilithic lichen Lecanora muralis: long-term field monitoring of CO2 exchange and its physiological interpretation. III. Diel, seasonal, and annual carbon budgets. Flora, 198, 277–292.Google Scholar
Lange, O. L. and Bertsch, A. (1965). Photosynthese der Wüstenflechte Ramalina maciformis nach Wasserdampfaufnahme aus dem Luftraum. Naturwissenschaften, 52, 215–216.CrossRefGoogle Scholar
Lange, O. L. and Green, T. G. A. (2003). Photosynthetic performance of a foliose lichen of biological soil crust communities: long-term monitoring of the CO2 exchange of Cladonia convoluta under temperate habitat conditions. Bibliotheca Lichenologica, 86, 257–280.Google Scholar
Lange, O. L. and Green, T. G. A. (2005). Lichens show that fungi can acclimate their respiration to seasonal changes in temperature. Oecologia, 142, 11–19.CrossRefGoogle ScholarPubMed
Lange, O. L. and Green, T. G. A. (2006). Nocturnal respiration in lichens in their natural habitat is not affected by preceding diurnal net photosynthesis. Oecologia, 148, 396–404.CrossRefGoogle Scholar
Lange, O. L. and Metzner, H. (1965). Lichtabhängiger Kohlenstoff-Einbau in Flechten bei tiefen Temperaturen. Naturwissenschaften, 52, 191.CrossRefGoogle Scholar
Lange, O. L. and Tenhunen, J. D. (1981). Moisture content and CO2 exchange of lichens. II. Depression of net photosynthesis in Ramalina maciformis at high water content is caused by increased thallus carbon dioxide diffusion resistance. Oecologia, 51, 426–429.CrossRefGoogle Scholar
Lange, O. L. and Wagenitz, G. (2004). Vernon Ahmadjian introduced the term “chlorolichen”. Lichenologist, 36, 171.CrossRefGoogle Scholar
Lange, O. L. and Ziegler, H. (1963). Der Schwermetallgehalt von Flechten aus dem Acarosporetum sinopicae auf Erzschlackenhalden des Harzes. q. Eisen und Kupfer. Mitteilungen der floristischsoziologischen Arbeitsgemeinschaft, neue Folge, 10, 156–183.Google Scholar
Lange, O. L., Bilger, W., Rimke, S. and Schreiber, U. (1989). Chlorophyll fluorescence of lichens containing green and blue-green algae during hydration by water vapor uptake and by addition of liquid water. Botanica Acta, 102, 306–313.CrossRefGoogle Scholar
Lange, O. L., Büdel, B., Heber, U., et al. (1993 a). Temperate rainforest lichens in New Zealand: high thallus water content can severely limit photosynthetic CO2 exchange. Oecologia, 95, 303–313.CrossRefGoogle ScholarPubMed
Lange, O. L., Büdel, B., Meyer, A. and Kilian, E. (1993 b). Further evidence that activation of net photosynthesis by dry cyanobacterial lichens requires liquid water. Lichenologist, 25, 175–189.CrossRefGoogle Scholar
Lange, O. L., Büdel, B., Meyer, A., Zellner, H. and Zotz, G. (2000). Lichen carbon gain under tropical conditions: water relations and CO2 exchange of three Leptogium species of a lower montane rainforest in Panama. Flora, 195, 172–190.CrossRefGoogle Scholar
Lange, O. L., Büdel, B., Zellner, H., Zotz, G. and Meyer, A. (1994). Field measurements of water relations and CO2 exchange of the tropical, cyanobacterial basidiolichen Dictyonema glabratum in a Panamanian rainforest. Botanica Acta, 107, 279–290.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A. and Heber, U. (2001). Hydration-dependent photosynthetic production of lichens: what do laboratory studies tell us about field performance?Journal of Experimental Botany, 52, 2033–2042.CrossRefGoogle ScholarPubMed
Lange, O. L., Green, T. G. A., Melzer, B., Meyer, A. and Zellner, H. (2006). Water relations and CO2 exchange of the terrestrial lichen Teloschistes capensis in the Namib fog desert: measurements during two seasons in the field and under controlled conditions. Flora, 201, 268–280.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A. and Reichenberger, H. (1999 b). The response of lichen photosynthesis to external CO2 concentration and its interaction with thallus water-status. Journal of Plant Physiology, 154, 157–166.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A. and Ziegler, H. (1988). Water status related photosynthesis and carbon isotope discrimination in species of the lichen genus Pseudocyphellaria with green or blue-green photobionts and in photosymbiodemes. Oecologia, 75, 494–501.CrossRefGoogle ScholarPubMed
Lange, O. L., Hahn, S. C., Meyer, A. and Tenhunen, J. D. (1998). Upland tundra in the foothills of the Brooks Range, Alaska, U.S.A.: lichen long-term photosynthetic CO2 uptake and net carbon gain. Arctic and Alpine Research, 30, 252–261.CrossRefGoogle Scholar
Lange, O. L., Kilian, E. and Ziegler, H. (1986). Water vapour uptake and photosynthesis of lichens: performance differences in species with green and blue-green algae as phycobionts. Oecologia, 71, 104–110.CrossRefGoogle Scholar
Lange, O. L., Kilian, E. and Ziegler, H. (1990 b). Photosynthese von Blattflechten mit hygroskopischen Thallusbewegungen bei Befeuchtung durch Wasserdampf oder mit flüssigem Wasser. Bibliotheca Lichenologica, 38, 311–323.Google Scholar
Lange, O. L., Leisner, J. M. R. and Bilger, W. (1999 a). Chlorophyll fluorescence characteristics of the cyanobacterial lichen Peltigera rufescens under field conditions. II. Diel and annual distribution of metabolic activity and possible mechanisms to avoid photoinhibition. Flora, 194, 413–430.CrossRefGoogle Scholar
Lange, O. L., Meyer, A., Zellner, H., Ullmann, I. and Wessels, D. C. J. (1990 a). Eight days in the life of a desert lichen: water relations and photosynthesis of Teloschistes capensis in the coastal fog zone of the Namib Desert. Madoqua, 17, 17–30.Google Scholar
Lange, O. L., Reichenberger, H. and Meyer, A. (1995). High thallus water content and photosynthetic CO2 exchange of lichens. Laboratory experiments with soil crust species from local xerothermic steppe formations in Franconia, Germany. In Flechten Follmann. Contributions to Lichenology in Honour of Gerhard Follmann, ed. Daniëls, F. J. A., Schulz, M., and Peine, J., pp. 139–153. Cologne: Geobotanical and Phytotaxonomical Study Group, University of Cologne.Google Scholar
Lange, O. L., Reichenberger, H. and Walz, H. (1997). Continuous monitoring of CO2 exchange of lichens in the field: short-term enclosure with an automatically operating cuvette. Lichenologist, 29, 259–274.CrossRefGoogle Scholar
Lange, O. L., Tenhunen, J. D., Harley, P. C. and Walz, H. (1985). Method for field measurements of CO2-exchange. The diurnal changes in net photosynthesis and photosynthetic capacity of lichens under mediterranean climatic conditions. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 23–39. New York: Plenum Press.CrossRefGoogle Scholar
Larcher, W. (2003). Physiological Plant Ecology. Berlin: Springer.CrossRefGoogle Scholar
Larocque, S. J. and Smith, D. J. (2004). Calibrated Rhizocarpon spp. growth curve for the Mount Waddington area, British Columbia coast mountains, Canada. Arctic, Antarctic, and Alpine Research, 36, 407–418.CrossRefGoogle Scholar
Larson, D. W. (1983). The pattern of production within individual Umbilicaria lichen thalli. New Phytologist, 94, 409–419.CrossRefGoogle Scholar
Larson, D. W. (1984). Thallus size as a complicating factor in the physiological ecology of lichens. New Phytologist, 97, 87–97.CrossRefGoogle Scholar
Larson, D. W. (1987). The absorption and release of water by lichens. Bibliotheca Lichenologica, 25, 351–360.Google Scholar
Larson, D. W. and Carey, C. K. (1986). Phenotypic variation with “individual” lichen thalli. American Journal of Botany, 73, 214–223.CrossRefGoogle Scholar
Larson, D. W. and Kershaw, K. A. (1974). Acclimation in arctic lichens. Nature, 254, 421–423.CrossRefGoogle Scholar
Larson, D. W. and Kershaw, K. A. (1975). Studies on lichen-dominated systems. XIII. Seasonal and geographical variation of net CO2 exchange of Alectoria ochroleuca. Canadian Journal of Botany, 53, 2598–2607.CrossRefGoogle Scholar
Larson, D. W., Matthes-Sears, U. and Nash, T. H. III (1985). The ecology of Ramalina menziesii. I. Geographical variation in form. Canadian Journal of Botany, 63, 2062–2068.CrossRefGoogle Scholar
Laufer, Z., Beckett, R. P. and Minibayeva, F. V. (2006 a). Co-occurrence of the multicopper oxidases tyrosinase and laccase in lichens in sub-order Peltigerineae. Annals of Botany, 98, 1035–1042.CrossRefGoogle ScholarPubMed
Laufer, Z., Beckett, R. P., Minibayeva, F. V., Lüthje, S. and Böttger, M. (2006 b). Occurrence of laccases in lichenized ascomycetes of the Peltigerineae. Mycological Research, 110, 846–853.CrossRefGoogle ScholarPubMed
Laundon, J. R. (1978). Haematomma chemotypes form fused thalli. Lichenologist, 10, 221–225.CrossRefGoogle Scholar
Lawrey, J. D. (1984). Biology of Lichenized Fungi. New York: Praeger.Google Scholar
Lawrey, J. D. (1986). Biological role of lichen substances. Bryologist, 89, 111–122.CrossRefGoogle Scholar
Lawrey, J. and Diederich, P. (2003). Lichenicolous fungi: interactions, evolution, and biodiversity. Bryologist, 106, 80–120.CrossRefGoogle Scholar
Laxen, D. P. H. and Thompson, M. N. A. (1987). Sulphur dioxide in Greater London, 1931–1985. Environmental Pollution, 43, 103–114.CrossRefGoogle Scholar
LeBlanc, F. and Sloover, J. (1970). Relation between industrialization and the distribution and growth of epiphytic lichens and mosses in Montreal. Canadian Journal of Botany, 48, 1485–1496.CrossRefGoogle Scholar
LeBlanc, F., Rao, D. N. and Comeau, G. (1972). Indices of atmospheric purity and fluoride pollution pattern in Arvida, Quebec. Canadian Journal of Botany, 50, 991–998.CrossRefGoogle Scholar
LeBlanc, F., Robitaille, G. and Rao, D. N. (1974). Biological response of lichens and bryophytes to environmental pollution in the Murdochville Copper Mine area, Quebec. Journal of the Hattori Botanical Laboratory, 38, 405–433.Google Scholar
Lechowicz, M. J. (1982). Ecological trends in lichen photosynthesis. Oecologia, 53, 330–336.CrossRefGoogle ScholarPubMed
Lee, D. E., Lee, W. G. and Mortimer, N. (2001). Where and why have all the flowers gone? Depletion and turnover in the New Zealand Cenozoic angiosperm flora in relation to palaeography and climate. Australian Journal of Botany, 49, 341–356.CrossRefGoogle Scholar
Legaz, M. E., Fontaniella, B., Millanes, A. M. and Vicente, C. (2004). Secreted arginases from phylogenetically far-related lichen species act as cross-recognition factors for two different algal cells. European Journal of Cell Biology, 83, 435–446.CrossRefGoogle Scholar
Leisner, J. M. R., Green, T. G. A. and Lange, O. L. (1997). Photobiont activity of a temperate crustose lichen: long-term chlorophyll fluorescence and CO2 exchange measurements in the field. Symbiosis, 23, 165–182.Google Scholar
Leprince, O., McKersie, B. D. and Hendry, G. A. (1993). The mechanisms of desiccation tolerance in developing seeds. Seed Science Research, 3, 231–246.CrossRefGoogle Scholar
Leuckert, C., Ahmadjian, V., Culberson, C. F. and Johnson, A. (1990). Xanthones and depsidones of the lichen Lecanaora dispersa in nature and of its mycobiont in culture. Mycologia, 82, 370–378.CrossRefGoogle Scholar
Leverenz, J. and Jarvis, P. G. (1979). Photosynthesis in Sitka spruce. VIII. The effects of light flux density and direction on the rate of net photosynthesis and the stomatal conductance of needles. Journal of Applied Ecology, 16, 919–932.CrossRefGoogle Scholar
Leverenz, J. W., Falk, S., Pilström, C.-M. and Samuelsson, G. (1990). The effects of photoinhibition on the photosynthetic light-response curve of green plant cells (Chlamydomonas reinhardtii). Planta, 182, 161–168.CrossRefGoogle Scholar
Lewis, D. H. (1973). Concepts in fungal nutrition and the origin of parasitism and mutualism. Biological Reviews, 48, 261–278.CrossRefGoogle Scholar
Lewis, D. H. and Smith, D. C. (1967). Sugar alcohols (polyols) in fungi and green plants. I. Distribution, physiology and metabolism. New Phytologist, 66, 143–184.CrossRefGoogle Scholar
Lewis, L. A. and McCourt, R. M. (2004). Green algae and the origin of land plants. American Journal of Botany, 91, 1535–1556.CrossRefGoogle ScholarPubMed
Smith, Lewis R. I. (1995). Colonization by lichens and the development of lichen-dominated communities in the maritime Antarctic. Lichenologist, 27, 473–483.CrossRefGoogle Scholar
Lex, M., Silvester, W. B. and Stewart, W. D. P. (1972). Photorespiration and nitrogenase activity in the blue-green alga, Anabaena cylindrica. Proceedings of the Royal Society of London B, 180, 87–102.CrossRefGoogle ScholarPubMed
Li, Y. [M.] (2006). Seasonal variation of diet and food availability in a group of Sichuan snub-nosed monkeys in Shennongjia Nature Reserve, China. American Journal of Primatology, 68, 217–233.Google Scholar
Liberatore, S., Garibotti, G. and Calvelo, S. (2002). Phytogeography of Argentinean lichens. Bibliotheca Lichenologica, 82, 221–234.Google Scholar
Lidén, K. and Gustafsson, M. (1967). Relationships and seasonal variation of 137Cs in lichen, reindeer and man in northern Sweden 1961–1965. In Radioecological Concentration Processes, ed. Aberg, B. and Hungate, F. P., pp. 193–208. Oxford: Pergamon Press.Google Scholar
Lilly, V. G. and Barnett, H. L. (1951). Physiology of the Fungi. New York: McGraw-Hill Book Co.Google Scholar
Lindblom, L. and Ekman, S. (2006). Genetic variation and population differentiation in the lichen-forming ascomycete Xanthoria parietina on the island Storfosna, central Norway. Molecular Ecology, 15, 1545–1559.CrossRefGoogle ScholarPubMed
Linder, M. B., Szilvay, G. R., Nakari-Setälä, T. and Penttilä, M. E. (2005). Hydrophobins: the protein-amphiphiles of filamentous fungi. FEMS Microbiology Reviews, 29, 877–896.CrossRefGoogle ScholarPubMed
Lines, C. E. M., Ratcliffe, R. G., Rees, T. A. V. and Southon, T., E. (1989). A 13C NMR study of photosynthate transport and metabolism in the lichen Xanthoria calcicola Oxner. New Phytologist, 111, 447–456.CrossRefGoogle Scholar
Link, S. O. and Nash, T. H. III (1984 a). Ecophysiological studies of the lichen, Parmelia praesignis Nyl. Population variation and the effect of storage conditions. New Phytologist, 96, 249–256.CrossRefGoogle Scholar
Link, S. O. and Nash, T. H. III (1984 b). A mathematical description of the effect of resaturation on net photosynthesis in the lichen, Parmelia praesignis Nyl. New Phytologist, 96, 257–262.CrossRefGoogle Scholar
Link, S. O., Nash, T. H., III and Driscoll, M. (1985). CO2 exchange in lichens: towards a mechanistic model. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 77–91. New York: Plenum Press.CrossRefGoogle Scholar
Linnaeus, C. (1753). Species plantarum, Vol. 2. Stockholm: L. Salvi.Google Scholar
Lipscomb, D. L., Farris, J. S., Källersjö, M. and Tehler, A. (1998). Support, ribosomal sequences, and the phylogeny of the eukaryotes. Cladistics, 14, 303–38.CrossRefGoogle Scholar
Litterski, B. and Ahti, T. (2004). World distribution of selected European Cladonia species. Symbolae Botanicae Upsalienses, 34, 205–236.Google Scholar
Litterski, B. and Otte, V. (2002). Biogeographical research on European species of selected lichen genera. Bibliotheca Lichenologica, 82, 83–90.Google Scholar
Liu, Y. J. and Hall, B. D. (2004). Body plan evolution of ascomycetes, as inferred from an RNA polymerase II phylogeny. Proceedings of the National Academy of Sciences, USA, 101, 4507–4512.CrossRefGoogle ScholarPubMed
Llimona, X. and Hladun, N. L. (2001). Checklist of the Lichens and lichenicolous Fungi of the Iberian Peninsula and the Balearic Islands. Bocconea, 14, 1–581.Google Scholar
Lodenius, M. and Laaksovirta, K. (1979). Mercury content of Hypogymnia physodes and pine needles affected by a chlor-alkali works in Kuusankoski, SE Finland. Annales Botanici Fennici, 16, 7–10.Google Scholar
Loewus, F. A. (1999). Biosynthesis and metabolism of ascorbic acid in plants and of analogs of ascorbic acid in fungi. Phytochemistry, 52, 193–210.CrossRefGoogle Scholar
Lohtander, K., Oksanen, I. and Rikkinen, J. (2003). Genetic diversity of green algal and cyanobacterial photobionts in Nephroma (Peltigerales). Lichenologist, 4, 325–339.CrossRefGoogle Scholar
Lomolino, M. V., Riddle, B. R. and Brown, J. H. (2006). Biogeography. 3rd edn. Sunderland: Sinauer Associates.Google Scholar
Longton, R. E. (1988). Biology of Polar Bryophytes and Lichens. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lorenzini, G., Landi, U., Loppi, S. and Nali, C. (2003). Lichen distribution and bioindicator tobacco plants give discordant response: a case study from Italy. Environmental Monitoring and Assessment, 82, 243–264.CrossRefGoogle ScholarPubMed
Loso, M. G. and Doak, D. F. (2006). The biology behind lichenometric dating curves. Oecologia, 147, 223–229.CrossRefGoogle ScholarPubMed
Louwhoff, S. H. J. J. (2001). Biogeography of Hypotrachyna, Parmotrema and allied genera (Parmeliaceae) in the Pacific islands. Bibliotheca Lichenologica, 78, 223–246.Google Scholar
Lücking, R. (1997). The use of foliicolous lichens as bioindicators in the tropics, with special reference to the microclimate. Abstracta Botanica, 21, 99–116.Google Scholar
Lücking, R. (2003). Takhtajan's floristic regions and foliicolous lichen biogeography: a compatibility analysis. Lichenologist 35, 33–54.CrossRefGoogle Scholar
Lücking, R. and Bernecker-Lücking, A. (2002). Distance, dynamics, and diversity in tropical rainforests: an experimental approach using foliicolous lichens on artificial leaves. I. Growth performance and succession. Ecotropica, 8, 1–13.Google Scholar
Lücking, R. and Kalb, K. (2001). New Caledonia, foliicolous lichens and island biogeography. Bibliotheca Lichenologica, 78, 247–273.Google Scholar
Lücking, R., Sérusiaux, E. and Vezda, A. (2005). Phylogeny and systematics of the lichen family Gomphillaceae (Ostropales) inferred from cladistic analysis of phenotype data. Lichenologist, 37, 123–170.CrossRefGoogle Scholar
Lücking, R., Wirth, V., Ferraro, L. and Caceres, M. E. S. (2003). Foliicolous lichens from Valdivian temperate rain forest of Chile and Argentina: evidence of an austral element, with the description of seven new taxa. Global Ecology and Biogeography, 12, 21–36.CrossRefGoogle Scholar
Lumbsch, H. T. (1998). The use of metabolic data in lichenology at the species and subspecific levels. Lichenologist, 30, 357–367.CrossRefGoogle Scholar
Lumbsch, H. T. and Elix, J. A. (1985). A new species of the lichen genus Diploschistes from Australia. Plant Systematics and Evolution, 150, 275–279.CrossRefGoogle Scholar
Lumbsch, H. T. and Kothe, H. W. (1988). Anatomical features of Chondropsis semiviridis (Nyl.) Nyl. in relation to its vagrant habit. Lichenologist, 20, 25–29.CrossRefGoogle Scholar
Lumbsch, H. T., Prado, R. and Kantvilas, G. (2005). Gregorella, a new genus to accommodate Moelleropsis humida and a molecular phylogeny of Arctomiaceae. Lichenologist, 37, 291–302.CrossRefGoogle Scholar
Lumbsch, H. T., Schmitt, I., Döring, H. and Wedin, M. (2001 a). ITS sequence data suggest variability of ascus types and support ontogenetic characters as phylogenetic discriminators in the Agyriales (Ascomycota). Mycological Research, 105, 265–274.CrossRefGoogle Scholar
Lumbsch, H. T., Schmitt, I., Döring, H. and Wedin, M. (2001 b). Molecular systematics supports the recognition of an additional order of Ascomycota: the Agyriales. Mycological Research, 105, 16–23.CrossRefGoogle Scholar
Lumbsch, H. T., Schmitt, I., Lücking, R., Wiklund, E. and Wedin, M. (2007). The phylogenetic placement of Ostropales within Lecanoromycetes (Ascomycota) revisited. Mycological Research, 111, 257–267.CrossRefGoogle ScholarPubMed
Lumbsch, H. T., Schmitt, I., Palice, Z., Wiklund, E. and Wedin, M. (2004). Supraordinal phylogenetic relationships of Lecanoromycetes based on a Bayesian analysis of combined nuclear and mitochondrial sequences. Molecular Phylogenetics and Evolution, 31, 822–832.CrossRefGoogle ScholarPubMed
Lumbsch, H. T., Wirtz, N., Lindemuth, R. and Schmitt, I. (2002). Higher level phylogenetic relationships of Euascomycetes (Pezizomycotina) inferred from a combined analysis of nuclear and mitochondrial sequence data. Mycological Progress, 1, 57–70.CrossRefGoogle Scholar
Lutzoni, F., Kauff, F., Cox, C., et al. (2004). Assembling the fungal tree of life: progress, classification, and evolution of subcellular traits. American Journal of Botany, 91, 1446–1480.CrossRefGoogle ScholarPubMed
Lutzoni, F., Pagel, M. and Reeb, V. (2001). Major fungal lineages are derived from lichen symbiotic ancestors. Nature, 411, 937–940.CrossRefGoogle ScholarPubMed
MacCracken, J. G., Alexander, L. E. and Uresk, D. W. (1983). An important lichen of southeastern Montana rangelands. Journal of Range Management, 36, 35–37.CrossRefGoogle Scholar
MacDonald, G. M. (2003). Biogeography: Space, Time, and Life. New York: John Wiley.Google Scholar
MacFarlane, J. D. and Kershaw, K. A. (1977). Physiological-environmental interactions in lichens. IV. Seasonal changes in the nitrogenase activity in Peltigera canina (L.) Willd. var. praetextata (Floerke in Somm.) Hue, and P. canina (L.) Willd. var. rufescens (Weiss) Mudd. New Phytologist, 69, 403–408.CrossRefGoogle Scholar
MacFarlane, J. D. and Kershaw, K. A. (1980). Physiological-environmental interactions in lichens. IX. Thermal stress and lichen ecology. New Phytologist, 84, 669–685.CrossRefGoogle Scholar
MacFarlane, J. D. and Kershaw, K. A. (1982). Physiological-environmental interactions in lichens. XIV. The environmental control of glucose movement from alga to fungus in Peltigera polydactyla, P. rufescens, and Collema furfuraceum. New Phytologist, 91, 93–101.CrossRefGoogle Scholar
MacGinitie, H. (1937). The flora of the Weaverville beds of Trinity County, California, with descriptions of the plant-bearing beds. In Eocene Flora of Western America, pp. 83–151. Publication 465. Washington: Carnegie Institution of Washington.Google Scholar
MacKenzie, T. D. B., Król, M., Huner, N. P. A. and Campbell, D. A. (2002). Seasonal changes in chlorophyll fluorescence quenching and the induction and capacity of the photoprotective xanthophyll cycle in Lobaria pulmonaria. Canadian Journal of Botany, 80, 255–261.CrossRefGoogle Scholar
MacKenzie, T. D. B., MacDonald, T. M., Dubois, L. A. and Campbell, D. A. (2001). Seasonal changes in temperature and light driven acclimation of photosynthetic physiology and macromolecular content in Lobaria pulmonaria. Planta, 214, 57–66.CrossRefGoogle Scholar
Madelin, M. F. (1968). Parasitism on other fungi and lichens. In The Fungi, Vol. III: The Fungal Population, ed. Ainsworth, G. C. and Sussman, A. S., pp. 253–269. New York: Academic Press.Google Scholar
Magan, N. (1997). Fungi in extreme environments. In Environmental and Microbial Relationships, Vol. IV: The Mycota, ed. Wicklow, D. and Soderstrom, B., pp. 99–114. Berlin: Springer.Google Scholar
Mägdefrau, K. (1957). Flechten und Moose im baltischen Bernstein. Berichte der deutschen botanischen Gesellschaft, 9, 433–435.Google Scholar
Máguas, C. and Griffiths, H. (2003). Applications of stable isotopes in plant ecology. Progress in Botany, 64, 472–505.Google Scholar
Máguas, C., Valladares, F. and Brugnoli, E. (1997). Effects of thallus size on morphology and physiology of foliose lichens: new findings with a new approach. Symbiosis, 23, 149–164.Google Scholar
Majerus, M. E. N. (1998). Melanism: Evolution in Action. Oxford: Oxford University Press.Google Scholar
Makkonen, S., Hurri, R. and Hyvärinen, M. (2007). Differential responses of lichen symbionts to enhanced nitrogen and phosphorous availability: an experiment with Cladina stellaris. Annals of Botany, 99, 877–884.CrossRefGoogle ScholarPubMed
Malcolm, W. M. and Galloway, D. J. (1997). New Zealand Lichens. Checklist, Key, and Glossary. Wellington: Museum of New Zealand Te Papa Tongarewa.Google Scholar
Malhotra, S. S. and Khan, A. A. (1983). Sensitivity to SO2 of various metabolic processes in an epiphytic lichen, Evernia mesomorpha. Biochemie und Physiologie der Pflanzen, 178, 121–130.CrossRefGoogle Scholar
Manodori, A. M. and Melis, A. (1984). Photochemical apparatus organization in Anacystis nidulans (Cyanophyceae). Effect of CO2 concentration during cell growth. Plant Physiology, 74, 67–71.CrossRefGoogle Scholar
Margot, J. (1973). Experimental study of the effects of sulphur dioxide on the soredia of Hypogymnia physodes. In Air Pollution and Lichens, ed. Ferry, B. W., Baddeley, M. S. and Hawksworth, D. L., pp. 314–329. Toronto: University of Toronto Press.Google Scholar
Margules, C. R. and Pressey, R. L. (2000). Systematics conservation planning. Nature, 405, 243–253.CrossRefGoogle Scholar
Margulis, L. and Fester, R. (eds.) (1991). Symbiosis as a Source of Evolutionary Innovation: Speciation and Morphogenesis. Cambridge: Massachusetts Institute of Technology Press.Google ScholarPubMed
Marsh, J. E. and Nash, T. H. III (1979). Lichens in relation to the Four Corners Power Plant in New Mexico. Bryologist, 82, 20–28.CrossRefGoogle Scholar
Marshall, W. A. (1996). Aerial dispersal of lichen soredia in the maritime antarctic. New Phytologist, 134, 523–530.CrossRefGoogle Scholar
Marti, J. (1983). Sensitivity of lichen phycobionts to dissolved air pollutants. Canadian Journal of Botany, 61, 1647–1653.CrossRefGoogle Scholar
Martin, D., Ciulla, R. and Roberts, M. (1999). Osmoadaptation in Archaea. Applied and Environmental Microbiology, 65, 1815–1825.Google ScholarPubMed
Masterson, C. L. and Murphy, P. M. (1984). The acetylene reduction technique. In Current Developments in Biological Nitrogen-fixation, ed. Rao, N. S. Subba, pp. 8–33. Baltimore: Edward Arnold.Google Scholar
Matthes-Sears, U. and Nash, T. H. III (1986). The ecology of Ramalina menziesii. V. Estimation of gross carbon and thallus hydration source from diurnal measurements and climatic data. Canadian Journal of Botany, 64, 1698–1702.CrossRefGoogle Scholar
Matthews, J. A. (2005). ‘Little Ice Age’ glacier variations in Jotunheimen, southern Norway: a study in regionally controlled lichenometric dating of recessional moraines with implications for climate and lichen growth rates. Holocene, 15, 1–19.CrossRefGoogle Scholar
Mattox, K. R. and Stewart, K. D. (1984). Classification of the green algae: a concept based on comparative cytology. In Systematics of the Green Algae, ed. Irvine, D. E. G. and John, D. M., pp. 29–72. London: Academic Press.Google Scholar
Mattson, J. E. (1991). Protein banding patterns in some American and European species of Cetraria. Bryologist, 94, 261–269.CrossRefGoogle Scholar
Mayaba, N. and Beckett, R. P. (2001). The effect of desiccation on the activities of antioxidant enzymes in lichens from habitats of contrasting water status. Symbiosis, 31, 113–121.Google Scholar
McCall, K. K. and Martin, C. E. (1991). Chlorophyll concentrations and photosynthesis in three forest understorey mosses in northeastern Kansas. Bryologist, 94, 25–29.CrossRefGoogle Scholar
McCarroll, D. (1993). Modelling late-Holocene snow-avalanche activity, incorporating a new approach to lichenometry. Earth Surface Processes and Landforms, 18, 527–539.CrossRefGoogle Scholar
McCarroll, D., Shakesby, R. and Matthews, J. A. (1998). Spatial and temporal patterns of late Holocene rockfall activity on a Norwegian talus slope: a lichenometric and simulation-modeling approach. Arctic and Alpine Research, 30, 51–60.CrossRefGoogle Scholar
McCarthy, D. P. (1999). A biological basis for lichenometry?Journal of Biogeography, 26, 379–386.CrossRefGoogle Scholar
McCarthy, P. M. (2001). Polyblastia. Flora of Australia, 58A, 171–172.
McCarthy, P. M. (2004). Maronina. Flora of Australia, 56A, 62–63.
McCarthy, P. M. (2006). Checklist of the Lichens of Australia and its Island Territories. Australian Biological Resources Study, Canberra. Version 6 April 2006. Online: www.anbg.gov.au/abrs/lichenlist/introduction.html.
McCarthy, P. M. and Healey, J. A. (1978). Dispersal of lichen propagules by slugs. Lichenologist, 10, 131–132.CrossRefGoogle Scholar
McCune, B. (1988). Lichen communities along O3 and SO2 gradients in Indianapolis. Bryologist, 91, 223–228.CrossRefGoogle Scholar
McCune, B. (1994). Using epiphytic litter to estimate epiphyte biomass. Bryologist, 97, 396–401.CrossRefGoogle Scholar
McCune, B. and Daly, W. J. (1994). Consumption and decomposition of lichen litter in a temperate coniferous rainforest. Lichenologist, 26, 67–71.Google Scholar
McCune, B. and Grace, J. B. (2002). Analysis of Ecological Communities. Gleneden Beach, OR: MjM Software.Google Scholar
McCune, B., Berryman, S. D., Cissel, J. H. and Gitelman, A. I. (2003). Use of a smoother to forecast occurrence of epiphytic lichens under alternative forest management plans. Ecological Applications, 13, 1110–1123.CrossRefGoogle Scholar
McDowall, R. M. (2004). What biogeography is: a place for process. Journal of Biogeography, 31, 345–351.CrossRefGoogle Scholar
McEvoy, M. (2006). Acclimation of the photobiont and mycobiont partners in lichens to high solar radiation. Ph.D. thesis. Ås, Norway: Department of Ecology and Natural Resource Management. Norwegian University of Life Sciences.
McGlone, M. S. (2005). Goodbye Gondwana. Journal of Biogeography, 32, 739–740.CrossRefGoogle Scholar
McGlone, M. S., Duncan, R. P. and Heenan, P. B. (2001). Endemism, species selection and the origin and distribution of the vascular plant flora of New Zealand. Journal of Biogeography, 28, 199–216.CrossRefGoogle Scholar
McKersie, B. D. and Lesham, Y. Y. (1994). Stress and Stress Coping in Cultivated Plants. Dordrecht: Kluwer.CrossRefGoogle Scholar
McNabb, D. H. and Geist, J. M. (1979). Acetylene reduction assay of symbiotic N2 fixation under field conditions. Ecology, 60, 1070–1072.CrossRefGoogle Scholar
Meier, F. A., Scherrer, S. and Honegger, R. (2002). Faecal pellets of lichenivorous mites contain viable cells of the lichen-forming ascomycete Xanthoria parietina and its green algal photobiont, Trebouxia arboricola. Biological Journal of the Linnean Society, 76, 259–268.CrossRefGoogle Scholar
Melkonian, M. (1990). Chlorophyte orders of uncertain affinities: Order Microthamniales. In Handbook of Protoclista, ed. Margulis, L., Corliss, J. O., Melkonian, M. and Chapman, D. J., pp. 652–654. Boston: Jones and Bartlett.Google Scholar
Melkonian, M. and Peveling, E. (1988). Zoospore ultrastructure in species of Trebouxia and Pseudotrebouxia (Chlorophyta). Plant Systematics and Evolution, 158, 183–210.CrossRefGoogle Scholar
Miadlikowska, J. and Lutzoni, F. (2004). Phylogenetic classification of peltigeralean fungi (Peltigerales, Ascomycota) based on ribosomal RNA small and large subunits. American Journal of Botany, 91, 449–464.CrossRefGoogle ScholarPubMed
Miadlikowska, J., Arnold, A. E., Hofstetter, V. and Lutzoni, F. (2004 a). High diversity of cryptic fungi inhabiting healthy lichen thalli in a temperate and tropical forest. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 43. Tartu: Tartu University Press.Google Scholar
Miadlikowska, J., Arnold, A. and Lutzoni, F. (2004 b). Diversity of cryptic fungi inhabiting healthy lichen thalli in a temperate and tropical forest. Ecological Society of America Annual Meeting, 89, 349–350.Google Scholar
Miao, V. P. W., Rabenau, A. and Lee, A. (1997). Cultural and molecular characterization of photobionts of Peltigera membranacea. Lichenologist, 29, 571–587.CrossRefGoogle Scholar
Micallef, A. and Colls, J. J. (1999). Analysis of long-term measurements of airborne concentrations of sulphur dioxide and SO42 − in the rural United Kingdom. Environmental Monitoring and Assessment, 57, 277–290.CrossRefGoogle Scholar
Mies, B. and Printzen, C. (1997). Notes on the lichens of Socotra (Yemen, Indian Ocean). Bibliothecia Lichenologica, 68, 223–239.Google Scholar
Mietzsch, E., Lumbsch, H. T. and Elix, J. A. (1993). Notice: a new computer program for the identification of lichen substances. Mycotaxon, 47, 475–479.Google Scholar
Mikhailova, I. N. and Scheidegger, C. (2001). Early development of Hypogymnia physodes (L.) Nyl. in response to emissions from a copper smelter. Lichenologist, 33, 527–538.CrossRefGoogle Scholar
Millbank, J. W. (1972). Nitrogen metabolism in lichens. IV. The nitrogenase activity of the Nostoc phycobiont in Peltigera canina. New Phytologist, 71, 1–10.CrossRefGoogle Scholar
Millbank, J. W. (1974). Nitrogen metabolism in lichens. V. The forms of nitrogen released by the blue-green phycobiont in Peltigera spp. New Phytologist, 73, 1171–1181.CrossRefGoogle Scholar
Millbank, J. W. (1976). Aspects of nitrogen metabolism in lichens. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 441–455. London: Academic Press.Google Scholar
Millbank, J. W. (1981). The assessment of nitrogen fixation and throughput by lichens. I. The use of a controlled environment chamber to relate acetylene reduction estimates to fixation. New Phytologist, 89, 647–655.CrossRefGoogle Scholar
Millbank, J. W. (1982 a). Nitrogenase and hydrogenase in cyanophilic lichens. New Phytologist, 92, 221–228.CrossRefGoogle Scholar
Millbank, J. W. (1982 b). The assessment of nitrogen fixation and throughput by lichens. III. Losses of nitrogenous compounds by Peltigera membranacea, P. polydactyla and Lobaria pulmonaria in simulated rainfall episodes. New Phytologist, 92, 229–234.CrossRefGoogle Scholar
Millbank, J. W. and Olsen, J. D. (1986). The assessment of nitrogen fixation and throughput by lichens. IV. Nitrogen losses from Peltigera membranacea (Ach.) Nyl. in autumn, winter and spring. New Phytologist, 104, 643–651.CrossRefGoogle Scholar
Miller, G. H. (1973). Variations in lichen growth from direct measurements: preliminary curves for Alectoria minuscula from eastern Baffin Island, N. W. T., Canada. Arctic and Alpine Research, 5, 33–42.CrossRefGoogle Scholar
Miller, J. E. and Brown, D. H. (1999). Studies of ammonia uptake and loss by lichens. Lichenologist, 31, 85–93.Google Scholar
Miller, P. R. and McBride, J. R. (eds.) (1998). Oxidant Air Pollution Impacts in the Montane Forests of Southern California. Ecological Studies 134. New York: Springer.Google Scholar
Miller, P. R., Longbotham, G. J. and Longbotham, C. R. (1983). Sensitivity of selected western conifers to ozone. Plant Disease, 67, 1113–1115.CrossRefGoogle Scholar
Miszalski, Z. and Niewiadomska, E. (1993). Comparison of sulphite oxidation mechanisms in three lichen species. New Phytologist, 123, 345–349.CrossRefGoogle Scholar
Mitchell, R. J., Truscot, A. M., Leith, I. D., et al. (2005). A study of the epiphytic communities of Atlantic oak woods along an atmospheric nitrogen deposition gradient. Journal of Ecology, 93, 482–492.CrossRefGoogle Scholar
Moberg, R. (1990). Waynea, a new lichen genus in the Bacidiaceae from California. Lichenologist, 22, 249–252.CrossRefGoogle Scholar
Modenesi, P. (1993). An SEM study of injury symptoms in Parmotrema reticulatum treated with paraquat or growing in sulphur dioxide-polluted air. Lichenologist, 25, 423–433.CrossRefGoogle Scholar
Mohr, F., Ekman, S. and Heegaard, E. (2004). Evolution and taxonomy of the marine Collemopsidium species (lichenized Ascomycota) in north-west Europe. Mycological Research, 108, 515–532.CrossRefGoogle Scholar
Molina, M. C., Crespo, A., Vicente, C. and Elix, J. A. (2003). Differences in the composition of phenolics and fatty acids of cultured mycobiont and thallus of Physconia distorta. Plant Physiology and Biochemistry, 41, 175–180.CrossRefGoogle Scholar
Mollenhauer, D. (1992). Geosiphon pyriforme. In Algae and Symbioses: Plants, Animals, Fungi, Viruses, Interactions Explored, ed. Reisser, W., pp. 339–351. Bristol: Biopress.Google Scholar
Mollenhauer, D., Mollenhauer, R. and Kluge, M. (1996). Studies on initiation and development of the partner association in Geosiphon pyriforme (Kütz.) v. Wettstein, a unique endocytobiotic system of a fungus (Glomales) and the cyanobacterium Nostoc punctiforme (Kütz.) Hariot. Protoplasma, 193, 3–9.CrossRefGoogle Scholar
Möller, C. and Dreyfuss, M. M. (1996). Microfungi from Antarctic lichens, mosses and vascular plants. Mycologia, 88, 922–933.CrossRefGoogle Scholar
Møller, I. M. (2001). Plant mitochondria and oxidative stress. Electron transport, NADPH turnover and metabolism of reactive oxygen species. Annual Review of Plant Physiology and Plant Molecular Biology, 52, 561–591.CrossRefGoogle ScholarPubMed
Monnet, F., Bordas, F., Deluchat, V., et al. (2005). Use of the aquatic lichen Dermatocarpon luridum as bioindicator of copper pollution: accumulation and cellular distribution tests. Environmental Pollution, 138, 455–461.CrossRefGoogle ScholarPubMed
Montalvo, A. M. and Ellstrand, N. C. (2001). Non-local transplantation and outbreeding depression in the subshrub Lotus scoparius (Fabaceae). American Journal of Botany, 88, 258–269.CrossRefGoogle Scholar
Montieth, J. L. (1977). Climate and the efficiency of crop production in Britain. Philosophical Transactions of the Royal Society of London B, 281, 277–294.CrossRefGoogle Scholar
Morrone, J. J. (2005). Cladistic biogeography: identity and place. Journal of Biogeography, 32, 1281–1284.CrossRefGoogle Scholar
Mosbach, K. (1969). Biosynthesis of lichen substances, products of a symbiotic association. Angewandte Chemie, International Edition, 8, 240–250.CrossRefGoogle Scholar
Moser, T. J., Nash, T. H. III and Clark, W. D. (1980). Effects of a long-term sulfur dioxide fumigation on arctic caribou forage lichens. Canadian Journal of Botany, 58, 2235–2240.CrossRefGoogle Scholar
Moser, T. J., Nash, T. H. III and Link, S. O. (1983 a). Diurnal gross photosynthetic patterns and potential seasonal CO2 assimilation in Cladonia stellaris and Cladonia rangiferina. Canadian Journal of Botany, 61, 642–655.CrossRefGoogle Scholar
Moser, T. J., Nash, T. H. III and Olafsen, A. G. (1983 b). Photosynthetic recovery in arctic caribou forage lichens following a long-term field sulfur dioxide fumigation. Canadian Journal of Botany, 61, 367–370.CrossRefGoogle Scholar
Moxham, T. H. (1980). Lichens and perfume manufacture. Bulletin of the British Lichen Society, 47, 1–2.Google Scholar
Muir, D. C. G., Segstro, M. D., Welbourn, P. M., et al. (1993). Patterns of accumulation of airborne organochlorine contaminants in lichens from the Upper Great Lakes Region of Ontario. Environmental Science and Technology, 27, 1201–1210.CrossRefGoogle Scholar
Muir, P. S., Shirazi, A. M. and Patrie, J. (1997). Seasonal growth dynamics in the lichen Lobaria pulmonaria. Bryologist, 100, 458–464.CrossRefGoogle Scholar
Mukhtar, A., Garty, J. and Galun, M. (1994). Does the lichen alga Trebouxia occur free-living in nature? – Further immunological evidence. Symbiosis, 17, 247–253.Google Scholar
Munne-Bosch, S. and Alegre, L. (2002). The function of tocopherols and tocotrienols in plants. Critical Reviews in Plant Science, 21, 31–57.CrossRefGoogle Scholar
Muñoz, J., Felicísmo, A., Cabezas, F., Burgaz, A. and Martínez, I. (2004). Wind as long-distance dispersal vehicle in the Southern Hemisphere. Science, 304, 1144–1147.CrossRefGoogle ScholarPubMed
Murashige, T. and Skoog, F. (1962). A revised medium for rapid growth and bioassays with tobacco tissue cultures. Physiologia Plantarum, 15, 473–494.CrossRefGoogle Scholar
Murtagh, G. J., Dyer, P. S. and Crittenden, P. D. (2000). Reproductive systems – sex and the single lichen. Nature, 404, 564.CrossRefGoogle Scholar
Murtagh, G. J., Dyer, P. S., McClure, P. C. and Crittenden, P. D. (1999). Use of randomly amplified polymorphic DNA markers as a tool to study variation in lichen- forming fungi. Lichenologist, 31, 257–267.CrossRefGoogle Scholar
Myachi, S., Nakayama, O., Yokohama, , et al. (1989). World Catalogue of Algae, 2nd. edn. Tokyo: Japan Scientific Societies Press.Google Scholar
Myllys, L., Högnabba, F., Lohtander, K., et al. (2005). Phylogenetic relationships of Stereocaulaceae based on simultaneous analysis of beta-tubulin, GAPDH and SSU rDNA sequences. Taxon, 54, 605–618.CrossRefGoogle Scholar
Myllys, L, Stenroos, S., Thell, A. and Ahti, T. (2003). Phylogeny of bipolar Cladonia arbuscula and Cladonia mitis (Lecanorales, Euascomycetes). Molecular Phylogenetics and Evolution, 27, 58–69.CrossRefGoogle Scholar
Naef, A., Roy, B. A., Kaiser, R. and Honegger, R. (2002). Insect-mediated reproduction of systemic infections by Puccinia arrhenatheri on Berberis vulgaris. New Phytologist, 154, 717–730.CrossRefGoogle Scholar
Nakano, T., Handa, S. and Takeshita, S. (1991). Some corticolous algae from the Taishaku-kyô Gorge, western Japan. Nova Hedwigia, 52, 427–451.Google Scholar
Nash, T. H. III (1971). Lichen sensitivity to hydrogen fluoride. Bulletin of the Torrey Botanical Club, 98, 103–106.Google Scholar
Nash, T. H. III (1972). Simplification of the Blue Mountain lichen communities near a zinc factory. Bryologist, 75, 315–324.Google Scholar
Nash, T. H. III (1973). Sensitivity of lichens to sulfur dioxide. Bryologist, 76, 333–339.Google Scholar
Nash, T. H. III (1975). Influence of effluents from a zinc factory on lichens. Ecological Monographs, 45, 183–196.CrossRefGoogle Scholar
Nash, T. H., III (1988). Correlating fumigation studies with field effects. In Lichens, Bryophytes and Air Quality ed. Nash, T. H. III and Wirth, V., pp. 201–216. Bibliotheca Lichenologica 30. Berlin: J. Cramer.Google Scholar
Nash, T. H., III (1989). Metal tolerance in lichens. In Heavy Metal Tolerance in Plants: Evolutionary Aspects, ed. Shaw, A. J., pp. 119–131. Boca Raton: CRC Press.Google Scholar
Nash, T. H., III (1996). Photosynthesis, respiration, productivity and growth. In Lichen Biology, ed. Nash, T. H. III, pp. 88–120. Cambridge: Cambridge University Press.Google Scholar
Nash, T. H. III and Gries, C. (2002). Lichens as bioindicators of sulfur dioxide. Symbiosis, 33, 1–21.Google Scholar
Nash, T. H. III and Lange, O. L. (1988). Responses of lichens to salinity: concentration and time-course relationships and variability among Californian species. New Phytologist, 109, 361–367.CrossRefGoogle Scholar
Nash, T. H III, and Moser, T. J. (1982). Vegetational and physiological patterns of lichens in North American deserts. Journal of the Hattori Botanical Laboratory, 53, 331–336.Google Scholar
Nash, T. H. III, and Olafsen, A. G. (1995). Climate change and the ecophysiological response of Arctic lichens. Lichenologist, 27, 559–565.CrossRefGoogle Scholar
Nash, T. H., III and Riddell, J. (2006). Historical perspectives and new opportunities in the use of lichens as air pollutant monitors. Botany 2006 Symposium Abstracts, p. 51. St. Louis: Botanical Society of America.
Nash, T. H. III and Sigal, L. L. (1979). Gross photosynthetic response of lichens to short-term ozone fumigations. Bryologist, 82, 280–285.Google Scholar
Nash, T. H., III and Sigal, L. L. (1980). Sensitivity of lichens to air pollution with an emphasis on oxidant air pollutants. In Proceedings of the Symposium on Effects of Air Pollution on Mediterranean and Temperate Forest Ecosystems, June 22–27, 1980, Riverside, California, U.S.A. Gen. Tech. Rep. PSW-43, ed. Miller, P. R. (prin. coord.), pp. 117–124. Berkley: Pacific Southwest Forest and Range Experiment Station, Forest Service, U.S. Department of Agriculture.Google Scholar
Nash, T. H. III, Moser, T. J. and Link, S. O. (1980). Nonrandom variation of gas exchange within arctic lichens. Canadian Journal of Botany, 58, 1181–1186.CrossRefGoogle Scholar
Nash, T. H. III, Moser, T. J., Link, S. O., et al. (1983). Lichen photosynthesis in relation to CO2 concentration. Oecologia, 58, 52–56.CrossRefGoogle ScholarPubMed
Nash, T. H. III, Reiner, A., Demmig-Adams, B., et al. (1990). The effect of atmospheric desiccation and osmotic water stress on photosynthesis and dark respiration of lichens. New Phytologist, 116, 269–276.CrossRefGoogle Scholar
Nash, T. H. III, Ryan, B. D., Diederich, P., Gries, C. and Bungartz, F., eds., (2004). Lichen Flora of the Greater Sonoran Desert Region. Vol. 2. Tempe: Lichens Unlimited.Google Scholar
Nash, T. H. III, Ryan, B. D., Gries, C. and Bungartz, F. (2002). Lichen Flora of the Greater Sonoran Desert Region. Vol. 1. Tempe: Lichens Unlimited.Google Scholar
Nash, T. H., III, Thomas, M. A., Hoober, J. K., Gries, C. and Zheng, S. X. (2001). Free amino acids in lichens and their symbionts. In Lichenological Contributions in Honour of Jack Elix, ed. McCarthy, P. M., Kantvilas, G. and Louwhoff, S. H. J. J., pp. 313–319. Bibliotheca Lichenologica 78. Berlin: J. Cramer.Google Scholar
Nash, T. H. III, White, S. L. and Marsh, J. E. (1977). Lichen and moss distribution and biomass in hot desert ecosystems. Bryologist, 80, 470–479.Google Scholar
Nathan, R. (2005). Long-distance dispersal research: building a network of yellow brick roads. Diversity and Distributions 11, 125–130.CrossRefGoogle Scholar
Nevo, E., Apelbaum-Elkahar, I., Garty, J. and Beiles, A. (1997). Natural selection caused microscale allozyme diversity in wild barley and a lichen at “Evolution Canyon”, Mt. Carmel, Israel. Heredity, 78, 373–382.CrossRefGoogle Scholar
Newmaster, S. T., Bell, F. W. and Vitt, D. H. (1999). The effects of glyphosate and triclopyr on common bryophytes and lichens in northwestern Ontario. Canadian Journal of Forest Research, 29, 1101–1111.CrossRefGoogle Scholar
Newsham, K. K., Low, M. N. R., McLeod, A. R., Greenslade, P. D. and Emmett, B. C. (1997). Ultraviolet-B radiation influences the abundance and distribution of phylloplane fungi on pedunculate oak (Quercus robur). New Phytologist, 136, 287–297.CrossRefGoogle Scholar
Nicholson, T. P., Rudd, B. A. M., Dawson, M., et al. (2001). Design and utility of oligonucleotide gene probes for fungal polyketide synthases. Chemistry and Biology, 8, 57–178.CrossRefGoogle ScholarPubMed
Nieboer, E. and Richardson, D. H. S. (1980). The replacement of the nondescript term ‘heavy metals’ by a biologically and chemically significant classification of metal ions. Environmental Pollution, 1, 3–26.Google Scholar
Nieboer, E. and Richardson, D. H. S. (1981). Lichens as monitors of atmospheric deposition. In Atmospheric Pollutants in Natural Waters, ed. Eisenreich, S. J., pp. 339–388. Ann Arbor: Ann Arbor Science.Google Scholar
Nieboer, E., MacFarlane, J. D. and Richardson, D. H. S. (1984). Modification of plant cell buffering capacities by gaseous air pollutants. In Gaseous Air Pollutants and Plant Metabolism, ed. Kozsol, M. J. and Whatley, F. R., pp. 313–330. London: Butterworths.Google Scholar
Nieboer, E., Richardson, D. H. S., Lavoie, P. and Padovan, D. (1979). The role of metal-ion binding in modifying the toxic effects of sulphur dioxide on the lichen Umbilicaria muhlenbergii. I. Potassium efflux studies. New Phytologist, 82, 621–632.CrossRefGoogle Scholar
Nieboer, E., Richardson, D. H. S., Puckett, K. J. and Tomassini, F. D. (1976). The phytotoxicity of sulphur dioxide in relation to measurable responses in lichens. In Effects of Air Pollutants on Plants, ed. Mansfield, T. A., pp. 61–85. Cambridge: Cambridge University Press.Google Scholar
Nieboer, E., Richardson, D. H. S. and Tomassini, F. D. (1978). Mineral uptake and release by lichens: an overview. Bryologist, 81, 226–246.CrossRefGoogle Scholar
Nieboer, E., Tomassini, F. D., Puckett, K. J. and Richardson, D. H. S. (1977). A model for the relationship between gaseous and aqueous concentrations of sulphur dioxide in lichen exposure studies. New Phytologist, 79, 157–162.CrossRefGoogle Scholar
Nikonov, A. A. and Shebalina, T. Y. (1979). Lichenometry and earthquake age determination in central Asia. Nature, 280, 675–677.CrossRefGoogle Scholar
Nimis, P. L. (1996). Towards a checklist of Mediterranean lichens. Bocconea, 6, 5–17.Google Scholar
Nimis, P. L. and Martellos, S. (2003). A second checklist of the lichens of Italy, with a thesaurus of synonyms. Museo Regionale di Scienze Naturali Monografie, 4, 1–192.Google Scholar
Nimis, P. L. and Poelt, J. (1987). The lichens and lichenicolous fungi of Sardinia (Italy). An annotated list. Studia Geobotanica Trieste, 7 (suppl.1), 1–269.Google Scholar
Nimis, P. L., Castello, M. and Perotti, M. (1990). Lichens as biomonitors of sulphur dioxide pollution in La Spezia (Northern Italy). Lichenologist, 22, 333–344.CrossRefGoogle Scholar
Nimis, P. L., Pinna, D. and Salvadori, O. (1992). Licheni e Conservazione dei Monumenti. Bologna: Cooperativa Libraria Universitaria Editrice Bologna.Google Scholar
Nimis, P. L., Scheidegger, C. and Wolseley, P. A., eds. (2002). Monitoring with Lichens – Monitoring Lichens. Nato Science Series IV: Earth and Environmental Sciences 7. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Noctor, G. and Foyer, C. (1998). Ascorbate and glutathione: keeping active oxygen under control. Annual Review of Plant Physiology and Plant Molecular Biology, 49, 249–279.CrossRefGoogle ScholarPubMed
Noeske, O., Läuchli, A., Lange, O. L., Vieweg, G. H. and Ziegler, H. (1970). Konzentration und Kokalisierung von Schwermetallen in Flechten der Erzschlackenhalden des Harzes. Berichte der deutschen botanischen Gesellschaft, 4, 67–79.Google Scholar
Nordberg, M. L. and Allard, A. (2002). A remote sensing methodology for monitoring cover. Canadian Journal of Remote Sensing, 28, 262–274.CrossRefGoogle Scholar
Notcutt, G. and Davies, F. (1993). Dispersion of gaseous volcanogenic fluoride, island of Hawaii. Journal of Volcanology and Geothermal Research, 56, 125–131.CrossRefGoogle Scholar
Notcutt, G. and Davies, F. (1999). Biomonitoring of volcanogenic fluoride, Furnas Caldera, Sao Miguel, Azores. Journal of Volcanology and Geothermal Research, 92, 209–214.CrossRefGoogle Scholar
Nriagu, J. O. and Pacyna, J. (1988). Quantitative assessment of worldwide contamination of air, water and soils by trace metals. Nature, 333, 134–139.CrossRefGoogle ScholarPubMed
Nyati, S. (2006). Photobiont diversity in Teloschistaceae. Ph.D. thesis, Mathematisch-Naturwissenschaftliche Fakultät, Universität Zürich.
Nybakken, L., Solhaug, K. A., Bilger, W. and Gauslaa, Y. (2004). The lichens Xanthoria elegans and Cetraria islandica maintain a high protection against UV-B radiation in Arctic habitats. Oecologia, 140, 211–216.CrossRefGoogle ScholarPubMed
Nylander, W. (1866). Circa novum in studio Lichenum criterium chemicum. Flora, 49, 198–201.Google Scholar
Obermayer, W. and Poelt, J. (1992). Contributions to the knowledge of the lichen flora of the Himalayas. III. On Lecanora somervellii Paulson (lichenised Ascomycotina, Lecanoraceae). Lichenologist, 24, 111–117.Google Scholar
Oberwinkler, F. (1984). Fungus-alga interactions in basidiolichens. Nova Hedwigia, 79, 739–774.Google Scholar
Oberwinkler, F. (2001). Basidiolichens. In Fungal Associations, The Mycota, Vol. IX, ed. Hock, B., pp. 211–225. Berlin: Springer.CrossRefGoogle Scholar
O'Brien, H., Miadlikowska, J. and Lutzoni, F. (2005). Assessing host specialization in symbiotic cyanobacteria associated with four closely related species of the lichen fungus Peltigera. European Journal of Phycology, 40, 363–378.CrossRefGoogle Scholar
Ochiai, E. (1977). Bioinorganic Chemistry: An Introduction. Boston: Allyn and Bacon.Google Scholar
Ockinger, E., Niklasson, M. and Nilsson, S. G. (2005). Is local distribution of the epiphytic lichen Lobaria pulmonaria limited by dispersal capacity or habitat quality?Biodiversity and Conservation, 14, 759–773.CrossRefGoogle Scholar
Ogren, E. (1993). Convexity of the photosynthetic light-response curve in relation to intensity and direction of light during growth. Plant Physiology, 101, 1013–1019.CrossRefGoogle ScholarPubMed
O'Hare, G. P. and Williams, P. (1975). Some effects of sulphur dioxide flow on lichens. Lichenologist, 7, 116–120.CrossRefGoogle Scholar
Ohmura, Y., Kawachi, M., Kasai, F. and Watanabe, M. (2006). Genetic combinations of symbionts in a vegetatively reproducing lichen, Parmotrema tinctorum, based on ITS rDNA sequences. Bryologist, 109, 43–59.CrossRefGoogle Scholar
Olafsen, A. G. (1989). Nitrogen and carbon fixation in two Arctic lichens, Stereocaulon tomentosum and Peltigera canina. M.S. thesis. Tempe: Arizona State University.
Oliver, A. E., Hinchab, D. K. and Crowe, J. H. (2002). Looking beyond sugars: the role of amphiphilic solutes in preventing adventitious reactions in anhydrobiotes at low water contents. Comparative Biochemistry and Physiology Part A, 131, 515–525.CrossRefGoogle ScholarPubMed
Oliver, A. E., Leprince, O., Wolkers, W. W., et al. (2001). Non-disaccharide-based mechanisms of protection during drying. Cryobiology, 43, 151–167.CrossRefGoogle ScholarPubMed
Olmez, I., Gulovali, M. C. and Gordon, G. E. (1985). Trace element concentrations in lichens near a coal-fired power plant. Atmospheric Environment, 19, 1663–1669.CrossRefGoogle Scholar
O'Neill, A. L. (1994). Reflectance spectra of microphytic soil crusts in semi-arid Australia. International Journal of Remote Sensing, 15, 675–681.CrossRefGoogle Scholar
Öquist, G., Brunes, L. and Hällgren, J.-E. (1982). Photosynthetic efficiency of Betula pendula acclimated to different quantum flux densities. Plant Cell and Environment, 5, 9–15.Google Scholar
Orange, A., James, P. W. and White, F. J. (2001). Microchemical Methods for the Identification of Lichens. London: British Lichen Society.Google Scholar
Ott, S. (1987 a). The juvenile development of lichen thalli from vegetative diaspores. Symbiosis, 3, 57–74.Google Scholar
Ott, S. (1987 b). Reproductive strategies in lichens. Bibliotheca Lichenologica, 25, 81–93.Google Scholar
Ott, S. and Zvoch, I. (1992). Ethylene production by lichens. Lichenologist, 24, 73–80.Google Scholar
Ott, S., Krieg, T., Spanier, U. and Schieleit, P. (2000 a). Phytohormones in lichens with emphasis on ethylene biosynthesis and functional aspects on lichen symbiosis. Phyton, 40, 83–94.Google Scholar
Ott, S., Schröder, T. and Jahns, H. M. (2000 b). Colonization strategies and interactions of lichens on twigs. Bibliotheca Lichenologica, 75, 445–455.Google Scholar
Ott, S., Treiber, K. and Jahns, H. M. (1993). The development of regenerative thallus structures. Botanical Journal of the Linnean Society, 113, 61–76.CrossRefGoogle Scholar
Otte, V., Esslinger, T. L. and Litterski, B. (2002). Biogeographical research on European species of the lichen genus Physconia. Journal of Biogeography, 29, 1125–1141.CrossRefGoogle Scholar
Otte, V., Esslinger, T. L. and Litterski, B. (2005). Global distribution of the European species of the lichen genus Melanelia Essl. Journal of Biogeography, 32, 1221–1241.CrossRefGoogle Scholar
Øvstedal, D. O. and Smith, Lewis R. I. (2001). The Lichens of Antarctica and South Georgia: A Guide to their Identification and Ecology. Cambridge: Cambridge University Press.Google Scholar
Pakarinen, P. and Häsänen, E. (1983). Mercury concentrations of bog mosses and lichens. Suo, 34, 17–20.Google Scholar
Palacios, D., Parrilla, G. and Zamorano, J. J. (1999). Paraglacial and postglacial debris flows on a Little Ice Age terminal moraine: Jamapa Glacier, Pico de Orizaba (Mexico). Geomorphology, 28, 95–118.CrossRefGoogle Scholar
Palice, Z. and Printzen, C. (2004). Genetic variability in tropical and temperate populations of Trapeliopsis glaucolepidea: evidence against long-range dispersal in a lichen with a disjunct distribution. Mycotaxon, 90, 43–54.Google Scholar
Palmer, H. E., Hanson, W. C., Griffin, B. I. and Roesch, W. C. (1963). Cesium-137 in Alaskan eskimos. Science, 142, 64–66.CrossRefGoogle ScholarPubMed
Palmer, R. J. Jr. and Friedmann, E. I. (1990). Water relations, thallus structure and photosynthesis in Negev Desert lichens. New Phytologist, 116, 597–603.CrossRefGoogle ScholarPubMed
Palmqvist, K. (1993). Photosynthetic CO2 use efficiency in lichens and their isolated photobionts: the possible role of a CO2 concentrating mechanism in cyanobacterial lichens. Planta, 191, 48–56.CrossRefGoogle Scholar
Palmqvist, K. (2000). Carbon economy in lichens. New Phytologist, 148, 11–36.CrossRefGoogle Scholar
Palmqvist, K. (2002). Carbon metabolism in cyanobacterial lichens. In Cyanobacteria in Symbiosis, ed. Rai, A. N., and, B. BergmanRasmussen, U., pp. 73–96. Amsterdam: Kluwer Academic.CrossRefGoogle Scholar
Palmqvist, K. and Dahlman, L. (2006). Responses of the green algal foliose lichen Platismatia glauca to increased nitrogen supply. New Phytologist, 171, 343–356.CrossRefGoogle ScholarPubMed
Palmqvist, K. and Sundberg, B. (2000). Light use efficiency of dry matter gain in five macro-lichens: relative impact of microclimate and species-specific traits. Plant Cell and Environment, 23, 1–14.CrossRefGoogle Scholar
Palmqvist, K., los Ríos, A., Ascaso, C. and Samuelsson, G. (1997). Photosynthetic carbon acquisition in the lichen photobionts Coccomyxa and Trebouxia (Chlorophyta). Physiologia Plantarum, 101, 67–76.CrossRefGoogle Scholar
Palmqvist, K., Campbell, D., Ekblad, A. and Johansson, H. (1998). Photosynthetic capacity in relation to nitrogen content and its partitioning in lichens with different photobionts. Plant, Cell and Environment, 21, 361–372.CrossRefGoogle Scholar
Palmqvist, K., Dahlman, L., Valladares, F., et al. (2002). CO2 exchange and thallus nitrogen across 75 contrasting lichen associations from different climate zones. Oecologia, 133, 295–306.CrossRefGoogle ScholarPubMed
Palmqvist, K., Samuelsson, G. and Badger, M. R. (1994). Photobiont-related differences in carbon acquisition among green-algal lichens. Planta, 195, 70–79.CrossRefGoogle Scholar
Pannewitz, S., Green, T. G. A., Schlensog, M., et al. (2006). Photosynthetic performance of Xanthoria mawsonii C. W. Dodge in coastal habitats, Ross Sea region, continental Antarctica. Lichenologist, 38, 67–81.CrossRefGoogle Scholar
Pannewitz, S., Schlensog, M., Green, T. G. A., Sancho, L. G. and Schroeter, B. (2003). Are lichens active under snow in continental Antarctica?Oecologia, 135, 30–38.CrossRefGoogle ScholarPubMed
Parguey-Leduc, A. and Janex-Favre, M. (1981). The ascocarps of ascohymenial pyrenomycetes. In Ascomycete Systematics: The Luttrellian Concept, ed. Reynolds, D. R., pp. 102–123. New York: Springer.CrossRefGoogle Scholar
Parmasto, E. (2004). Integrating molecular and morphological data in the systematics of fungi. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 5. Tartu: Tartu University Press.Google Scholar
Peake, J. F. and James, P. W. (1967). Lichens and mollusca. Lichenologist, 3, 425–428.CrossRefGoogle Scholar
Pearce, C. (1997). Biologically active fungal metabolites. Advances in Applied Microbiology, 44, 1–80.CrossRefGoogle ScholarPubMed
Pearson, L. C. and Henriksson, E. (1981). Air pollution damage to cell membranes in lichens. II. Laboratory experiments. Bryologist, 84, 515–520.CrossRefGoogle Scholar
Pearson, L. C. and Skye, E. (1965). Air pollution affects patterns of photosynthesis in Parmelia sulcata, a corticolous lichen. Science, 148, 1600–1602.CrossRefGoogle Scholar
Peiser, G. and Yang, S. F. (1985). Biochemical and physiological effects of SO2 on nonphotosynthetic processes in plants. In Sulfur Dioxide and Vegetation: Physiology, Ecology, and Policy Issues, ed. Winner, W. E., Mooney, H. A. and Goldstein, R. A., pp. 148–161. Stanford: Stanford University Press.Google Scholar
Pennycook, S. R. and Galloway, D. J. (2004). Checklist of New Zealand “Fungi”. In Fungi of New Zealand, Vol. 1: Introduction to Fungi of New Zealand, ed. McKenzie, E. H. C., pp. 401–488. Hong Kong: Fungal Diversity Press.Google Scholar
Pérez, F. L. (1994). Vagrant cryptogams in a paramo of the high Venezuelan Andes. Flora, 189, 263–276.CrossRefGoogle Scholar
Perkins, D. F. (1992). Relationship between fluoride contents and loss of lichens near an aluminium works. Water, Air, and Soil Pollution, 64, 503–510.CrossRefGoogle Scholar
Perkins, D. F. and Millar, R. O. (1987). Effects of airborne fluoride emissions near an aluminium works in Wales: part 2 – saxicolous lichens growing on rocks and walls. Environmental Pollution, 48, 185–196.CrossRefGoogle ScholarPubMed
Peršoh, D. (2004). Diversity of lichen inhabiting fungi in the Letharietum vulpinae. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 34. Tartu: Tartu University Press.Google Scholar
Peršoh, D., Beck, A. and Rambold, G. (2004). The distribution of ascus types and photobiontal selection in Lecanoromycetes (Ascomycota) against the background of a revised SSU nrDNA phylogeny. Mycological Progress, 3, 103–121.CrossRefGoogle Scholar
Peterson, E. B. (2000). An overlooked fossil lichen (Lobariaceae). Lichenologist, 32, 298–300.CrossRefGoogle Scholar
Peterson, E. B. and McCune, B. (2003). The importance of hotspots for lichen diversity in forests of western Oregon. Bryologist, 106, 246–256.CrossRefGoogle Scholar
Peterson, R. B. and Burris, R. H. (1976). Conversion of acetylene reduction rates in natural populations of blue-green algae. Analytical Biochemistry, 73, 404–410.CrossRefGoogle ScholarPubMed
Petrini, O., Hake, U. and Dreyfuss., M. M. (1990). An analysis of fungal communities isolated from fruticose lichens. Mycologia, 82, 444–451.CrossRefGoogle Scholar
Petzold, D. E. and Goward, S. N. (1988). Reflectance spectra of subarctic lichens. Remote Sensing of Environment, 24, 481–492.CrossRefGoogle Scholar
Pfanz, H., Martinoia, E, Lange, O. L. and Heber, U. (1987). Flux of SO2 into leaf cells and cellular acidification by SO2. Plant Physiology, 85, 928–933.CrossRefGoogle Scholar
Piercey-Normore, M. D. (2004). Selection of algal genotypes by three species of lichen fungi in the genus Cladonia. Canadian Journal of Botany, 82, 947–961.CrossRefGoogle Scholar
Piercey-Normore, M. D. (2006). The lichen-forming ascomycete Evernia mesomorpha associates with multiple genotypes of Trebouxia jamesii. New Phytologist, 169, 331–344.CrossRefGoogle ScholarPubMed
Piercey-Normore, M. D. and DePriest, P. T. (2001). Algal switching among lichen symbioses. American Journal of Botany, 88, 1490–1498.CrossRefGoogle ScholarPubMed
Pike, L. H. (1978). The importance of epiphytic lichens in mineral cycling. Bryologist, 81, 247–257.CrossRefGoogle Scholar
Pike, L. H. (1981). Estimation of lichen biomass and production with special reference to the use of ratios. In The Fungal Community: Its Organization and Role in the Ecosystem, ed. Wicklow, D. and Carroll, G., pp. 533–552. New York: Marcel Dekker.Google Scholar
Pinna, D., Salvadori, O. and Tretiach, M. (1998). An anatomical investigation of calcicolous endolithic lichens from the Trieste karst (NE Italy). Plant Biosystems, 132, 183–195.CrossRefGoogle Scholar
Pintado, A. and Sancho, L. G. (2002). Ecological significance of net photosynthesis activation by water vapour uptake in Ramalina capitata from rain-protected habitats in central Spain. Lichenologist, 34, 403–413.CrossRefGoogle Scholar
Pišút, I. and Lisicka, E. (1985). A study of cryptogamic epiphytes on an oak trunk in the vicinity of Bratislava in the years 1973–1983. Ekologia (CSSR), 4, 225–234.Google Scholar
Plakunova, O. V. and Plakunova, V. G. (1987). Ultrastructure of Cladina stellaris lichen components in normal conditions and at SO2 environment pollution. Izvestiya Akademii Nauk SSR, Seriya Biologicheskaya, 1987, 361–369.Google Scholar
Platt, J. L. and Spatafora, J. W. (2000). Evolutionary relationships of nonsexual lichenized fungi: molecular phylogenetic hypotheses for the genera Siphula and Thamnolia from SSU and LSU rDNA. Mycologia, 92, 475–487.CrossRefGoogle Scholar
Plessl, A. (1963). Über die Beziehungen von Pilz und Alge im Flechtenthallus. Österreichische Botanische Zeitschrift, 110, 194–269.CrossRefGoogle Scholar
Poelt, J. (1970). Das Konzept der Artenpaare bei den Flechten. Vorträge aus dem Gesamtgebeit der Botanik herausgegeben von der deutschen botanischen Gesellschaft, Berlin, 4, 187–198.Google Scholar
Poelt, J. (1972). Die Taxonomische Behandlung von Artenpaaren bei den Flechten. Botaniska Notiser, 125, 77–81.Google Scholar
Poelt, J. (1980). Physcia opuntiella und die Lebensform der sprossenden Flechten. Flora, 169, 22–31.CrossRefGoogle Scholar
Poelt, J. (1985). Über auf Moosen parasitierende Flechten. Sydowia, Annales Mycologici Series II, 38, 241–254.Google Scholar
Poelt, J. (1986). Morphologie der Flechten – Fortschritte und Probleme. Berichte der deutschen botanischen Gesellschaft, 99, 3–29.Google Scholar
Poelt, J. (1989). Die Entstehung einer Strauchflechte aus einem Formenkreis krustiger Verwandter. Flora, 183, 65–72.CrossRefGoogle Scholar
Poelt, J. (1993). La riproduzione asessuale nei licheni. Notiziario della Società Lichenologica Italiana, 6, 9–28.Google Scholar
Poelt, J. (1994). On lichenized asexual diaspores in foliose lichens – a contribution towards a more differentiated nomenclature (Lichens, Lecanorales). Cryptogamic Botany, 5, 150–162.Google Scholar
Poelt, J. and Doppelbauer, H. (1956). Über parasitische Flechten. Planta, 46, 467–480.CrossRefGoogle Scholar
Poelt, J. and Mayrhofer, H. (1988). Über Cyanotrophie bei Flechten. Plant Systematics and Evolution, 158, 265–281.CrossRefGoogle Scholar
Poelt, J. and Obermayer, W. (1990 a). Über Thallosporen bei einigen Krustenflechten. Herzogia, 8, 273–288.Google Scholar
Poelt, J. and Obermayer, W. (1990 b). Lichenisierte Bulbillen als Diasporen bei der Basidiolichene Multiclavula vernalis spec. coll. Herzogia, 6, 289–294.Google Scholar
Poelt, J. and Vězda, A. (1990). Über kurzlebige Flechten. Bibliotheca Lichenologica, 38, 377–394.Google Scholar
Pöggeler, S. (1999). Phylogenetic relationships between mating-type sequences from homothallic and heterothallic ascomycetes. Current Genetics, 36, 222–231.Google ScholarPubMed
Poinar, G., Peterson, E. and Platt, J. (2000). Fossil Parmelia in New World amber. Lichenologist, 32, 263–269.CrossRefGoogle Scholar
Pole, M. (1993). Keeping in touch: vegetation prehistory on both sides of the Tasman. Australian Systematic Botany, 6, 387–397.CrossRefGoogle Scholar
Pole, M. (1994). The New Zealand flora – entirely long-distance dispersal?Journal of Biogeography, 21, 625–635.CrossRefGoogle Scholar
Ponzetti, J. M. and McCune, B. P. (2001). Biotic soil crusts of Oregon's shrub steppe: community composition in relation to soil chemistry, climate, and livestock activity. Bryologist, 104, 212–225.CrossRefGoogle Scholar
Popp, M. and Smirnoff, N. (1995). Polyol accumulation and metabolism during water deficit. In Environment and Plant Metabolism, ed. Smirnoff, N., pp. 199–215. Oxford: BIOS Scientific Publishers.Google Scholar
Potts, M. (1994). Desiccation tolerance in prokaryotes. Microbiology Reviews, 58, 755–805.Google ScholarPubMed
Potts, M. and Bowman, M. (1985). Sensitivity of Nostoc commune UTEX 584 (Cyanobacteria) to water stress. Archiv für Microbiologie, 141, 51–56.CrossRefGoogle Scholar
Pressel, S., Ligrone, R. and Duckett, J. G. (2006). The effects of de- and rehydration on food-conducting cells in the moss Polytrichum formosum Hedw.: a cytological study. Annals of Botany, 98, 67–76.CrossRefGoogle Scholar
Price, S. and Long, S. P. (1989). An in vivo analysis of the effect of SO2 fumigation on photosynthesis in Zea mays. Physiologica Plantarum, 76, 193–200.CrossRefGoogle Scholar
Prillinger, H. (1982). Zur genetischen Kontrolle und Evolution der sexuellen Fortplanzung und Heterothallie der Chitinpilzen. Zeitschrift für Mykologie, 48, 297–324.Google Scholar
Printzen, C. and Ekman, S. (2002). Genetic variability and its geographical distribution in the widely disjunct Cavernularia hultenii. Lichenologist, 34, 101–111.CrossRefGoogle Scholar
Printzen, C. and Ekman, S. (2003). Local population subdivision in the lichen Cladonia subcervicornis as revealed by mitochondrial cytochrome oxidase subunit 1 intron sequences. Mycologia, 95, 399–406.CrossRefGoogle ScholarPubMed
Printzen, C. and Kantvilas, G. (2004). Hertelidea, genus novum Stereocaulaearum (Ascomycetes lichenisati). Bibliotheca Lichenologica, 88, 539–553.Google Scholar
Printzen, C. and Lumbsch, H. T. (2000). Molecular evidence for the diversification of extant lichens in the late Cretaceous and Tertiary. Molecular Phylogenetics and Evolution, 17, 379–387.CrossRefGoogle ScholarPubMed
Printzen, C., Ekman, S. and Tønsberg, T. (2003). Phylogeography of Cavernularia hultenii: evidence of slow genetic drift in a widely disjunct lichen. Molecular Ecology, 12, 1473–1486.CrossRefGoogle Scholar
Proctor, M. C. F., Ligrone, R. and Duckett, J. G. (2006). Desiccation tolerance in the moss Polytrichum formosum Hedw.: physiological and fine-structural changes during desiccation and recovery. Annals of Botany, 99, 75–93.CrossRefGoogle Scholar
Puckett, K. J. (1976). The effect of heavy metals in some aspects of lichen physiology. Canadian Journal of Botany, 54, 2695–2703.CrossRefGoogle Scholar
Puckett, K. J. (1978). Element Levels in Lichens from the Northwest Territories. Report ARQA–56–76. Downsview: Atmospheric Environment Service, Environment Canada.Google Scholar
Puckett, K. J. (1988). Bryophytes and lichens as monitors of metal deposition. Bibliotheca Lichenologica, 30, 231–267.Google Scholar
Puckett, K. J. and Burton, M. A. S. (1981). The effect of trace elements on lower plants. In Effect of Heavy Metal Pollution on Plants, Vol. 2: Metals in the Environment, ed. Lepp, N. W., pp. 213–238. London: Applied Science Publishers.Google Scholar
Puckett, K. J. and Finegan, E. J. (1980). An analysis of the element content of lichens from the Northwest Territories, Canada. Canadian Journal of Botany, 58, 2073–2089.CrossRefGoogle Scholar
Puckett, K. J., Nieboer, E., Flora, W. P. and Richardson, D. H. S. (1973). Sulphur dioxide: its effect on photosynthetic 14C fixation in lichens and suggested mechanisms of phytotoxicity. New Phytologist, 72, 141–154.CrossRefGoogle Scholar
Puckett, K. J., Richardson, D. H. S., Flora, W. P. and Nieboer, E. (1974). Photosynthetic 14C fixation by the lichen Umbilicaria muhlenbergii (Ach.) Tuck. following short exposures to aqueous sulphur dioxide. New Phytologist, 73, 1183–1192.CrossRefGoogle Scholar
Puckett, K. J., Tomassini, F. D., Nieboer, E. and Richardson, D. H. S. (1977). Potassium efflux by lichen thalli following exposure to aqueous sulphur dioxide. New Phytologist, 79, 135–145.CrossRefGoogle Scholar
Punz, W. (1979). Der Einfluss isolierter und kombinierter Schadstoffe auf die Flechtenphotosynthese. Photosynthetica, 13, 428–433.Google Scholar
Purvis, A., Gittleman, J. L. and Brooks, T., eds. (2005). Phylogeny and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Purvis, O. W. (1984). The occurrence of copper oxalate in lichens growing on copper sulphide-bearing rocks in Scandinavia. Lichenologist, 16, 197–204.CrossRefGoogle Scholar
Purvis, O. W. (2000). Lichens. London: Natural History Museum and Washington: Smithsonian Institution.Google Scholar
Purvis, O. W., Coppins, B. J., Hawksworth, D. J., James, P. W. and Moore, M. D. (1992). The Lichen Flora of Great Britain and Ireland. London: Natural History Publications.Google Scholar
Purvis, O. W., Elix, J. A.Broomhead, J. A. and Jones, G. C. (1987). The occurrence of copper-norstictic acid in lichens from cupriferous substrata. Lichenologist, 19, 193–203.CrossRefGoogle Scholar
Qui, B. S. and Gao, K. S. (2001). Photosynthetic characteristics of the terrestrial blue-green alga Nostoc flagelliforme. European Journal of Phycology, 36, 147–156.Google Scholar
Quilhot, W., Fernández, E., Rubio, C., Goddard, M. and Hidalgo, M. E. (1998). Lichen secondary products and their importance in environmental studies. In Lichenology in Latin America: History, Current Knowledge and Applications, ed. Marcelli, M. P. and Seaward, M. R. D., pp. 171–179. São Paulo: CETESB.Google Scholar
Quispel, A. (1960). Respiration of lichens. In Pflanzenatmung einschliesslich Gärung und Säurestoffwechsel (Handbuch der Pflanzenphysiologie vol XII/2), ed. Wolf, J., pp. 455–460. Berlin: Springer.Google Scholar
Rai, A. N. (1988). Nitrogen metabolism. In Handbook of Lichenology, Vol. I, ed. Galun, M., pp. 201–237. Boca Raton: CRC Press.Google Scholar
Rai, A. N. (2002). Cyanolichens: nitrogen metabolism. In Cyanobacteria in Symbiosis, ed. Rai, A. N., Bergman, B. and Rasmussen, U., pp. 97–115. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Rai, A. N., Rowell, P. and Stewart, W. D. P. (1980). NH4+ assimilation and nitrogenase regulation in the lichen Peltigera aphthosa Willd. New Phytologist, 85, 545–555.CrossRefGoogle Scholar
Rai, A. N., Rowell, P. and Stewart, W. D. P. (1981). Nitrogenase activity and dark CO2 fixation in the lichen Peltigera aphthosa Willd. Planta, 151, 256–264.CrossRefGoogle ScholarPubMed
Rai, A. N., Rowell, P. and Stewart, D. P. (1983). Mycobiont-cyanobiont interactions during dark nitrogen fixation by the lichen Peltigera aphthosa. Physiologia Plantarum, 57, 285–290.CrossRefGoogle Scholar
Rambold, G. and Triebel, D. (1992). The inter-lecanoralean associations. Bibliotheca Lichenologica, 48, 3–201.Google Scholar
Rambold, G., Friedl, T. and Beck, A. (1998). Photobionts in lichens: possible indicators of phylogenetic relationships?Bryologist, 101, 392–397.CrossRefGoogle Scholar
Ramstad, S. and Hestmark, G. (2001). Population structure and size-dependent reproductive effort in Umbilicaria sporochroa. Mycologia, 93, 453–458.CrossRefGoogle Scholar
Rancan, F., Rosan, S., Boehm, K., et al. (2002). Protection against UVB irradiation by natural filters extracted from lichens. Journal of Photochemistry and Photobiology B: Biology, 68, 133–139.CrossRefGoogle ScholarPubMed
Randlane, T. and Saag, A. (1998). Synopsis of the genus Nephromopsis (Fam. Parmeliaceae, lichenized Ascomycota). Cryptogamie, Bryologie et Lichénologie, 19, 175–191.Google Scholar
Randlane, T., Saag, A. and Thell, A. (1997). A second updated world list of cetrarioid lichens. Bryologist, 100, 109–122.CrossRefGoogle Scholar
Rao, D. N. and LeBlanc, F. (1966). Effects of sulfur dioxide on the lichen alga with special reference to chlorophyll. Bryologist, 69, 69–75.CrossRefGoogle Scholar
Raven, J. A. (1992). Energy and nutrient acquisition by autotrophic symbioses and their asymbiotic ancestors. (Review). Symbiosis, 14, 33–60.Google Scholar
Raven, J. A., Johnston, A. M., Handley, L. L. and McInroy, S. G. (1990). Transport and assimilation of inorganic carbon by Lichina pygmea under emersed and submersed conditions. New Phytologist, 114, 407–417.CrossRefGoogle Scholar
Reeb, V., Lutzoni, F. and Roux, C. (2004). Contribution of RPB2 to multilocus phylogenetic studies of the euascomycetes (Pezizomycotina, Fungi) with special emphasis on the lichen-forming Acarosporaceae and evolution of polyspory. Molecular Phylogenetics and Evolution, 32, 1036–1060.CrossRefGoogle ScholarPubMed
Reich, P. B., Ellsworth, D. S., Walters, M. B., et al. (1999). Generality of leaf trait relationships: a test across six biomes. Ecology, 80, 1955–1969.CrossRefGoogle Scholar
Reich, P. B., Walters, M. B., Ellsworth, D. S., et al. (1998). Relationships of leaf dark respiration to leaf nitrogen, specific leaf area and leaf life-span: a test across biomes and functional groups. Oecologia, 114, 471–482.CrossRefGoogle ScholarPubMed
Reiners, W. A. and Olson, R. K. (1984). Effects of canopy components on throughfall chemistry: an experimental analysis. Oecologia, 63, 320–330.CrossRefGoogle ScholarPubMed
Reinke, J. (1894–1896). Abhandlungen über Flechten. Jahrbücher wissenschaftliche Botanik, 26, 28, 29.Google Scholar
Reisser, W. (1992). Endosymbiotic associations of algae with freshwater protozoa and invertebrates. In Algae and Symbioses: Plants, Animals, Fungi, Viruses, Interactions Explored, ed. Reisser, W., pp. 1–19. Bristol: Biopress.Google Scholar
Reiter, R. and Türk, R. (2000 a). Investigations on the CO2 exchange of lichens in the alpine belt. I. Comparative patterns of net CO2 exchange in Cladonia mitis, Thamnolia vermicularis and Umbilicaria cylindrica. Bibliotheca Lichenologica, 75, 333–351.Google Scholar
Reiter, R. and Türk, R. (2000 b). Investigations on the CO2 exchange of lichens in the alpine belt. II. Comparative patterns of net CO2 exchange in Cetraria islandica and Flavocetraria nivalis. Phyton (Austria), 40, 161–177.Google Scholar
Renhorn, K. E., Esseen, P.-A., Palmqvist, K. and Sundberg, B. (1997). Growth and vitality of epiphytic lichens. I. Responses to microclimate along a forest edge-interior gradient. Oecologia, 109, 1–9.CrossRefGoogle Scholar
Rennenberg, H. (1984). The fate of excess sulfur in higher plants. Annual Review of Plant Physiology, 35, 121–153.CrossRefGoogle Scholar
Rennenberg, H. and Polle, A. (1994). Metabolic consequences of atmospheric sulphur influx into plants. In Plant Responses to the Gaseous Environment, ed. Alscher, R. G. and Wellburn, A. R., pp. 165–180. London: Chapman and Hall.CrossRefGoogle Scholar
Reutimann, P. and Scheidegger, C. (1987). Importance of lichen secondary products in food choice of two oribatic mites (Acari) in an alpine meadow ecosystem. Journal of Chemical Ecology, 13, 363–369.CrossRefGoogle Scholar
Rhoades, F. M. (1977). Growth rates of Lobaria oregana as determined from sequential photographs. Canadian Journal of Botany, 55, 2226–2233.CrossRefGoogle Scholar
Rhoades, F. M. (1981). Biomass of epiphytic lichens and bryophytes on Abies lasiocarpa on a Mt. Baker lava flow. Bryologist, 84, 39–47.CrossRefGoogle Scholar
Rhoades, F. M. (1983). Distribution of thalli in a population of the epiphytic lichen Lobaria oregana and a model of population dynamics and production. Bryologist, 86, 309–331.CrossRefGoogle Scholar
Richardson, D. H. S. (1974). The Vanishing Lichens. New York: Hafner Press.Google Scholar
Richardson, D. H. S. (1988). Medicinal and other economic aspects of lichens. In Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 93–108. Boca Raton: CRC Press.Google Scholar
Richardson, D. H. S. (1991). Lichens and man. In Frontiers in Mycology, ed. Hawksworth, D. L., pp. 187–210. Kew: CAB International.Google Scholar
Richardson, D. H. S. (1999). War in the world of lichens: parasitism and symbiosis as exemplified by lichens and lichenicolous fungi. Mycological Research, 6, 641–650.CrossRefGoogle Scholar
Richardson, D. H. S. and Cameron, R. P. (2004). Cyanolichens: their response to pollution and possible management strategies for their conservation in northeastern North America. Northeastern Naturalist, 11, 1–22.CrossRefGoogle Scholar
Richardson, D. H. S. and Nieboer, E. (1983). Ecophysiological responses of lichens to sulphur dioxide. Journal of the Hattori Botanical Laboratory, 54, 331–351.Google Scholar
Richardson, D. H. S. and Smith, D. C. (1966). The physiology of the symbiosis in Xanthoria aureola (Ach.) Erichs. Lichenologist, 3, 202–206.CrossRefGoogle Scholar
Richardson, D. H. S. and Smith, D. C. (1968). Lichen physiology. IX. Carbohydrate movement from the Trebouxia symbiont of Xanthoria aureola. New Phytologist, 67, 61–68.CrossRefGoogle Scholar
Richardson, D. H. S. and Young, C. M. (1977). Lichens and vertebrates. In Lichen Ecology, ed. Seaward, M. R. D., pp. 121–144. London: Academic Press.Google Scholar
Richardson, D. H. S., Beckett, P. J. and Nieboer, E. (1980). Nickel in lichens, bryophytes, fungi and algae. In Nickel in the Environment, ed. Nriagu, J. O., pp. 367–406. New York: John Wiley.Google Scholar
Richardson, D. H. S., Hill, D. J. and Smith, D. C. (1968). Lichen physiology. XI. The role of the alga in determining the pattern of carbohydrate movement between lichen symbionts. New Phytologist, 67, 469–486.CrossRefGoogle Scholar
Richardson, D. H. S., Nieboer, E., Lavoie, P. and Padovan, D. (1979). The role of metal-ion binding in modifying the toxic affects of sulphur dioxide on the lichen Umbilicaria muhlenbergii. II. 14C-fixation studies. New Phytologist, 82, 633–643.CrossRefGoogle Scholar
Richardson, D. H. S., Nieboer, E., Lavoie, P. and Padovan, D. (1984). Anion accumulation by lichens. I. The characteristics and kinetics of arsenate uptake by Umbilicaria muhlenbergii. New Phytologist, 96, 71–82.CrossRefGoogle Scholar
Riddle, B. R. (2005). Is biogeography emerging from its identity crisis?Journal of Biogeography, 32, 185–186.CrossRefGoogle Scholar
Riddle, B. R. and Hafner, D. J. (2004). The past and future roles of phylogeography. In Frontiers of Biogeography: New Directions in the Geography of Nature, ed. Lomolino, M. V. and Heaney, L. R., pp. 93–110. Sunderland: Sinauer Associates.Google Scholar
Rikkinen, J. (2003). Calicioid lichens from European tertiary amber. Mycologia, 95, 1032–1036.CrossRefGoogle ScholarPubMed
Rikkinen, J., Oksanen, I. and Lohtander, K. (2002). Lichen guilds share related cyanobacterial symbionts. Science, 297, 357.CrossRefGoogle ScholarPubMed
Rindi, F. and Guiry, M. D. (2003). Composition and distribution of subaerial algal assemblages in Galway City, western Ireland. Cryptogamie Algologie, 24, 245–267.Google Scholar
Roberts, B. A. and Thompson, L. K. (1980). Lichens as indicators of fluoride emission from a phosphorous plant, Long Harbour, Newfoundland, Canada. Canadian Journal of Botany, 58, 2218–2228.CrossRefGoogle Scholar
Robinson, C. H. (2001). Cold adaptation in Arctic and Antarctic fungi. New Phytologist, 151, 341–353.CrossRefGoogle Scholar
Rogers, R. W. (1990). Ecological strategies of lichens. Lichenologist, 22, 149–162.CrossRefGoogle Scholar
Rogers, R. W. (1992). Lichen ecology and biogeography. Flora of Australia 54, 30–42.Google Scholar
Rollin, E. M., Milton, E. J. and Roche, P. (1994). The influence of weathering and lichen cover on the reflectance spectra of granitic rocks. Remote Sensing of Environment, 50, 194–199.CrossRefGoogle Scholar
Rolstad, J., Gjerde, I., Storaunet, K. O. and Rolstad, E. (2001). Epiphytic lichens in Norwegian coastal spruce forest: historical logging and present forest structure. Ecological Applications, 11, 421–436.CrossRefGoogle Scholar
Romagni, J. G., Thomas, M. A., Gries, C. and Nash, T. H. III (1997). Sulfite reductase activity in six lichen species as a response to fumigations with sulfur dioxide. Supplement to Plant Physiology, 114, 58.Google Scholar
Romagni, J. G., Thomas, M. A. and Nash, T. H. III (1998). Detoxification of SO2 in lichens: total glutathione. Supplement to Plant Physiology, 115, 89.Google Scholar
Romeike, J., Friedl, T., Helms, G. and Ott, S. (2002). Genetic diversity of algal and fungal partners in four species of Umbilicaria (lichenized ascomycetes) along a transect of the Antarctic peninsula. Molecular Biology and Evolution, 19, 1209–1217.CrossRefGoogle ScholarPubMed
Rorat, T. (2006). Plant dehydrins: tissue location, structure and function. Cell Molecular Biology Letters, 11, 536–556.CrossRefGoogle ScholarPubMed
Rose, C. I. and Hawksworth, D. L. (1981). Lichen recolonization in London's cleaner air. Nature, 289, 289–292.CrossRefGoogle Scholar
Rose, F. (1976). Lichenological indicators of age and environmental quality in woodlands. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 279–307. London: Academic Press.Google Scholar
Rosentreter, R. (1993). Vagrant lichens in North America. Bryologist, 96, 333–338.CrossRefGoogle Scholar
Rosentreter, R. and Eldridge, D. J. (2002). Monitoring biodiversity and ecosystem function: grasslands, deserts, and steppe. In Monitoring with Lichens – Monitoring Lichens. Nato Science Series IV: Earth and Environmental Sciences, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 223–237. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Roser, D. J., Mellick, D. R., Ling, H. U. and Seppelt, R. D. (1992). Polyol and sugar content of terrestrial plants from continental Antarctica. Antarctic Science, 4, 413–420.CrossRefGoogle Scholar
Ross, L. J. (1982). Lichens on coastal live oak in relation to ozone. M. S. Thesis. Arizona State University, Tempe, Arizona.
Ross, L. J. and Nash, T. H. III (1983). Effect of ozone on gross photosynthesis of lichens. Environmental and Experimental Botany, 23, 71–77.CrossRefGoogle Scholar
Rosso, A. L., McCune, B. and Rambo, T. R. (2000). Ecology and conservation of a rare, old-growth-associated canopy lichen in a silvicultural landscape. Bryologist, 103, 117–127.CrossRefGoogle Scholar
Roux, C. (1981). Étude écologique et phytosociologique des peuplements lichéniques saxicoles–calcicoles du sud-est de la France. Bibliotheca Lichenologica, 15, 1–557.Google Scholar
Rubio, C., Fernández, E., Hidalgo, M. E. and Quilhot, W. (2002). Effects of solar UV-B radiation in the accumulation of rhizocarpic acid in a lichen species from alpine zones of Chile. Boletín, Sociedad Chilena de Química, 47, 67–72.CrossRefGoogle Scholar
Ruchty, A., Rosso, A. L. and McCune, B. (2001). Changes in epiphyte communities as the shrub, Acer circinatum, develops and ages. Bryologist, 104, 272–281.CrossRefGoogle Scholar
Rundel, P. W. (1969). Clinal variation in the production of usnic acid in Cladonia subtenuis along light gradients. Bryologist, 72, 40–44.CrossRefGoogle Scholar
Rundel, P. W. (1978). The ecological role of secondary lichen substances. Biochemical Systematics and Ecology, 6, 157–170.CrossRefGoogle Scholar
Rundel, P. W. (1982). Water uptake by organs other than roots. In Physiological Plant Ecology. Vol. II: Water Relations and Carbon Assimilation, ed. Lange, O. L., Nobel, P. S., Osmond, C. B. and Ziegler, H., pp. 111–134. Encyclopedia of Plant Physiology 12B. Berlin: Springer.CrossRefGoogle Scholar
Rundel, P. W. (1988). Water relations. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 17–36. Boca Raton: CRC Press.Google Scholar
Ruoss, E. (1987). Species differentiation in a group of reindeer lichens (Cladonia subg. Cladina). Bibliotheca Lichenologica, 25, 197–206.Google Scholar
Rychert, R. C. and Skujins, J. (1974). Nitrogen fixation by blue-green algae–lichen crusts in the Great Basin Desert. Proceedings of the Soil Science Society of America, 38, 768–771.CrossRefGoogle Scholar
Salisbury, F. B. and Ross, C. W. (1992). Plant Physiology. 4th edn. Belmont: Wadsworth Publishing.Google Scholar
Sancho, L. G. and Kappen, L. (1989). Photosynthesis and water relations and the role of anatomy in Umbilicaricaeae (Lichenes) from central Spain. Oecologia, 81, 473–480.CrossRefGoogle ScholarPubMed
Sancho, L. G. and Pintado, A. (2004). Evidence of high annual growth rate for lichens in the maritime Antarctic. Polar Biology, 27, 312–319.CrossRefGoogle Scholar
Sancho, L. G., Torre, R., Horneck, G., et al. (2007). Lichens survive in space: Results from the 2005 LICHENS experiment. Astrobiology, 7, 443–454.CrossRefGoogle ScholarPubMed
Sancho, L. G., Pintado, A., Blanquer, J. M., Raggio, J. and Vilches, R. (2004). Lichen morphology, thallus water content and photosynthetic performance. Looking for a single trait. In Book of Abstracts of the 5th IAL Symposium: Lichens in Focus, ed. Randlane, T. and Saag, A., p. 47. Tartu: Tartu University Press.Google Scholar
Sancho, L. G., Pintado, A., Green, T. G. A., Pannewitz, S. and Schroeter, B. (2003). Photosynthetic and morphological variation within and among populations of the Antarctic lichen Umbilicaria aprina: implications of the thallus size. Bibliotheca Lichenologica, 86, 299–311.Google Scholar
Sancho, L. G., Pintado, A., Valladares, F., Schroeter, B. and Schlensog, M. (1997). Photosynthetic performance of cosmopolitan lichens in the maritime Antarctic. Bibliotheca Lichenologica, 67, 197–210.Google Scholar
Sancho, L. G., Schroeter, B. and Del-Prado, R. (2000 a). Ecophysiology and morphology of the globular erratic lichen Aspicilia fruticulosa (Eversm.) Flag. from central Spain. Bibliotheca Lichenologica, 75, 137–147.Google Scholar
Sancho, L. G., Valladares, F., Schroeter, B. and Kappen, L. (2000 b). Ecophysiology of Antarctic versus temperate populations of a bipolar lichen: the key role of the photosynthetic partner. In Antarctic Ecosystems: Models for Wider Ecological Understanding, ed. Davison, W., Williams, C. H. and Broady, P., pp. 190–194. Christchurch: New Zealand Natural Sciences Publications.Google Scholar
Sanders, W. B. (1989). Growth and development of the reticulate thallus in the lichen Ramalina menziesii. American Journal of Botany, 76, 666–678.CrossRefGoogle Scholar
Sanders, W. B. (2001 a). Preliminary light microscope observations of fungal and algal colonization and lichen thallus initiation on glass slides placed near foliicolous lichen communities within a lowland tropical forest. Symbiosis, 31, 85–94.Google Scholar
Sanders, W. B. (2001 b). Lichens: the interface between mycology and plant morphology. BioScience, 51, 1025–1035.CrossRefGoogle Scholar
Sanders, W. B. (2005). Observing microscopic phases of lichen life cycles on transparent substrata placed in situ. Lichenologist, 37, 373–382.CrossRefGoogle Scholar
Sanders, W. B. and Lücking, R. (2002). Reproductive strategies, relichenization and thallus development observed in situ in leaf-dwelling lichen communities. New Phytologist, 155, 425–435.CrossRefGoogle Scholar
Sanders, W. B., Moe, R. L. and Ascaso, C. (2004). The intertidal marine lichen formed by the pyrenomycete fungus Verrucaria tavaresiae (Ascomycotina) and the brown alga Petroderma maculiforme (Phaeophyceae): thallus organization and symbiont interaction. American Journal of Botany, 91, 511–522.CrossRefGoogle ScholarPubMed
Sanmartín, I. and Ronquist, F. (2004). Southern Hemisphere biogeography inferred by event-based models: plant versus animal patterns. Systematic Biology, 53, 216–243.CrossRefGoogle ScholarPubMed
Santesson, J. (1969). Chemical studies on lichens. 10. Mass spectrometry on lichens. Arkiv för Chemie, 30, 363–377.Google Scholar
Santesson, R. (1939). Amphibious pyrenolichens I. Arkiv för Botanik, 29A, 1–67.Google Scholar
Santesson, R. (1952). Foliicolous lichens. I. A revision of the taxonomy of the obligately foliicolous, lichenized fungi. Symbolae Botanicae Upsalienses, 12, 1–590.Google Scholar
Santesson, R., Moberg, R., Nordin, A., Tønsberg, T. and Vitikainen, O. (2004). Lichen-forming and Lichenicolous Fungi of Fennoscandia. Uppsala: Museum of Evolution, Uppsala University.Google Scholar
Sanz, M.-J., Gries, C. and Nash, T. H. III (1992). Dose-response relationships for SO2 fumigations in the lichens Evernia prunastri (L.) Ach. and Ramalina fraxinea (L.) Ach. New Phytologist, 122, 313–319.CrossRefGoogle Scholar
Schaper, T. and Ott, S. (2003). Photobiont selectivity and interspecific interactions in lichen communities. I. Culture experiments with the mycobiont Fulgensia bracteata. Plant Biology, 5, 441–450.CrossRefGoogle Scholar
Scheidegger, C. (1985). Systematische Studien zur Krustenflechte Anzina carneonivea (Trapeliacae, Lecanorales). Nova Hedwigia, 41, 191–218.Google Scholar
Scheidegger, C. (1993). A revision of saxicolous species of the genus Buellia De Not. and formerly included genera in Europe. Lichenologist, 25, 315–364.CrossRefGoogle Scholar
Scheidegger, C. (1994 a). Low-temperature scanning electron microscopy: the localization of free and perturbed water and its role in the morphology of the lichen symbionts. Cryptogamic Botany, 4, 290–299.Google Scholar
Scheidegger, C. (1994 b). Reproductive strategies in Vezdaea (Lecanorales, lichenized Ascomycetes): a low-temperature scanning electron microscopy study of a ruderal species. Cryptogamic Botany, 5, 163–171.Google Scholar
Scheidegger, C. (1998). Erioderma pedicellatum: a critically endangered lichen species. Species, 30, 68–69.Google Scholar
Scheidegger, C. and Schroeter, B. (1995). Effects of ozone fumigation on epiphytic macrolichens: ultrastructure, CO2 gas exchange and chlorophyll fluorescence. Environmental Pollution, 88, 345–354.CrossRefGoogle ScholarPubMed
Scheidegger, C., Schroeter, B. and Frey, B. (1995 a). Structural and functional processes during water vapour uptake and desiccation in selected lichens with green algal photobionts. Planta, 197, 399–409.CrossRefGoogle Scholar
Scheidegger, C., Wolseley, P. A. and Thor, G., eds., (1995 b). Conservation biology of lichenised fungi. Mitteilungen der Eidgenössischen Forschungsanstalt für Wald, Schnee und Landschaft, 70, 1–173.Google Scholar
Scherrer, S. and Honegger, R. (2003). Inter- and intraspecific variation of homologous hydrophobin (H1) gene sequences among Xanthoria spp. (lichen-forming ascomycetes). New Phytologist, 158, 375–389.CrossRefGoogle Scholar
Scherrer, S., Vries, O. M. H., Dudler, R., Wessels, J. G. H. and Honegger, R. (2000). Interfacial self-assembly of fungal hydrophobins of the lichen-forming ascomycetes Xanthoria parietina and X. ectaneoides. Fungal Genetics and Biology, 30, 81–93.CrossRefGoogle ScholarPubMed
Scherrer, S., Haisch, A. and Honegger, R. (2002). Characterization and expression of XPH1, the hydrophobin gene of the lichen-forming ascomycete Xanthoria parietina. New Phytologist, 154, 175–184.CrossRefGoogle Scholar
Scherrer, S., Zippler, U. and Honegger, R. (2005). Characterisation of the mating-type locus in the genus Xanthoria (lichen-forming ascomycetes, Lecanoromycetes). Fungal Genetics and Biology, 42, 976–988.CrossRefGoogle Scholar
Schlee, D., Kandzia, R., Tintemann, H. and Türk, R. (1995). Activity of superoxide dismutase and malondialdehyde content in lichens along an altitude profile. Phyton, 35, 233–242.Google Scholar
Schlensog, M., Pannewitz, S., Green, T. G. A. and Schroeter, B. (2004). Metabolic recovery of continental antarctic cryptogams after winter. Polar Biology, 27, 399–408.CrossRefGoogle Scholar
Schlensog, M., Schroeter, B. and Green, T. G. A. (2000). Water dependent photosynthetic activity of lichens from New Zealand: differences in the green algal and the cyanobacterial thallus parts of photosymbiodemes. Bibliotheca Lichenologica, 75, 149–160.Google Scholar
Schlensog, M., Schroeter, B., Sancho, L. G., Pintado, A. and Kappen, L. (1997). Effect of strong irradiance on photosynthetic performance of the melt-water dependent cyanobacterial lichen Leptogium puberulum (Collemataceae) Hue from the maritime Antarctic. Bibliotheca Lichenologica, 67, 235–246.Google Scholar
Schmitt, I. and Lumbsch, H. T. (2004). Molecular phylogeny of the Pertusariaceae supports secondary chemistry as an important systematic character set in lichen-forming ascomycetes. Molecular Phylogenetics and Evolution, 33, 43–55.CrossRefGoogle ScholarPubMed
Schmitt, I., Lumbsch, H. T. and Søchting, U. (2003). Phylogeny of the lichen genus Placopsis and its allies based on Bayesian analyses of nuclear and mitochondrial sequences. Mycologia, 95, 827–835.CrossRefGoogle ScholarPubMed
Schmitt, I., Martin, M. P., Kautz, S. and Lumbsch, T. H. (2005 a). Diversity of non-reducing polyketide synthase genes in the Pertusariales (lichenized Ascomycota). A phylogenetic perspective. Phytochemistry, 66, 1241–1253.CrossRefGoogle ScholarPubMed
Schmitt, I., Messuti, M. I., Feige, G. B. and Lumbsch, H. T. (2001). Molecular data support rejection of the generic concept in the Coccotremataceae (Ascomycota). Lichenologist, 33, 315–321.CrossRefGoogle Scholar
Schmitt, I., Mueller, G. M. and Lumbsch, H. T. (2005 b). Ascoma morphology is homoplaseous and phylogenetically misleading in some pyrenocarpous lichens. Mycologia, 97, 362–374.CrossRefGoogle ScholarPubMed
Schmitt, I., Yamamoto, Y. and Lumbsch, H T. (2006). Phylogeny of Pertusariales (Ascomycotina): resurrection of Ochrolechiaceae and a new circumscription of Megasporaceae. Journal of the Hattori Botanical Laboratory, 100, 753–764.Google Scholar
Schmull, M. and Hauck, M. (2003). Element microdistribution in the bark of Abies balsamea and Picea rubens and its impact on epiphytic lichen abundance on Whiteface Mountain, New York. Flora, 198, 293–303.CrossRefGoogle Scholar
Schreiber, U., Bilger, W. and Neubauer C. (1994). Chlorophyll fluorescence as a nonintrusive indicator for rapid assessment of in vivo photosynthesis. In Ecophysiology of Photosynthesis ed. Schulze, E.-D. and Caldwell, M. M., pp. 49–70. Ecological Studies 100. Berlin: Springer.CrossRefGoogle Scholar
Schroeter, B. and Scheidegger, C. (1995). Water relations in lichens at subzero temperatures: structural changes and carbon dioxide exchange in the lichen Umbilicaria aprina from continental Antarctica. New Phytologist, 131, 273–285.CrossRefGoogle Scholar
Schroeter, B., Green, T. G. A., Kappen, L. and Seppelt, D. (1994). Carbon dioxide exchange at subzero temperatures. Field measurements on Umbilicaria aprina in Antarctica. Cryptogamic Botany, 4, 233–241.Google Scholar
Schroeter, B., Kappen, L., Schulz, F. and Sancho, L. G. (2000). Seasonal variation in the carbon balance of lichens in the maritime Antarctic: long-term measurements of photosynthetic activity in Usnea aurantiaco-atra. In Antarctic Ecosystems: Model for Wider Ecological Understanding, ed. Davison, W., Howard-Williams, C. and Broady, P., pp. 220–224. Christchurch: Caxton Press.Google Scholar
Schroeter, B., Schulz, F. and Kappen, L. (1997). Hydration-related spatial and temporal variation of photosynthetic activity in Antarctic lichens. In Antarctic Communities: Species, Structure and Survival, ed. Battaglia, B., Valencia, J. and Walton, D. W. H., pp. 221–225. Cambridge: Cambridge University Press.Google Scholar
Schulz, M. (1995). Protein and ubiquitin conjugate patterns of Peltigera horizontalis (Huds.) Baumg. during desiccation and rehydration. In Flechten Follmann. Contributions to Lichenology in Honour of Gerhard Follmann, ed. Daniels, F. J. A., Schultz, M. and Peine, J., pp. 87–96. Cologne: Botanical Institute, University of Cologne.Google Scholar
Schulze, E.-D., Beck, E. and Müller-Hohenstein, K. (2002). Pflanzenökologie. Heidelberg: Spektrum Akademischer.Google Scholar
Schulze, E.-D. and Chapin, F. S., III (1987). Plant specialization to environments of different resource availability. In Potentials and Limitations of Ecosystem Analysis, ed. Schulze, E.-D. and Zwolfer, H., pp. 120–148. Berlin: Springer.CrossRefGoogle Scholar
Schüssler, A. (2002). Molecular phylogeny, taxonomy, and evolution of Geosiphon pyriformis and arbuscular mycorrhizal fungi. Plant and Soil, 244, 75–83.CrossRefGoogle Scholar
Schüssler, A., Schnepf, E., Mollenhauer, D. and Kluge, M. (1995). The fungal bladders of the endocyanosis Geosiphon pyriforme, a Glomus-related fungus: cell wall permeability indicates a limiting pore radius of only 0.5 nm. Protoplasma, 185, 131–139.CrossRefGoogle Scholar
Schüssler, A., Schwarzott, D. and Walker, C. (2001). A new fungal phylum, the Glomeromycota: phylogeny and evolution. Mycological Research, 105, 1413–1421.CrossRefGoogle Scholar
Schuster, G., Ott, S. and Jahns, H. M. (1985). Artificial cultures of lichens in the natural environment. Lichenologist, 17, 247–253.CrossRefGoogle Scholar
Schwartzman, D. W. and Volk, T. (1989). Biotic enhancement of weathering and the habitability of Earth. Nature, 340, 457–460.CrossRefGoogle Scholar
Schwendener, S. (1867). Über die wahre Natur der Flechtengonidien. Verhandlungen der schweizerischen naturforschenden Gesellschaft, 57, 9–11.Google Scholar
Schwendener, S. (1869). Die Algentypen der Flechtengonidien. Basel: Schultze.Google Scholar
Scotter, G. W. (1965). Chemical composition of forage lichens from northern Saskatchewan as related to use by barren-ground caribou in the taiga of northern Canada. Canadian Journal of Plant Sciences, 45, 246–250.CrossRefGoogle Scholar
Seaward, M. R. D. (1973). Lichen ecology of the Scunthorpe heathlands. I. Mineral accumulation. Lichenologist, 5, 423–433.CrossRefGoogle Scholar
Seaward, M. R. D. (1976). Performance of Lecanora muralis in an urban environment. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 323–357. London: Academic Press.Google Scholar
Seaward, M. R. D. (1982 a). Lichen ecology of changing urban environments. In Urban Ecology, ed. Bornkamm, R., Lee, J. A. and Seaward, M. R. D., pp. 181–189. Oxford: Blackwell Scientific.Google Scholar
Seaward, M. R. D. (1982 b). Principles and priorities of lichen conservation. Journal of the Hattori Botanical Laboratory, 52, 401–406.Google Scholar
Seaward, M. R. D. (1988). Contribution of lichens to ecosystems. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 107–129. Boca Raton: CRC Press.Google Scholar
Seaward, M. R. D. (1993). Lichens and sulphur dioxide air pollution: field studies. Environmental Reviews, 1, 73–91.CrossRefGoogle Scholar
Seaward, M. R. D. (1996 a). Checklist of Tunisian lichens. Bocconea, 6, 115–148.Google Scholar
Seaward, M. R. D. (1996 b). Lichens and the environment. In A Century of Mycology, ed. Sutton, B. C., pp. 293–320. Cambridge: Cambridge University Press.Google Scholar
Seaward, M. R. D. (1997). Urban deserts bloom: a lichen renaissance. Bibliotheca Lichenologica, 67, 297–309.Google Scholar
Seaward, M. R. D. (ed.) (1998). Lichen Atlas of the British Isles. London: British Lichen Society.Google Scholar
Seaward, M. R. D. (2004). The use of lichens for environmental impact assessment. Symbiosis, 37, 293–305.Google Scholar
Seaward, M. R. D. and Aptroot, A. (2003). Lichens of Silhouette Island (Seychelles). Bibliotheca Lichenologica, 86, 423–439.Google Scholar
Seaward, M. R. D. and Coppins, B. J. (2004). Lichens and hypertrophication. Bibliotheca Lichenologica, 88, 561–572.Google Scholar
Seaward, M. R. D. and Edwards, H. G. M. (1997). Biological origin of major chemical disturbances on ecclesiastical architecture studied by Fourier Transform Raman spectroscopy. Journal of Raman Spectroscopy, 28, 691–696.3.0.CO;2-4>CrossRefGoogle Scholar
Seaward, M. R. D. and Letrouit-Galinou, M. (1991). Lichens return to the Jardin du Luxembourg after an absence of almost a century. Lichenologist, 23, 181–186.CrossRefGoogle Scholar
Sedelnikova, N. V. and Cheremisin, D. V. (2001). The use of lichens for dating of petroglyphs. Siberian Journal of Ecology, 8, 479–481.Google Scholar
Sensen, M. and Richardson, D. H. S. (2002). Mercury levels in lichens from different host trees around a chlor-alkali plant in New Brunswick, Canada. Science of the Total Environment, 293, 31–45.CrossRefGoogle ScholarPubMed
Sérusiaux, E. (1985). Goniocysts, goniocystangia and Opegrapha lambinonii and related species. Lichenologist, 17, 1–25.CrossRefGoogle Scholar
Sérusiaux, E. (1986). The nature and origin of campylidia in lichenized fungi. Lichenologist, 18, 1–35.CrossRefGoogle Scholar
Sérusiaux, E. (1989). Liste Rouge des Macrolichens dans la Communauté Européenne. Liège: Centre des Recherches sur les Lichens.Google Scholar
Seymour, F. A., Crittenden, P. D., Dickinson, M. J., et al. (2005 a). Breeding systems in the lichen-forming fungal genus Cladonia. Fungal Genetics and Biology, 42, 554–563.CrossRefGoogle ScholarPubMed
Seymour, F. A., Crittenden, P. D. and Dyer, P. S. (2005 b). Sex in the extremes: lichen-forming fungi. Mycologist, 19, 51–58.CrossRefGoogle Scholar
Sharma, P., Bergman, B., Hallbom, L. and Hofsten, A. (1982). Ultrastructural changes of Nostoc of Peltigera canina in presence of SO2. New Phytologist, 92, 573–579.CrossRefGoogle Scholar
Sheridan, R. P. (1979). Impact of emissions from coal-fired electricity generating facilities on N2-fixing lichens. Bryologist, 82, 54–58.CrossRefGoogle Scholar
Shibata, S. (1973). Some aspects of lichen chemotaxonomy. In Chemistry in Botanical Classification, Vol. 25, ed. Bendz, G. and Santesson, J., pp. 241–249. Stockholm: Nobel Symposia: Medicine and Natural Sciences.Google Scholar
Shibata, S. (1992). Studies on some lichen metabolites and their development. Journal of Japanese Botany, 67, 63–71.Google Scholar
Shibuya, M., Ebizuka, Y., Noguchi, H., Iitaka, Y. and Sankawa, U. (1983). Inhibition of prostaglandin biosynthesis by 4-O-methylcryptochlorophaeic acid: synthesis of monomeric arylcarboxylic acids for inhibitory activity testing and X-ray analysis of 4-O-methylcryptochlorophaeic acid. Chemical Pharmaceutical Bulletin of Tokyo, 31, 407–413.CrossRefGoogle ScholarPubMed
Shuvalov, V. A. and Heber, U. (2003). Photochemical reactions in dehydrated photosynthetic organisms, leaves, chloroplasts, and photosystem II particles: reversible reduction of pheophytin and chlorophyll and oxidation of β-carotene. Chemical Physics, 294, 227–237.CrossRefGoogle Scholar
Sigal, L. L. and Johnston, J. W. Jr. (1986). Effects of acidic rain and ozone on nitrogen fixation and photosynthesis in the lichen Lobaria pulmonaria (L.) Hoffm. Environmental and Experimental Botany, 26, 59–64.CrossRefGoogle Scholar
Sigal, L. L. and Nash, T. H. III (1983). Lichen communities on conifers in southern California: an ecological survey relative to oxidant air pollution. Ecology, 64, 1343–1354.CrossRefGoogle Scholar
Sigal, L. L. and Taylor, O. C. (1979). Preliminary studies of the gross photosynthetic response of lichens to peroxyacetylnitrite fumigations. Bryologist, 82, 564–575.CrossRefGoogle Scholar
Silberstein, L., Siegel, B. Z., Siegel, S. M., Mukhtar, A. and Galun, M. (1996 a). Comparative studies on Xanthoria parietina, a pollution-resistant lichen, and Ramalina duriaei, a sensitive species. I. Effects of air pollution on physiological processes. Lichenologist, 28, 355–365.CrossRefGoogle Scholar
Silberstein, L., Siegel, B. Z., Siegel, S. M., Mukhtar, A. and Galun, M. (1996 b). Comparative studies on Xanthoria parietina, a pollution-resistant lichen, and Ramalina duriaei, a sensitive species. II. Evaluation of possible air pollution-protection mechanisms. Lichenologist, 28, 367–383.CrossRefGoogle Scholar
Sillett, S. C. and Goslin, M. N. (1999). Distribution of epiphytic macrolichens in relation to remnant trees in a multiple-age Douglas-fir forest. Canadian Journal of Forest Research, 29, 1204–1215.CrossRefGoogle Scholar
Sillett, S. C., McCune, B., Peck, J. E., Rambo, T. R. and Ruchty, A. (2000). Dispersal limitations of epiphytic lichens result in species dependent on old-growth forests. Ecological Applications, 10, 789–799.CrossRefGoogle Scholar
Simpson, T. J. (1995). Polyketide biosynthesis. Chemistry and Industries, 1995, 407–411.Google Scholar
Sinnemann, S. J., Andrésson, Ó. S., Brown, D. W. and Miao, V. P. (2000). Cloning and heterologous expression of Solorina crocea pyrG. Current Genetics, 37, 333–338.CrossRefGoogle ScholarPubMed
Sipman, H. J. M. (1994). Foliicolous lichens on plastic tape. Lichenologist, 26, 311–312.CrossRefGoogle Scholar
Sipman, H. J. M. (1997). Observations on the foliicolous lichen and bryophyte flora in the canopy of a semi-deciduous tropical forest. Abstracta Botanica, 21, 153–161.Google Scholar
Sipman, H. J. M. (2002). The significance of the northern Andes for lichens. Botanical Review, 68, 88–99.CrossRefGoogle Scholar
Sipman, H. J. M. (2006 a). Diversity and biogeography of lichens in neotropical montane oak forests. In Ecology and Conservation of Neotropical Oak Forests, ed. Kappelle, M., pp. 69–81. Ecological Studies 185. Berlin: Springer.CrossRefGoogle Scholar
Sipman, H. J. M. (2006b). Identification key and literature guide to the genera of lichenized fungi (Lichens) in the Neotropics (provisional version). Online: www.bgbm.org/sipman/keys/neokeyA.htm.
Sipman, H. J. M. and Harris, R. C. (1989). Lichens. In Tropical Rain Forest Ecosystems, ed. Lieth, H. and Werger, M. J. A., pp. 303–309. Amsterdam: Elsevier.Google Scholar
Skujins, J. and Klubek, B. (1978). Nitrogen fixation and cycling by blue-green algae-lichen-crusts in arid rangeland soils. Ecological Bulletin (Stockholm), 26, 164–171.Google Scholar
Skulachev, V. P. (1998). Uncoupling: new approaches to an old problem of bioenergetics. Biochimica Biophysica Acta, 1363, 100–124.CrossRefGoogle Scholar
Skult, H. (1984). The Parmelia omphaloides (Ascomycetes) complex in Eastern Fennoscandia. Annales Botanici Fennici, 21, 117–142.Google Scholar
Skye, E. (1968). Lichens and air pollution: a study of cryptogamic epiphytes and environment in the Stockholm region. Acta Phytogeographica Suecica, 52, 8–123.Google Scholar
Slocum, R. D., Ahmadijan, V. and Hildreth, K. C. (1980). Zoosporogenesis in Trebouxia gelatinosa: ultrastrucutral potential for zoospore release and implications for the lichen association. Lichenologist, 12, 173–187.CrossRefGoogle Scholar
Sluiman, H. J. (1989). The green algal class Ulvophyceae – an ultrastructural survey and classification. Cryptogamic Botany, 1, 83–94.Google Scholar
Sluiman, H. J., Kouwets, F. A. C. and Blommers, P. C. J. (1989). Classification and definition of cytokinetic patterns in Green Algae: sporulation versus (vegetative) cell division. Archiv für Protistenkunde, 137, 277–90.CrossRefGoogle Scholar
Smith, D. C. and Douglas, A. (1987). The Biology of Symbiosis. London: Edward Arnold.Google Scholar
Smith, D. C. and Molesworth, S. (1973). Lichen physiology. XIII. Effects of rewetting of dry lichens. New Phytologist, 72, 525–533.CrossRefGoogle Scholar
Smith, E. C. and Griffiths, H. (1996). The occurrence of the chloroplast pyrenoid is correlated with the activity of a CO2 concentrating mechanism and carbon isotope discrimination in lichens and bryophytes. Planta, 198, 6–16.CrossRefGoogle Scholar
Snelgar, W. P. (1981). The ecophysiology of New Zealand forest lichens with special reference to carbon dioxide exchange. Ph.D. thesis, Waikato University, Hamilton, New Zealand.
Snelgar, W. P. and Green, T. G. A. (1980). Carbon dioxide exchange in lichens. II. Low carbon dioxide compensation levels and lack of apparent photorespiratory activity in some lichens. Bryologist, 83, 505–507.CrossRefGoogle Scholar
Snelgar, W. P. and Green, T. G. A. (1981). Carbon dioxide exchange in lichens: apparent photorespiration and the possible role of CO2 refixation in some members of the Stictaceae (Lichenes). Journal of Experimental Botany, 32, 661–668.CrossRefGoogle Scholar
Snelgar, W. P. and Green, T. G. A. (1982). Growth rates of Stictaceae lichens in New Zealand beech forests. Bryologist, 85, 301–306.CrossRefGoogle Scholar
Snelgar, W. P., Green, T. G. A. and Beltz, C. K. (1981 a). Carbon dioxide exchange in lichens: estimation of internal thallus CO2 transport resistances. Physiologia Plantarum, 52, 417–422.CrossRefGoogle Scholar
Snelgar, W. P., Green, T. G. A. and Wilkins, A. L. (1981 b). Carbon dioxide exchange in lichens. I. Resistances to CO2 uptake at different thallus water contents. New Phytologist, 88, 353–361.CrossRefGoogle Scholar
Søchting, U. (1995). Lichens as monitors of nitrogen deposition. Cryptogamic Botany, 5, 264–269.Google Scholar
Søchting, U. and Johnsen, I. (1978). Lichen transplants as biological indicators of SO2 air pollution in Copenhagen. Bulletin of Environmental Contamination and Toxicology, 19, 1–7.CrossRefGoogle ScholarPubMed
Solhaug, K. A. and Gauslaa, Y. (1996). Parietin, a photoprotective secondary product of the lichen Xanthoria parietina. Oecologia, 108, 412–418.CrossRefGoogle ScholarPubMed
Solhaug, K. A. and Gauslaa, Y. (2004 a). Testing of ecological roles of secondary lichen compounds. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 40. Tartu: Tartu University Press.Google Scholar
Solhaug, K. A. and Gauslaa, Y. (2004 b). Photosynthates stimulate the UV-B induced fungal anthraquinone synthesis in the foliose lichen Xanthoria parietina. Plant Cell and Environment, 27, 167–176.CrossRefGoogle Scholar
Solhaug, K. A., Gauslaa, Y., Nybakken, L. and Bilger, W. (2003). UV-induction of sun-screening pigments in lichens. New Phytologist, 158, 91–100.CrossRefGoogle Scholar
Solomina, O. and Calkin, P. E. (2003). Lichenometry as applied to moraines in Alaska, U.S.A., and Kamchatka, Russia. Arctic, Antarctic, and Alpine Research, 35, 129–143.CrossRefGoogle Scholar
Sommerkorn, M. (2000). The ability of lichens to benefit from natural CO2 enrichment under a spring snow-cover: a study with two arctic-alpine species from contrasting habitats. Bibliotheca Lichenologica, 75, 365–380.Google Scholar
Sonesson, M., Osborne, C. and Sandberg, G. (1994). Epiphytic lichens as indicators of snow depth. Arctic and Alpine Research, 26, 159–165.CrossRefGoogle Scholar
Sonesson, M., Schipperges, B. and Carlsson, B. Å. (1992). Seasonal patterns of photosynthesis in alpine and subalpine populations of the lichen Nephroma arcticum. Oikos, 65, 3–12.CrossRefGoogle Scholar
Spiro, B., Morrisson, J. and Purvis, O. W. (2002). Sulphur isotopes in lichens as indicators of sources. In Monitoring with Lichens – Monitoring Lichens. Nato Science Series IV: Earth and Environmental Sciences, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 311–315. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Clair, St. L. L. and Seaward, M. R. D., eds. (2004). Biodeterioration of Stone Surfaces. Dordrecht: Kluwer.CrossRefGoogle Scholar
Clair, St., Johansen, L. L., , J. R. and Rushforth, S. R. (1993). Lichens of soil crust communities in the intermountain area of the western United States. Great Basin Naturalist, 53, 5–12.Google Scholar
Staiger, B. (2002). Die Flechtenfamilie Graphidaceae. Studien in Richtung einer natürlicheren Gliederung. Bibliotheca Lichenologica, 85, 1–526.Google Scholar
Staiger, B., Kalb, K. and Grube, M. (2006). Phylogeny and phenotypic variation in the lichen family Graphidaceae (Ostropomycetidae, Ascomycota). Mycological Research, 110, 765–772.CrossRefGoogle Scholar
Stålfelt, M. G. (1939). Der Gasaustausch der Flechten. Planta, 29, 11–31.CrossRefGoogle Scholar
Stamatakis, A. (2006). RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics, 22, 2688–2690.CrossRefGoogle ScholarPubMed
Stamp, N. (2004). Can the growth-differentiation balance hypothesis be tested rigorously?Oikos, 107, 439–448.CrossRefGoogle Scholar
Steinkötter, J., Bhattacharya, D., Semmelroth, I., Bibeau, C. and Melkonian, M. (1994). Prasinophytes form independent lineages within the Chlorophyta: evidence from ribosomal RNA sequence comparisons. Journal of Phycology, 30, 340–345.CrossRefGoogle Scholar
Steinnes, E. and Krog, H. (1977). Mercury, arsenic and selenium fall-out from an industrial complex studied by means of lichen transplants. Oikos, 28, 160–164.CrossRefGoogle Scholar
Stenroos, S. (1993). Taxonomy and distribution of the lichen family Cladoniaceae in the Antarctic and peri-Antarctic regions. Cryptogamic Botany, 3, 310–344.Google Scholar
Stenroos, S. and DePriest, P. T. (1998). SSU rDNA phylogeny of cladoniiform lichens. American Journal of Botany, 85, 1548–1559.CrossRefGoogle ScholarPubMed
Stenroos, S., Feuerer, T. and Ahti, T. (2002 a). Chilean Cladoniaceae online. Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg, 30–32, 241–251.Google Scholar
Stenroos, S., Hyvönen, J., Myllys, L, Thell, A. and Ahti, T. (2002 b). Phylogeny of the genus Cladonia s. lat. (Cladoniaceae, Ascomycetes) inferred from molecular, morphological, and chemical data. Cladistics, 18, 237–278.CrossRefGoogle Scholar
Stenroos, S., Stocker-Wörgötter, E., Yoshimura, I., et al. (2003). Culture experiments and DNA sequence data confirm the identity of Lobaria photomorphs. Canadian Journal of Botany, 81, 232–247.CrossRefGoogle Scholar
Stewart, W. D. P. (1980). Some aspects of structure and function in N2-fixing cyanobacteria. Annual Reviews in Microbiology, 34, 497–536.CrossRefGoogle ScholarPubMed
Stewart, W. D. P. and Rodgers, G. A. (1978). Studies on the symbiotic blue-green algae of Anthoceros, Blasia, and Peltigera. In Environmental Role of Nitrogen-fixing Blue-green Algae and Asymbiotic Bacteria, ed. Granhall, U., pp. 247–259. Stockholm: Swedish Natural Science Research Council.Google Scholar
Stewart, W. D. P. and Rowell, P. (1975). Effects of L-methionine DL-sulphoximine on the assimilation of newly fixed NH3, C2H2 reduction and heterocyst production in Anabaena cylindrica. Biochemical Biophysical Research Communications, 65, 846–856.CrossRefGoogle Scholar
Stewart, W. D. P. and Rowell, P. (1977). Modifications of nitrogen-fixing algae in lichen symbioses. Nature (London), 265, 371–372.CrossRefGoogle Scholar
Stewart, W. D. P., Fitzgerald, G. P. and Burris, R. H. (1967). In situ studies on N2 fixation using the acetylene reduction technique. Proceedings of the National Academy of Sciences, USA, 58, 2071–2088.CrossRefGoogle ScholarPubMed
Stocker, O. (1927). Physiologische und ökologische Untersuchungen an Laub- und Strauchflechten. Flora, 21, 334–415.Google Scholar
Stocker-Wörgötter, E. (1995). Experimental cultivation of lichens and lichen symbionts. Canadian Journal of Botany, 73, S579–S589.CrossRefGoogle Scholar
Stocker-Wörgötter, E. (2001). Experimental lichenology and microbiology of lichens: culture experiments, secondary chemistry of cultured mycobionts, resynthesis and thallus morphogenesis. Bryologist, 104, 576–581.CrossRefGoogle Scholar
Stocker-Wörgötter, E. (2002 a). Resynthesis of Photosymbiodemes. In Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 47–60. Berlin: Springer.Google Scholar
Stocker-Wörgötter, E. (2002 b). Analysis of secondary compounds in cultured mycobionts. In Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 296–306. Berlin: Springer.Google Scholar
Stocker-Wörgötter, E. (2005). Approaches to a biotechnology of lichen-forming fungi: induction of polyketide pathways and chemosyndromes in axenically cultured mycobionts. Recent Research Developments in Phytochemistry, 9, 115–131.Google Scholar
Stocker-Wörgötter, E. (2008). Metabolic diversity of lichen-forming ascomycetous fungi: culturing, polyketide and shikimate metabolite production, and PKS genes. Natural Product Reports, 25, 188–200.CrossRefGoogle ScholarPubMed
Stocker-Wörgötter, E. and Elix, J. A. (2004). Experimental studies of lichenized fungi: formation of rare depsides and dibenzofurans by the cultured mycobiont of Bunodophoron patagonicum (Sphaerophoraceae, lichenized Ascomycota). In Contributions to Lichenology, Festschrift in Honour of Hannes Hertel, ed. Döbbeler, P. and Rambold, G., pp. 659–669. Bibliotheca Lichenologica 88. Berlin: J. Cramer.Google Scholar
Stocker-Wörgötter, E. and Elix, J. A. (2006). Morphogenetic strategies and induction of secondary metabolite biosynthesis in cultured lichen-forming Ascomycota, as exemplified by Cladia retipora (Labill.) Nyl. and Dactylina arctica (Richards.) Nyl. Symbiosis, 41, 9–20.Google Scholar
Stocker-Wörgötter, E., Elix, J. A. and Grube, M. (2004). Secondary chemistry of lichen-forming fungi: chemosyndromic variation and DNA-analyses of cultures and chemotypes in the Ramalina farinacea complex. Bryologist, 107, 152–162.CrossRefGoogle Scholar
Stone, D. F. (1989). Epiphytic succession on Quercus garryana branches in the Willamette Valley of western Oregon. Bryologist, 92, 81–94.CrossRefGoogle Scholar
Stulen, I. and De Kok, L. J. (1993). Whole plant regulation of sulfur-metabolism – a theoretical approach and comparison with current ideas on regulation of nitrogen metabolism. In Sulfur Nutrition and Assimilation in Higher Plants, ed. Kok, L. J., Stulen, I., Rennenberg, H., Brunold, C. and Rauser, W. E., pp. 77–91. The Hague: SPB Academic Publishing.Google Scholar
Sugiyama, K., Kurokawa, S. and Okada, G. (1976). Studies on lichens as a bioindicator of air pollution. I. Correlation of distribution of Parmelia tinctorum with SO2 air pollution. Japanese Journal of Ecology, 26, 209–212.Google Scholar
Sun, H. J. and Friedmann, E. I. (2005). Communities adjust their temperature optima by shifting producer-to-consumer ratio, shown in lichens as models. I. Experimental verification. Microbial Ecology, 49, 528–535.CrossRefGoogle ScholarPubMed
Sun, H. J., DePriest, P. T., Gargas, A., Rossman, A. Y. and Friedmann, E. I. (2002). Pestalotiopsis maculans: a dominant parasymbiont in North American lichens. Symbiosis, 33, 215–226.Google Scholar
Sundberg, B., Ekblad, A, Näsholm, T. and Palmqvist, K. (1999 a). Lichen respiration in relation to active time, nitrogen and ergosterol concentrations. Functional Ecology, 13, 119–125.CrossRefGoogle Scholar
Sundberg, B., Lundberg, P., Ekblad, A. and Palmqvist, K. (1999 b). In vivo13C NMR spectroscopy of carbon fluxes in four diverse lichens. In Physiological Ecology of Lichen Growth, by Sundberg, B.. Ph.D. thesis paper VI. Umeå: Umeå University.Google Scholar
Sundberg, B., Näsholm, T. and Palmqvist, K. (2001). The effect of nitrogen on growth and key thallus components in the two tripartite lichens, Nephroma arcticum and Peltigera aphthosa. Plant, Cell and Environment, 24, 517–527.CrossRefGoogle Scholar
Sundberg, B., Palmqvist, K., Esseen, P.-A. and Renhorn, K.-E. (1997). Growth and vitality of epiphytic lichens. II. Modelling of carbon gain using field and laboratory data. Oecologia, 109, 10–18.CrossRefGoogle Scholar
Suryanarayanan, T. S., Thirunavukkarasu, N., Hariharan, G. N. and Balaji, P. (2005). Occurrence of non-obligate microfungi inside lichen thalli. Sydowia, 57, 120–130.Google Scholar
Svenning, M. M., Eriksson, T. and Rasmussen, U. (2005). Phylogeny of symbiotic cyanobacteria within the genus Nostoc based on 16S rDNA sequence analyses. Archiv für Mikrobiologie, 183, 19–26.CrossRefGoogle ScholarPubMed
Swanson, A. and Fahselt, D. (1997). Effects of ultraviolet light on the polyphenolics of Umbilicaria americana Poelt & Nash. Canadian Journal of Botany, 75, 284–289.CrossRefGoogle Scholar
Syers, J. K. and Iskander, I. K. (1973). Pedogenetic significance of lichens. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 225–248. New York: Academic Press.Google Scholar
Szabo, I., Bergantino, E. and Giacometti, G. M. (2005). Light and oxygenic photosynthesis: energy dissipation as a protection mechanism against photo-oxidation. EMBO Reports, 6, 629–634.CrossRefGoogle ScholarPubMed
Takala, K., Kaurenen, P. and Olkkonen, H. (1978). Fluorine content of two lichen species in the vicinity of a fertilizer factory. Annales Botanici Fennici, 15, 158–167.Google Scholar
Takala, K., Olkkonen, H., Ikonen, J., Jääskeläinen, J. and Puumalainen, P. (1985). Total sulphur contents of epiphytic and terricolous lichens in Finland. Annales Botanici Fennici, 22, 91–100.Google Scholar
Tanabe, Y., Watanabe, M. M. and Sugiyama, J. (2002). Are Microsporidia really related to Fungi?: a reappraisal based on additional gene sequences from basal fungi. Mycological Research, 106, 1380–1391.CrossRefGoogle Scholar
Tapper, R. (1976). Dispersal and changes in the local distribution of Evernia prunastri and Ramalina farinacea. New Phytologist, 77, 725–734.CrossRefGoogle Scholar
Tarhanen, S., Holopainen, T. and Oksanen, J. (1997). Ultrastructural changes and electrolyte leakage from ozone fumigated epiphytic lichens. Annals of Botany, 80, 611–621.CrossRefGoogle Scholar
Taylor, O. C. (1969). Importance of peroxyacetyl nitrate (PAN) as a phytotoxic air pollutant. Journal of the Air Pollution Control Association, 19, 347–351.CrossRefGoogle Scholar
Taylor, R. J. and Bell, M. (1983). Effects of SO2 on the lichen flora in an industrial area, northwest Whatcom County, Washington. Northwest Science, 57, 157–166.Google Scholar
Taylor, T. N., Hass, H. and Kerp, H. (1997). A cyanolichen from the Lower Devonian Rhynie Chert. American Journal of Botany, 84, 992–1004.CrossRefGoogle ScholarPubMed
Taylor, T. N., Hass, H., Remy, W. and Kerp, H. (1995 b). The oldest fossil lichen. Nature, 378, 244.CrossRefGoogle Scholar
Taylor, T. N., Remy, W., Hass, H. and Kerp, H. (1995 a). Fossil arbuscular mycorrhizae from the early Devonian. Mycologia, 87, 560–573.CrossRefGoogle Scholar
Taylor, W. A., Free, C., Boyce, C., Helgemo, R. and Ochoada, J. (2004). SEM analysis of Spongiophyton interpreted as a fossil lichen. International Journal of Plant Sciences, 165, 875–881.CrossRefGoogle Scholar
Tehler, A. (1982). The species pair concept in lichenology. Taxon, 31, 708–714.CrossRefGoogle Scholar
Tehler, A. (1988). A cladistic outline of the Eumycota. Cladistics, 4, 227–277.CrossRefGoogle Scholar
Tehler, A. (1990). A new approach to the phylogeny of Euascomycetes with a cladistic outline of Arthoniales focussing on Roccellaceae. Canadian Journal of Botany, 68, 2458–2592.CrossRefGoogle Scholar
Tehler, A. (1995). Morphological data, molecular data and total evidence in phylogenetic analysis. Canadian Journal of Botany, 73, S667–S676.CrossRefGoogle Scholar
Tehler, A. (1996). Systematics, phylogeny and classification. In Lichen Biology, ed. Nash, T. H. III, pp. 217–239. Cambridge: Cambridge University Press.Google Scholar
Tehler, A. and Irestedt, M. (2007). Parallel evolution of lichen growth forms in the family Roccellaceae (Arthoniales, Ascomycota). Cladistics, 23, 432–454.CrossRefGoogle Scholar
Tehler, A., Dahlkild, Å., Eldenäs, P. and Feige, G. B. (2004). The phylogeny and taxonomy of Macaronesian, European and Mediterranean Roccella (Roccellaceae, Arthoniales). Symbolae Botanicae Upsalienses, 34, 405–428.Google Scholar
Tehler, A., Farris, J. S., Lipscomb, D. L. and Källersjö, M. (2000). Phylogenetic analyses of the fungi based on large rDNA data sets. Mycologia, 92, 459–474.CrossRefGoogle Scholar
Tehler, A., Little, D. P. and Farris, J. S. (2003). The full-length phylogenetic tree from 1551 ribosomal sequences of chitinous fungi, Fungi. Mycological Research, 107, 901–916.CrossRefGoogle ScholarPubMed
Tel-Or, E. and Stewart, W. D. P. (1976). Photosynthetic electron transport, ATP synthesis and nitrogenase activity in isolated heterocysts of Anabaena cylindrica. Biochimica et Biophysica Acta, 423, 189–195.CrossRefGoogle ScholarPubMed
Tel-Or, E. and Stewart, W. D. P. (1977). Photosynthetic components of activities of nitrogen-fixing isolated heterocysts of Anabaena cylindrica. Proceedings of the Royal Society of London B, 198, 61–96.CrossRefGoogle Scholar
Théau, J. and Duguay, C. R. (2003). Mapping lichen changes in the summer range of the George River caribou herd (Québec-Labrador, Canada) using Landsat imagery (1976–1998). Rangifer, 24, 31–49.CrossRefGoogle Scholar
Théau, J. and Duguay, C. R. (2004). Lichen mapping in the summer range of the George River caribou herd using Landsat TM imagery. Canadian Journal of Remote Sensing, 30, 867–881.CrossRefGoogle Scholar
Théau, J., Peddle, D. R. and Duguay, C. R. (2005). Mapping lichen in a caribou habitat of northern Quebec, Canada, using an enhancement-classification method and spectral mixture analysis. Remote Sensing of Environment, 94, 232–243.CrossRefGoogle Scholar
Thell, A. and Goward, T. (1996). The new cetrarioid genus Kaernefeltia and related groups in the Parmeliaceae (lichenized Ascomycotina). Bryologist, 99, 125–136.CrossRefGoogle Scholar
Thell, A., Feuerer, T.Kärnefelt, I., Myllys, L. and Stenroos, S. (2004 a). Monophyletic groups within the Parmeliaceae identified by IST rDNA, β-tubulin and GAPDH sequences. Mycological Progress, 3, 297–314.CrossRefGoogle Scholar
Thell, A., Goward, T., Randlane, T., Kärnefewlt, E. I. and Saag, A. (1995). A revision of the North American lichen genus Ahtiana (Parmeliaceae). Bryologist, 98, 596–605.CrossRefGoogle Scholar
Thell, A., Randlane, T. and Saag, A. (2005). A new circumscription of the lichen genus Nephromopsis (Parmeliaceae, lichenized Ascomycetes). Mycological Progress, 4, 303–316.CrossRefGoogle Scholar
Thell, A., Westberg, M. and Kärnefelt, I. (2004 b). Biogeography of the lichen family Parmeliaceae in the Nordic countries with taxonomic remarks. Symbolae Botanicae Upsalienses, 34, 429–452.Google Scholar
Thomas, J., Wolk, C. P., Shaffer, P. W., Austin, S. M. and Galonsky, A. (1975). The initial organic product of fixation of 15N-labeled nitrogen gas by the blue-green alga Anabaena cylindrica. Biochemical Biophysical Research Communication, 67, 501–507.CrossRefGoogle Scholar
Thomas, M. A. (1999). Effect of sulfur dioxide and ozone on glutathione reductase and superoxide dismutase in lichens. Ph.D. dissertation. Tempe: Arizona State University.
Thomas, M. A., Romagni, J. G., Gries, C. and Nash, T. H. III (1997). The effects of sulfur dioxide exposure on glutathione reductase activity in the cyanolichen Peltigera canina. Supplement to Plant Physiology, 114, 57.Google Scholar
Thomson, J. W. (1972). Distributional patterns in American Arctic lichens. Canadian Journal of Botany, 50, 1135–1156.CrossRefGoogle Scholar
Tibell, L. (1984). A reappraisal of the taxonomy of Caliciales. Nova Hedwigia, Beiheft, 79, 597–713.Google Scholar
Tibell, L. (1998). Practice and prejudice in lichen classification. Lichenologist, 30, 439–453.CrossRefGoogle Scholar
Tibell, L. (1999). Calicioid lichens and fungi. Nordic Lichen Flora, 1, 20–71.Google Scholar
Tibell, L. (2001). Photobiont association and molecular phylogeny of the lichen genus Chaenotheca. Bryologist, 104, 191–198.CrossRefGoogle Scholar
Tibell, L. (2003). Tholurna dissimilis and generic delimitations in Caliciaceae inferred from nuclear IST and LSU rDNA phylogenies (Lecanorales, lichenized Ascomycetes). Mycological Research, 107, 1403–1418.CrossRefGoogle Scholar
Tibell, L. and Beck, A. (2001). Morphological variation, photobiont association and ITS phylogeny of Chaenotheca phaeocephala and C. subroscida (Coniocybaceae, lichenized Ascomycetes). Nordic Journal of Botany, 22, 651–660.CrossRefGoogle Scholar
Timdal, E. (1991). A monograph of the genus Toninia (Lecideaceae, Ascomycetes). Opera Botanica, 110, 1–137.Google Scholar
Timdal, E. and Tønsberg, T. (2006). Psoroma paleaceum comb. nov. the only hairy Psoroma in northern Europe. Graphis Scripta, 18, 54–57.Google Scholar
Timoney, K. P. and Marsh, J. E. (2004). Lichen trimlines in northern Alberta: establishment, growth rates, and historic water levels. Bryologist, 107, 429–440.CrossRefGoogle Scholar
Tomassini, F. D., Lavoie, P., Puckett, K. J., Nieboer, E. and Richardson, D. H. S. (1977). The effect of time of exposure to sulphur dioxide on potassium loss from and photosynthesis in the lichen, Cladina rangiferina (L.) Harm. New Phytologist, 79, 147–155.CrossRefGoogle Scholar
Tomassini, F. D., Puckett, K. J., Nieboer, E., Richardson, D. H. S. and Grace, B. (1976). Determination of copper, iron, nickel, and sulphur by x-ray fluorescence in lichens from the Mackenzie Valley, Northwest Territories, and the Sudbury District, Ontario. Canadian Journal of Botany, 54, 1591–1603.CrossRefGoogle Scholar
Tomitani, A., Knoll, A. H., Cavanaugh, C. M. and Ohna, T. (2006). The evolutionary diversification of cyanobacteria: molecular-phylogenetic and paleontological perspectives. Proceedings of the National Academy of Sciences, USA, 103, 5442–5447.CrossRefGoogle ScholarPubMed
Tønsberg, T. (1992). The sorediate and isidiate, corticolous, crustose lichens in Norway. Sommerfeltia, 14, 1–331.Google Scholar
Tønsberg, T. and Holtan-Hartwig, J. (1983). Phycotype pairs in Nephroma, Peltigera and Lobaria in Norway. Nordic Journal of Botany, 3, 681–688.CrossRefGoogle Scholar
Topham, P. B. (1977). Colonization, growth, succession and competition. In Lichen Ecology, ed. Seaward, M. R. D., pp. 31–68. London: Academic Press.Google Scholar
Tormo, R., Recio, D., Silva, I. and Muñoz, A. (2001). A quantitative investigation of airborne algae and lichen soredia obtained from pollen traps in south-west Spain. European Journal of Phycology, 36, 385–390.CrossRefGoogle Scholar
Trembley, M. L., Ringli, C. and Honegger, R. (2002 a). Differential expression of hydrophobins DGH1, DGH2 and DGH3 and immunolocalization of DGH1 in strata of the lichenized basidiocarp of Dictyonema glabratum. New Phytologist, 154, 185–195.CrossRefGoogle Scholar
Trembley, M. L., Ringli, C. and Honegger, R. (2002 b). Hydrophobins DGH1, DGH2, and DGH3 in the lichen-forming basidiomycete Dictyonema glabratum. Fungal Genetics and Biology, 35, 247–259.CrossRefGoogle ScholarPubMed
Trembley, M. L., Ringli, C. and Honegger, R. (2002 c). Morphological and molecular analysis of early stages in the resynthesis of the lichen Baeomyces rufus. Mycological Research, 106, 768–776.CrossRefGoogle Scholar
Tretiach, M. and Carpanelli, A. (1992). Chlorophyll content and morphology as factors influencing the photosynthetic rate of Parmelia caperata. Lichenologist, 24, 81–90.Google Scholar
Triebel, D. and Rambold, G. (1988). Cecidonia und Phacopsis (Lecanorales): zwei lichenicole Pilzgattungen mit cecidogenen Arten. Nova Hedwigia, 47, 279–309.Google Scholar
Tschermak, E. (1941). Untersuchungen über die Beziehungen von Pilz und Alge im Flechtenthallus. Österreichische Botanische Zeitschrift, 90, 233–307.CrossRefGoogle Scholar
Tschermak-Woess, E. (1976). Algal taxonomy and the taxonomy of lichens: the phycobiont of Verrucaria adriatica. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 79–87. Orlando: Academic Press.Google Scholar
Tschermak-Woess, E. (1978). Myrmecia reticulata as a phycobiont and free-living Trebouxia – the problem of Stenocybe septata. Plant Systematics and Evolution, 129, 185–208.CrossRefGoogle Scholar
Tschermak-Woess, E. (1980). Chaenothecopsis consociata – kein parasitischer oder parasymbiotischer Pilz, sondern lichenisiert mit Dictyochloropsis symbiontica, spec. nova. Plant Systematics and Evolution, 136, 287–306.CrossRefGoogle Scholar
Tschermak-Woess, E. (1988). The algal partner. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 39–92. Boca Raton: CRC Press.Google Scholar
Tschermak-Woess, E. and Poelt, J. (1976). Vezdea, a peculiar lichen genus, and its phycobiont. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 89–105. Orlando: Academic Press.Google Scholar
Tuba, Z., Csintalan, Z., Szente, K., Nagy, Z. and Grace, J. (1998). Carbon gains by desiccation-tolerant plants at elevated CO2. Functional Ecology, 12, 39–44.CrossRefGoogle Scholar
Tuba, Z., Proctor, M. C. F. and Takács, Z. (1999). Desiccation-tolerant plants under elevated air CO2: a review. Zeitschrift für Naturforschung, Section C, 54, 788–796.Google Scholar
Tupa, D. D. (1974). An investigation of certain chaetophoralean algae. Nova Hedwigia, 46, 1–155.Google Scholar
Turgeon, B. and Yoder, O. (2000). Proposed nomenclature for mating type genes in filamentous ascomycetes. Fungal Genetics and Biology, 31, 1–5.CrossRefGoogle ScholarPubMed
Türk, R. and Christ, R. (1980). Untersuchungen des CO2-Gaswechsels von Flechtenexplantaten zur Indikation von SO2-Belastung im Stadtgebiet von Salzburg. In Bioindikation auf subzellularen und zellular Ebene (Bioindikation 2), ed. Schubert, R. and Schuh, J., pp. 39–45. Halle-Wittenberg: Martin-Luther-Universitat.Google Scholar
Türk, R. and Wirth, V. (1975). The pH dependence of SO2 damage to lichens. Oecologia, 19, 285–291.CrossRefGoogle Scholar
Türk, R., Wirth, V. and Lange, O. L. (1974). CO2-Gaswechsel-Untersuchungen zur SO2-Resistenz von Flechten. Oecologia, 15, 33–64.Google Scholar
Turner, S., Huang, T. C., and Chaw, S.-M. (2001). Molecular phylogeny of nitrogen-fixing unicellular cyanobacteria. Botanical Bulletin of Academia Sinica, 42, 181–186.Google Scholar
Turner, W. B. and Aldridge, D. C. (1983). Fungal Metabolites II. London: Academic Press.Google Scholar
US EPA (1996). Air Quality Criteria for Ozone and Related Photochemical Oxidants. Vol. II. EPA/600/p-93/004bF. Research Triangle Park, N.C. National Center for Environmental Assessment, Office of Research and Development.
Vainio, E. A. (1890). Étude sur classification naturelle et morphologie des lichens du Brésil. Acta Societatis pro Fauna et Flora Fennica, Helsinki, 7, 1–256.Google Scholar
Valladares, F., Sancho, L. G. and Ascaso, C. (1997). Water storage in the lichen family Umbilicariaceae. Botanica Acta, 111, 1–9.Google Scholar
Hoek, C., Jahns, H. M. and Mann, D. G. (1993). Algen, 3. Auflage. Stuttgart: Thieme.Google Scholar
Eerden, L., Vries, W. and Dobben, H. (1998). Effects of ammonia deposition on forests in The Netherlands. Atmospheric Environment, 32, 525–532.CrossRefGoogle Scholar
van Dobben, H. (1993) Vegetation as a monitor for deposition of nitrogen and acidity. Ph.D. dissertation, University of Utrecht.
Dobben, H. F. (1996). Decline and recovery of epiphytic lichens in an agricultural area in The Netherlands (1900–1988). Nova Hedwigia, 62, 477–485.Google Scholar
Dobben, H. F. and Bakker, A. J. (1996). Re-mapping epiphytic lichen biodiversity in The Netherlands: effects of decreasing SO2 and increasing NH3. Acta Botanica Neerlandica, 45, 55–71.CrossRefGoogle Scholar
Dobben, H. F. and Braak, C. J. F. (1998). Effects of atmospheric NH3 on epiphytic lichens in The Netherlands: the pitfalls of biological monitoring. Atmospheric Environment, 32, 551–557.CrossRefGoogle Scholar
Dobben, H. F. and Braak, C. J. F. (1999). Ranking of epiphytic lichen sensitivity to air pollution using survey data: a comparison of indicator scales. Lichenologist, 31, 27–39.Google Scholar
van Herk, C. M. (2002). Epiphytes on wayside trees as an indicator of eutrophication in the Netherlands. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 285–289. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Herk, C. M., Aptroot, A. and Dobben, H. F. (2002). Long-term monitoring in the Netherlands suggests that lichens respond to global warming. Lichenologist, 34, 141–154.CrossRefGoogle Scholar
Vězda, A. (1979). Flechtensystematische Studien. XI. Beiträge zur Kenntnis der Familie Asterothyriaceae (Discolichenes). Folia Geobotanica Phytotaxonomica Bohemoslovaca, Praha, 14, 43–94.CrossRefGoogle Scholar
Vězda, A. (1980). Foliikole Flechten aus Zaire. Die Arten der Sammelgattungen Catillaria und Bacidia. Folia Geobotanica Phytotaxonomica Bohemoslovaca, Praha, 15, 75–94.CrossRefGoogle Scholar
Villeneuve, J. P. and Holm, E. (1984). Atmospheric background of chlorinated hydrocarbons studied in Swedish lichens. Chemosphere, 13, 1133–1138.CrossRefGoogle Scholar
Virtala, M. (1992). Optimal harvesting of a plant–herbivore system: lichen and reindeer in northern Finland. Ecological Modelling, 60, 233–255.CrossRefGoogle Scholar
Vitikainen, O. (1998). Taxonomic notes on neotropical species of Peltigera. In Lichenology in Latin America: History, Current Knowledge and Applications, ed. Marcelli, M. P. and Seaward, M. R. D., pp. 135–139. São Paulo: CETESB.Google Scholar
Vitikainen, O. (2001). William Nylander (1822–1899) and lichen chemotaxonomy. Bryologist, 104, 263–267.CrossRefGoogle Scholar
Vitousek, P. M., Walker, L. R., Whittaker, L. D., Mueller-Dombois, D. and Matson, P. A. (1987). Biological invasion by Myrica faga alters ecosystem development in Hawaii. Science, 238, 802–804.CrossRefGoogle Scholar
Vobis, G. and Hawksworth, D. L. (1981). Conidial lichen-forming fungi. In The Biology of Conidial Fungi, ed. Cole, G. T. and Kendrick, B., pp. 245–273. New York: Academic Press.Google Scholar
Vogel, H. J. (1964). Distribution of lysine pathways among fungi: evolutionary implications. American Naturalist, 98, 435–446.CrossRefGoogle Scholar
Vogel, S. (1955). “Niedere Fensterpflanzen” in der südafrikanischen Wüste. Eine ökologische Schilderung. Beiträge zur Biologie der Pflanzen, 31, 45–135.Google Scholar
Arb, C., Mueller, C., Ammann, K. and Brunold, C. (1990). Lichen physiology and air pollution. II. Statistical analysis of the correlation between SO2, NO2, NO and O3, and chlorophyll content, net photosynthesis, sulphate uptake and protein synthesis of Parmelia sulcata Taylor. New Phytologist, 115, 431–437.CrossRefGoogle Scholar
Vráblíková, H., McEvoy, M., Solhaug, K. A., Barták, M. and Gauslaa, Y. (2006). Annual variation in photoacclimation and photoprotection of the photobiont in the foliose lichen Xanthoria parietina. Journal of Photochemistry and Photobiology B: Biology, 83, 151–162.CrossRefGoogle Scholar
Wachtmeister, C. A. (1956). Identification of lichen acids by paper chromatography, Botaniser Notiser, 109, 313–324.Google Scholar
Wainright, P. O., Hinkle, G., Sogin, M. L. and Stickel, S. K. (1993). Monophyletic origins of the Metazoa: an evolutionary link with fungi. Science, 260, 340–342.CrossRefGoogle ScholarPubMed
Walker, D. A., Webber, P. J., Everett, K. R. and Brown, J. (1978). Effects of crude and diesel oil spills on plant communities at Prudhoe Bay, Alaska, and the derivation of oil spill sensitivity maps. Arctic, 31, 242–259.CrossRefGoogle Scholar
Walker, F. J. (1985). The lichen genus Usnea subgenus Neuropogon. Bulletin of the British Museum (Natural History), 13, 1–130.Google Scholar
Walker, T. R., Crittenden, P. D. and Young, S. D. (2003). Regional variation in the chemical composition of winter snow pack and terricolous lichens in relation to sources of acid emissions in the Usa River basin, northeast European Russia. Environmental Pollution, 125, 401–412.CrossRefGoogle ScholarPubMed
Walser, J.-C., Holderegger, R., Gugerli, F., Hoebee, S. E. and Scheidegger, C. (2005). Microsatellites reveal regional population differentiation and isolation in Lobaria pulmonaria, an epiphytic lichen. Molecular Ecology, 14, 457–468.CrossRefGoogle ScholarPubMed
Walser, J.-C., Sperisen, C., Soliva, M. and Scheidegger, C. (2003). Fungus-specific microsatellite primers of lichens: application for the assessment of genetic variation on different spatial scales in Lobaria pulmonaria. Fungal Genetics and Biology, 40, 72–82.CrossRefGoogle ScholarPubMed
Warén, H. (1918–19) [1920]. Reinkulturen von Flechtengonidien. Öfversigt af Finska Vetenskaps-Societetens Förhandlingar (Helsingfors), 61, 1–79.Google Scholar
Waring, R. H. and Schlesinger, W. H. (1985). Forest Ecosystems: Concepts and Management. Orlando: Academic Press.Google Scholar
Warren, S. D. and Eldridge, D. J. (2001). Biological soil crusts and livestock in arid ecosystems: are they compatible?. In Biological Soil Crusts: Structure, Function and Management, ed. Belnap, J. and Lange, O. L., pp. 401–415. Berlin: Springer.Google Scholar
Washburn, S. (2005 [2006]). The Epiphytic Macrolichens of the Greater Cincinnati, Ohio, Metropolitan Area. M.S. thesis. Cincinnati: University of Cincinnati.
Washburn, S. (2006). Ozone exposure indices correlated with lichen abundance data in greater Cincinnati metropolitan area, Ohio. Botany 2006 Abstracts, p. 51. St. Louis: Botanical Society of America.
Watanabe, A. (1960). List of algal strains in collection at the Institute of Applied Microbiology, University of Tokyo. Journal of General and Applied Microbiology, 6, 283–292.CrossRefGoogle Scholar
Waterbury, J. B. and Stanier, R. Y. (1978). Patterns of growth and development in pleurocapsalean cyanobacteria. Microbiological Reviews, 42, 2–44.Google ScholarPubMed
Waters, J. M. and Craw, D. (2006). Goodby Gondwana? New Zealand biogeography, geology, and the problem of circularity. Systematic Biology, 55, 351–356.CrossRefGoogle Scholar
Weber, W. A. (1977). Environmental modification and lichen taxonomy. In Lichen Ecology, ed. Seaward, M. R. D., pp. 9–29. London: Academic Press.Google Scholar
Weber, W. A. (2003). The middle Asian element in the southern Rocky Mountain flora of the western United States: a critical biogeographical review. Journal of Biogeography, 30, 649–685.CrossRefGoogle Scholar
Wedin, M. (1995). The lichen family Sphaerophoraceae (Caliciales, Ascomycotina) in temperate areas of the Southern Hemisphere. Symbolae Botanicae Upsalienses, 31, 1–102.Google Scholar
Wedin, M. and Döring, H. (1999). The phylogenetic relationship of the Spaerophoraceae, Austropeltum and Neophyllis (lichenized Ascomycota) inferred by SSU rDNA sequences. Mycological Research, 109, 1131–1137.CrossRefGoogle Scholar
Wedin, M. and Wiklund, E. (2004). The phylogenetic relationships of Lecanorales suborder Peltigerineae revisited. Symbolae Botanicae Upsalienses, 34, 469–475.Google Scholar
Wedin, M., Döring, H. and Ekman, S. (2000 a). Molecular phylogeny of the lichen families Cladoniaceae, Sphaerophoraceae, and Stereocaulaceae (Lecanorales, Ascomycotina). Lichenologist, 32, 171–187.CrossRefGoogle Scholar
Wedin, M., Döring, H. and Gilenstam, G. (2004). Saprotrophy and lichenization as options for the same fungal species on different substrata: environmental plasticity and fungal lifestyles in the Stictis-Conotrema complex. New Phytologist, 164, 459–465.CrossRefGoogle Scholar
Wedin, M., Döring, H., Könberg, K. and Gilenstam, G. (2005 a). Generic delimitations in the family Stictidaceae (Ostropales, Ascomycota): the Stictis-Conotrema problem. Lichenologist, 37, 67–75.CrossRefGoogle Scholar
Wedin, M., Döring, H., Nordin, A. and Tibell, L. (2000 b). Small subunit rDNA phylogeny shows the lichen families Caliciaceae and Physciaceae (Lecanorales, Ascomycotina) to form a monophyletic group. Canadian Journal of Botany, 78, 246–254.CrossRefGoogle Scholar
Wedin, M., Wiklund, E., Crewe, A., et al. (2005 b). Phylogenetic relationships of Lecanoromycetes (Ascomycota) as revealed by analyses of mtSSU and nLSU rDNA sequence data. Mycological Research, 109, 159–172.CrossRefGoogle ScholarPubMed
Wedin, M., Wiklund, E. and Jørgensen, P. M. (2007). Massalongiaceae, fam. nov., an overlooked monophyletic group among the cyanobacterial lichens (Peltigerales, Lecanoromycetes, Ascomycota). Lichenologist, 39, 61–67.CrossRefGoogle Scholar
Wein, R. W. and Speer, J. E. (1975). Lichen biomass in Acadian and boreal forests of Cape Breton Island, Nova Scotia. Bryologist, 78, 328–333.CrossRefGoogle Scholar
Weissman, J. C. and Benemann, J. R. (1977). Hydrogen products by nitrogen-starved cultures of Anabaena cylindrica. Applied Environmental Microbiology, 33, 123–131.Google Scholar
Weissman, L., Garty, J. and Hochman, A. (2005 a). Rehydration of the lichen Ramalina lacera results in production of reactive oxygen species and nitric oxide and a decrease in antioxidants. Applied and Environmental Microbiology, 71, 2121–2129.CrossRefGoogle Scholar
Weissman, L., Garty, J. and Hochman, A. (2005 b). Characterization of enzymatic antioxidants in the lichen Ramalina lacera and their response to rehydration. Applied and Environmental Microbiology, 71, 6508–6514.CrossRefGoogle ScholarPubMed
Werth, S., Wagner, H. H., Holderegger, J. M., Kalwij, J. and Scheidegger, C. (2005). Genetic diversity of an old-forest associated lichen is affected by stand-replacing disturbances. In Dispersal and Persistence of an Epiphytic Lichen in a Dynamic Pasture-woodland Landscape. Ph.D. dissertation, by S. Werth, pp. 23–48. Bern: Universität Bern.Google Scholar
Wessels, D. C. J. and Schoeman, P. (1988). Mechanism and rate of weathering of Clarens sandstone by an endolithic lichen. South African Journal of Science, 84, 274–277.Google Scholar
Wessels, D. C. J. and Wessels, L. A. (1991). Erosion of biogenically weathered Clarens sandstone by lichenophagous bagworm larvae (Lepidoptera: Psychidae). Lichenologist, 23, 283–291.CrossRefGoogle Scholar
Wessels, J. G. H. (1999). Fungi in their own right. Fungal Genetics and Biology, 27, 134–145.CrossRefGoogle ScholarPubMed
Wetmore, C. M. (1973). Multiperforate septa in lichens. New Phytologist, 72, 535–538.CrossRefGoogle Scholar
White, F. J. and James, P. W. (1985). A new guide to microchemical techniques for the identification of lichen substances. British Lichen Society Bulletin, 57 (supp.), 1–41.Google Scholar
Whiteford, J. and Spanu, P. (2002). Hydrophobins and the interactions between fungi and plants. Molecular Plant Pathology, 3, 391–400.CrossRefGoogle ScholarPubMed
Whiton, J. C. and Lawrey, J. D. (1984). Inhibition of crustose lichen spore germination by lichen acids. Bryologist, 87, 42–43.CrossRefGoogle Scholar
Wiklund, E. and Wedin, M. (2003). The phylogenetic relationships of the cyanobacterial lichens in the Lecanorales suborder Peltigerineae. Cladistics, 19, 419–431.CrossRefGoogle Scholar
Will-Wolf, S. (1980). Structure of corticolous lichen communities before and after exposure to emissions from a “clean” coal-fired generating station. Bryologist, 83, 281–295.CrossRefGoogle Scholar
Will-Wolf, S. (2002). Monitoring regional status and trends in forest health with lichen communities: the United States Forest Service approach. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 353–357. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Wilmotte, A. and Golubic, S. (1991). Morphological and genetic criteria in the taxonomy of Cyanophyta/Cyanobacteria. Algological Studies, 64, 1–24.Google Scholar
Winchester, V. and Harrison, S. (1994). A development of the lichenometric method applied to the dating of glacially influenced debris flows in southern Chile. Earth Surface Processes and Landforms, 19, 137.CrossRefGoogle Scholar
Winkworth, R. C., Wagstaff, S. J., Glenny, D. and Lockhart, P. J. (2002). Plant dispersal N.E.W.S. from New Zealand. Trends in Ecology and Evolution, 17, 514–520.CrossRefGoogle Scholar
Winkworth, R. C., Wagstaff, S. J., Glenny, D. and Lockhart, P. J. (2005). Evolution of the New Zealand mountain flora: origins, diversification and dispersal. Organisms, Diversity and Evolution, 5, 237–247.CrossRefGoogle Scholar
Winner, W. E., Atkinson, C. J. and Nash, T. H. III. (1988). Comparisons of SO2 absorption capacities of mosses, lichens, and vascular plants in diverse habitats. Bibliotheca Lichenologica, 30, 217–230.Google Scholar
Wirth, V. (1972). Die Silikatflechten – Gemeinschaften im ausseralpinen Zentraleuropa. Dissertationes Botanicae, 17, 1–306.Google Scholar
Wirth, V. (1987). The influence of water relations on lichen SO2-resistance. In Progress and Problems in Lichenology in the Eighties, ed. Peveling, E., pp. 347–350. Bibliotheca Lichenologica 25. Berlin-Stuttgart: J. Cramer.Google Scholar
Wirth, V. (1993). Trenwende bei der Ausbreitung der anthropogen geförderten Flechte Lecanora conizaeoides? Phytocoenologia, 23, 625–636.CrossRefGoogle Scholar
Wirth, V. (1995) Die Flechten Baden-Würtembergs. Verbreitungsatlas. Vols. I and II. Stuttgart: Eugen Ulmer.Google Scholar
Wirth, V. (2001). Zeiberwerte von Flechten. Scripta Geobotanica, 18, 221–243.Google Scholar
Wirtz, N., Lumbsch, H. T., Green, T. G. A., et al. (2003). Lichen fungi have low cyanobiont selectivity in maritime Antarctica. New Phytologist, 160, 177–183.CrossRefGoogle Scholar
Wirtz, N., Printzen, C., Sancho, L. G. and Lumbsch, H. T. (2006). The phylogeny and classification of Neuropogon and Usnea (Parmeliaceae, Ascomycota) revisited. Taxon, 55, 367–376.CrossRefGoogle Scholar
Wise, M. J. and Tunnacliffe, A. (2004). POPP the question: what do LEA proteins do? Trends in Plant Science, 9, 13–17.CrossRefGoogle Scholar
Wolf, J. H. D. (1993). Diversity patterns and biomass of epiphytic bryophytes and lichens along an altitudinal gradient in the northern Andes. Annals of the Missouri Botanical Garden, 80, 928–960.CrossRefGoogle Scholar
Wolk, C. P., Ernst, A. and Elhai, J. (1994). Heterocyst metabolism and development. In Molecular Genetics of Cyanobacteria, ed. Bryant, D., pp. 769–823. Dordrecht: Kluwer Academic Publishing.CrossRefGoogle Scholar
Wolken, G. J. (2006). High-resolution multispectral techniques for mapping former Little Ice Age terrestrial ice cover in the Canadian High Arctic. Remote Sensing of Environment, 101, 104–114.CrossRefGoogle Scholar
Wolken, G. J., England, J. H. and Dyke, A. S. (2005). Re-evaluating the relevance of vegetation trimlines in the Canadian Arctic as an indicator of Little Ice Age paleoenvironments. Arctic, 58, 341–353.Google Scholar
Wolseley, P. A. (1995). A global perspective on the status of lichens and their conservation. Mitteilungen der Eidgenössischen Forshungsanstalt für Wald, Schnee und Landschaft, 70, 11–22.Google Scholar
Wolseley, P. A. (2002). Using corticolous lichens of tropical forests to assess environmental changes. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 373–378. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Wolseley, P. A., James, P. W., Theobald, M. R. and Sutton, M. A. (2006). Detecting changes in epiphytic lichen communities at sites affected by atmospheric ammonia from agricultural sources. Lichenologist, 38, 161–176.CrossRefGoogle Scholar
Wösten, H. A. B. (2001). Hydrophobins: multipurpose proteins. Annual Reviews in Microbiology, 55, 625–646.CrossRefGoogle ScholarPubMed
Wösten, H. A. B., Devries, O. M. H. and Wessels, J. G. H. (1993). Interfacial self-assembly of a fungal hydrophobin into a hydrophobic rodlet layer. Plant Cell, 5, 1567–1574.CrossRefGoogle ScholarPubMed
Yahr, R., Vilgalys, R. and DePriest, P. T. (2004). Strong fungal specificity and selectivity for algal symbionts in Florida scrub Cladonia lichens. Molecular Ecology, 13, 3367–3378.CrossRefGoogle ScholarPubMed
Yahr, R., Vilgalys, R. and DePriest, P. T. (2006). Geographic variation in algal partners of Cladonia subtenuis (Cladoniaceae) highlights the dynamic nature of a lichen symbiosis. New Phytologist, 171, 847–860.CrossRefGoogle ScholarPubMed
Yamamoto, Y. (1990). Studies of Cell Aggregates and the Production of Natural Pigments in Plant Cell Culture. Osaka: Nippon Paint Publication.Google Scholar
Yamamoto, Y., Kinoshita, Y. and Yoshimura, I. (2002). Culture of thallus fragments and redifferentiation of lichens. In Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 34–46. Berlin: Springer.CrossRefGoogle Scholar
Yoshimura, I. (1998). Lobaria in Latin America: taxonomic, geographic and evolutionary aspects. In Lichenology in Latin America: History, Current Knowledge and Applications, ed. Marcelli, M. P. and Seaward, M. R. D., pp. 129–134. São Paulo: CETESB.Google Scholar
Yoshimura, I. and Arvidsson, L. (1994). Taxonomy and chemistry of the Lobaria crenulata group in Ecuador. Acta Botanica Fennica, 150, 223–233.Google Scholar
Yoshimura, I., Yamamoto, Y., Nakano, T. and Finnie, J. (2002). Isolation and culture of lichen photobionts and mycobionts. Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 3–33. Berlin: Springer:CrossRefGoogle Scholar
Young, K. (2005). Hardy lichen shown to survive in space. New Scientist, November 10th, 2005.
Yuan, X., Xiao, S. and Taylor, T. N. (2005). Lichen-like symbiosis 600 million years ago. Science, 308, 1017–1020.CrossRefGoogle ScholarPubMed
Yun, S., Berbee, M., Yoder, O. and Turgeon, B. (1999). Evolution of the fungal self-fertile reproductive life style from self-sterile ancestors. Proceedings of the National Academy of Sciences, USA, 96, 5592–5597.CrossRefGoogle ScholarPubMed
Yurchenko, E. and Golubkov, V. (2003). The morphology, biology, and geography of a necrotrophic basidiomycete Athelia arachnoidea in Belarus. Mycological Progress, 2, 275–284.CrossRefGoogle Scholar
Zachariassen, K. E. and Kristiansen, E. (2000). Ice nucleation and antinucleation in nature. Cryobiology, 41, 257–279.CrossRefGoogle ScholarPubMed
Zahlbruckner, A. (1907). Die naturlichen Pflanzenfamilien, I Teil, 1 Apt, ed. Engler, A. and Prantl, K.. Leipzig: Borntraeger.Google Scholar
Zahlbruckner, A. (1926). Die naturlichen Pflanzenfamilien, Vol. 8, ed. Engler, A. and Prantl, K.. Leipzig: Borntraeger.Google Scholar
Zambrano, A., Nash, T. H. III and Gries, C. (2000). Responses of Ramalina farinacea (L.) Ach. to transplanting in southern California and to gaseous formaldehyde. Bibliotheca Lichenologica, 75, 219–230.Google Scholar
Zavarzina, A. and Zavarzin, A. (2006). Laccase and tyrosinase activity in lichens. Microbiology (Rus), 75, 630–641.Google ScholarPubMed
Zeitler, I. (1954). Untersuchungen über die Morphologie, Entwicklungsgeschichte und Systematik von Flechtengonidien. Österreichische Botanische Zeitschrift, 101, 453–487.CrossRefGoogle Scholar
Zhu, Y. Y., Machleder, E. M., Chenchik, A., Li, R. and Siebert, P. M. (2001). Reverse transcriptase template switching: a SMART™ approach for full-length cDNA library construction. BioTechniques, 30, 892–897.Google Scholar
Ziegler, I. (1977). Sulfite action on ribulosediphosphate carboxylase in the lichen Pseudevernia furfuracea. Oecologia, 29, 63–66.CrossRefGoogle ScholarPubMed
Zoller, S. and Lutzoni, F. (2003). Slow algae, fast fungi: exceptionally high nucleotide substitution rate differences between lichenized fungi Omphalina and their symbiotic green algae Coccomyxa. Molecular Phylogenetics and Evolution, 29, 629–640.CrossRefGoogle ScholarPubMed
Zoller, S., Lutzoni, F. and Scheidegger, C. (1999). Genetic variation within and among populations of the threatened lichen Lobaria pulmonaria in Switzerland and implications for its conservation. Molecular Ecology, 8, 2049–2059.CrossRefGoogle ScholarPubMed
Zopf, W. (1897). Über Nebensymbiose (Parasymbiose). Berichte der deutschen Botanischen Gesellschaft, 15, 90–92.Google Scholar
Zopf, W. (1907). Die Flechtenstoffe in chemischer, botanischer, pharmakologischer, und technischer Beziehung. Jena: Gustav Fischer.Google Scholar
Zotz, G. (1999). Altitudinal changes in diversity and abundance of non-vascular epiphytes in the tropics – an ecophysiological explanation. Selbyana, 20, 256–260.Google Scholar
Zotz, G. and Winter, K. (1994). Photosynthesis and carbon gain of the lichen, Leptogium azureum, in a lowland tropical forest. Flora, 189, 179–186.CrossRefGoogle Scholar
Zotz, G., Büdel, B., Meyer, A., Zellner, H. and Lange, O. L. (1998). In situ studies of water relations and CO2 exchange of the tropical macrolichen, Sticta tomentosa. New Phytologist, 139, 525–535.CrossRefGoogle Scholar
Zotz, G., Schultz, S. and Rottenberger, S. (2003). Are tropical lowlands a marginal habitat for macrolichens? Evidence from a field study with Parmotrema endosulphureum in Panama. Flora, 198, 71–77.CrossRefGoogle Scholar
Zukal, H. (1895). Morphologische und biologische Untersuchungen über die Flechten (II. Abhandlung). Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften Wien, Math.-naturw. Cl., 104/1, 1–68.Google Scholar
Aarrestad, P. A. and Aamlid, D. (1999). Vegetation monitoring in South-Varanger, Norway – species composition of ground vegetation and its relation to environmental variables and pollution impact. Environmental Monitoring and Assessment, 58, 1–21.CrossRefGoogle Scholar
Abba', S., Ghignone, S. and Bonfante, P. (2006). A dehydration-inducible gene in the truffle Tuber borchii identifies a novel group of dehydrins. BMC Genomics, 7, 39.CrossRefGoogle ScholarPubMed
Acharius, E. (1798). Lichenographiae Svecicae Prodromus. Linköping: D.G. Björn.CrossRefGoogle Scholar
Acharius, E. (1803). Methodus qua omnus detectos Lichenes. Stockholm: F.D.D. Ulrich.Google Scholar
Acharius, E. (1810). Lichenographia universalis. Göttingen: F. Dandewerts.Google Scholar
Acharius, E. (1814). Synopsis methodica lichenum. Lund.Google Scholar
Adams, G. C. and Kropp, B. R. (1996). Athelia arachnoidea, the sexual state of Rhizoctonia carotae, a pathogen of carrot in cold storage. Mycologia, 88, 459–472.CrossRefGoogle Scholar
Adler, M. T. and Calvelo, S. (2002). Parmeliaceae s. str. (lichenized Ascomycetes) from Tierra del Fuego (southern South America) and their world distribution patterns. Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg, 30–32, 9–24.Google Scholar
Adler, M. T., Fazio, A., Bertoni, M. D., et al. (2004). Culture experiments and DNA verification of a mycobiont isolated from Punctelia praesignis (Parmeliaceae, lichenized Ascomycotina). Bibliotheca Lichenologica, 88, 1–8.Google Scholar
Aguiar, L. W., Martau, L., Oliveira, M. L. A. A. and Martins-Mazzitelli, S. M. A. (1998). Efeitos do dióxido de enxofre (SO2) em liquens, Rio Grande do Sul, Brasil. Iheringia Botanica, 50, 67–73.Google Scholar
Ahmadjian, V. (1967 a). The Lichen Symbiosis. Toronto: Blaisdell Publishing.Google Scholar
Ahmadjian, V. (1967 b). A guide to the algae occurring as lichen symbionts: isolation, culture, cultural physiology and identification. Phycologia, 6, 127–160.CrossRefGoogle Scholar
Ahmadjian, V. (1973). Methods of isolating and culturing lichen symbionts and thalli. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 653–659. London: Academic Press.Google Scholar
Ahmadjian, V. (1988). The lichen alga Trebouxia: does it occur free-living?Plant Systematics and Evolution, 158, 243–247.CrossRefGoogle Scholar
Ahmadjian, V. (1993). The Lichen Symbiosis. New York: John Wiley.Google Scholar
Ahmadjian, V. (1995). Lichens are more important than you think. BioScience, 45, 123–124.CrossRefGoogle Scholar
Ahmadjian, V. (2001). Trebouxia: reflections on a perplexing and controversial lichen photobiont. In Symbiosis, ed. Seckbach, J., pp. 373–383. Dordrecht: Kluwer Academic.Google Scholar
Ahmadjian, V. and Heikkilä, H. (1970). The culture and synthesis of Endocarpon pusillum and Staurothelse clopima. Lichenologist, 4, 259–267.CrossRefGoogle Scholar
Ahmadjian, V. and Jacobs, J. B. (1981). Relationship between fungus and alga in the lichen Cladonia cristatella Tuck. Nature (London), 389, 169–172.CrossRefGoogle Scholar
Ahti, T. (1999). Biogeography. Nordic Lichen Flora, 1, 7–8.Google Scholar
Ahti, T. (2000). Cladoniaceae. Flora Neotropica Monograph, 78, 1–362.Google Scholar
Alebic-Juretic, A. and Arko-Pijevac, M. (1989). Air pollution damage to cell membranes in lichens – results of simple biological test applied in Rijeka, Yugoslavia. Water, Air, and Soil Pollution, 47, 25–33.CrossRefGoogle Scholar
Alexopoulos, C. J., Mims, C. W. and Blackwell, M. (1996). Introductory Mycology. 4th edn. New York: John Wiley.Google Scholar
Allan, A. and Fluhr, R. (1997). Two distinct sources of elicited reactive oxygen species in tobacco epidermal cells. Plant Cell, 9, 1559–1572.CrossRefGoogle ScholarPubMed
Allen, J. F., Mullineaux, C. W., Sanders, C. E. and Melis, A. (1989). State transitions, photosystem stoichiometry adjustment and non-photochemical quenching in cyanobacterial cells acclimated to light absorbed by photosystem I or photosystem II. Photosynthesis Research, 22, 157–166.CrossRefGoogle ScholarPubMed
Alpert, P. (1988). Survival of a desiccation-tolerant moss, Grimmia laevigata, beyond its observed microdistributional limits. Journal of Bryology, 15, 219–227.CrossRefGoogle Scholar
Alscher, R. (1984). Effects of SO2 on light-modulated enzyme reactions. In Gaseous Air Pollutants and Plant Metabolism, ed. Koziol, M. J. and Whatley, F. R., pp. 181–200. London: Butterworths.Google Scholar
Alstrup, V. (1992). Effects of pesticides on lichens. Bryonora, 9, 2–4.Google Scholar
Alstrup, V. and Hansen, E. S. (1977). Three species of lichens tolerant of high concentrations of copper. Oikos, 29, 290–293.CrossRefGoogle Scholar
Alstrup, V. and Hawksworth, D. L. (1990). The lichenicolous fungi from Greenland. Meddelelser om Grønland, Bioscience, 31, 1–90.Google Scholar
Amthor, J. S. (1995). Higher plant respiration and its relationships to photosynthesis. In Ecophysiology of Photosynthesis, ed. Schulze, E. D. and Caldwell, M. M., pp. 71–101. Berlin: Springer.Google Scholar
Anagnostidis, K. and Komárek, J. (1985). Modern approach to the classification system of cyanophytes. 1 – Introduction. Archiv für Hydrobiologie, Supplementband, 71 (Algological Studies, 38/39), 291–302.Google Scholar
Anagnostidis, K. and Komárek, J. (1988). Modern approach to the classification system of cyanophytes. 3 – Oscillatoriales. Archiv für Hydrobiologie, Supplementband, 80 (Algological Studies, 50/53), 327–472.Google Scholar
Anagnostidis, K. and Komárek, J. (1990). Modern approach to the classification system of cyanophytes. 5 – Stigonematales. Algological Studies, 59, 1–73.Google Scholar
Anand, M., Laurence, S. and Rayfield, B. (2005). Diversity relationships among taxonomic groups in recovering and restored forests. Conservation Biology, 19, 955–962.CrossRefGoogle Scholar
Andersen, H. L. and Ekman, S. (2005). Disintegration of the Micareaceae (lichenized Ascomycota): a molecular phylogeny based on mitochondrial rDNA sequences. Mycological Research, 109, 21–30.CrossRefGoogle ScholarPubMed
Anonymous (2004). Vagrant lichen charged in elk deaths. Castilleja, 23, 1.
Antoine, M. E. (2004). An ecophysiological approach to quantifying nitrogen fixation by Lobaria oregana. Bryologist, 107, 82–87.CrossRefGoogle Scholar
Antoine, M. E. and McCune, B. (2004). Contrasting fundamental and realized ecological niches with epiphytic lichen transplants in an old-growth Pseudotsuga forest. Bryologist, 107, 163–173.CrossRefGoogle Scholar
Aptroot, A. (1987). Terpenoids in tropical Pyxinaceae (lichenized fungi). In XIV International Botanical Congress, Abstracts, Berlin, ed. Greuter, W., Zimmer, B. and Behnke, H.-D., p. 5–04–7.Google Scholar
Aptroot, A. (2001). Lichenized and saprobic fungal biodiversity of a single Elaeocarpus tree in Papua New Guinea, with the report of 200 species of ascomycetes associated with one tree. Fungal Diversity, 6, 1–11.Google Scholar
Aptroot, A. and Seaward, M. R. D. (2003). Freshwater lichens. In Freshwater Mycology, ed. Tsui, C. K. and Hyde, K. D., pp. 101–110. Hong Kong: Fungal Diversity Press.Google Scholar
Aptroot, A. and Sipman, H. J. M. (1997). Diversity of lichenized fungi in the tropics. In Biodiversity of Tropical Microfungi, ed. Hyde, K. D., pp. 93–106. Hong Kong: Hong Kong University Press.Google Scholar
Aptroot, A. and Sparrius, L. B. (2006). Additions to the lichen flora of Vietnam, with an annotated checklist and bibliography. Bryologist, 109, 358–371.CrossRefGoogle Scholar
Archer, A. W. and Elix, J. A. (1993). Additional new taxa and a new report of Pertusaria (lichenised Ascomycotina) from Australia. Mycotaxon, 49, 143–150.Google Scholar
Archer, D., Eggink, G., Schweizer, M. Stymne, S. and Rathledge, G. (1999). Manipulation of Lipid Metabolism Aimed at the Production of Fatty Acids and Polyketides. Final report, Contr. No. Air 2-CT94-967. Internet Publication www.biomatnet.org.secure/Air/S1045.htm.
Archibald, P. A. (1975). Trebouxia DePuymaly (Chlorophyceae, Chlorococcales) and Pseudotrebouxia (Chlorophyceae, Chlorosarcinales). Phycologia, 14, 125–137.CrossRefGoogle Scholar
Armaleo, D. (1993). Why do lichens make secondary products? In XV International Botanical Congress Abstracts, ed. Furuya, M., p. 11. Yokohama, Japan: International Union of Biological Sciences.Google Scholar
Armaleo, D. and Clerc, P. (1990). Lichen chimeras: DNA analysis suggests that one fungus forms two morphotypes. Experimental Mycology, 15, 1–10.CrossRefGoogle Scholar
Armaleo, D. and Clerc, P. (1995). A rapid and inexpensive method for the purification of DNA from lichens and their symbionts. Lichenologist, 27, 207–213.CrossRefGoogle Scholar
Armstrong, R. A. (1974). A comparison of the growth-curves of the foliose lichen Parmelia conspersa determined by a cross-sectional study and by direct measurement. Environmental and Experimental Botany, 32, 221–227.CrossRefGoogle Scholar
Armstrong, R. A. (1988). Substrate colonization, growth, and competition. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 3–16. Boca Raton: CRC Press.Google Scholar
Armstrong, R. A. (1992). Soredial dispersal from individual soralia in the lichen Hypogymnia physodes (L.) Nyl. Environmental and Experimental Botany, 32, 55–63.CrossRefGoogle Scholar
Armstrong, R. A. (1993). Factors determining lobe growth in foliose lichen thalli. New Phytologist, 124, 675–679.CrossRefGoogle Scholar
Armstrong, R. A. and Smith, S. N. (1992). Lobe growth variation and the maintenance of symmetry in foliose lichen thalli. Symbiosis, 12, 145–158.Google Scholar
Aronson, J. M. (1977). Cell walls and intracellular polysaccharides of Leptomitales. Abstracts Second International Mycological Congress A–L, ed. Bigelow, H. E. and Simmons, E. G., p. 19. Tampa: IMC-2, Inc.Google Scholar
Arup, U. and Grube, M. (1998). Molecular systematics of Lecanora subgenus Placodium. Lichenologist, 30, 415–425.CrossRefGoogle Scholar
Arup, U., Ekman, S., Grube, M., Mattsson, J. E. and Wedin, M. (2007). The sister group relation of Parmeliaceae (Lecanorales, Ascomycota). Mycologia, 99, 42–49.CrossRefGoogle Scholar
Arup, U., Ekman, S., Lindblom, L. and Mattsson, J. E. (1993). High performance thin layer chromatography (HPTLC), an improved technique for screening lichen substances. Lichenologist, 25, 61–71.CrossRefGoogle Scholar
Asahina, Y. and Shibata, S. (1954). Chemistry of Lichen Substances. Tokyo: Japan Society for the Promotion of Science.Google Scholar
Ascaso, C., Wierzchos, J. and los Rios, A. (1995). Cytological investigations of lithobiontic microorganisms in granitic rocks. Botanica Acta, 108, 474–481.CrossRefGoogle Scholar
Aspray, T., Jones, E., Whipps, J. and Bending, G. (2006). Importance of mycorrhization helper bacteria cell density and metabolite localization for the Pinus sylvestris / Lactarius rufus symbiosis. FEMS Microbiology Ecology, 56, 25–33.CrossRefGoogle ScholarPubMed
Augusto, S., Pinho, P., Branquinho, C., et al. (2004). Atmospheric dioxin and furan deposition in relation to land-use and other pollutants: a survey with lichens. Journal of Atmospheric Chemistry, 49, 53–65.CrossRefGoogle Scholar
Avise, J. C. (1998). The history and purview of phylogeography: a personal reflection. Molecular Ecology, 7, 371–379.CrossRefGoogle Scholar
Bacci, E., Calamari, D., Gaggi, C., et al. (1986). Chlorinated hydrocarbons in lichen and moss samples from the Antarctic Peninsula. Chemosphere, 15, 747–754.CrossRefGoogle Scholar
Baćkor, M. and Dzubaj, A. (2004). Short-term and chronic effects of copper, zinc and mercury on the chlorophyll content of four lichen photobionts and related alga. Journal of the Hattori Botanical Laboratory, 95, 271–284.Google Scholar
Baćkor, M. and Fahselt, D. (2004). Physiological attributes of the lichen Cladonia pleurota in heavy metal-rich and control sites near Sudbury (Ont., Canada). Environmental and Experimental Botany, 52, 149–159.CrossRefGoogle Scholar
Baćkor, M. and Váczi, P. (2002). Copper tolerance in the lichen photobiont Trebouxia erici (Chlorophyta). Environmental and Experimental Botany, 48, 11–20.CrossRefGoogle Scholar
Baćkor, M. and Zetikova, J. (2003). Effects of copper, cobalt and mercury on the chlorophyll content of lichens Cetraria islandica and Flavocetraria cucullata. Journal of the Hattori Botanical Laboratory, 93, 175–187.Google Scholar
Baćkor, M., Dvorsky, K. and Fahselt, D. (2003). Influence of invertebrate feeding on the lichen Cladonia pocillum. Symbiosis, 34, 281–291.Google Scholar
Baddeley, M. S., Ferry, B. W. and Finegan, E. J. (1973). Sulphur dioxide and respiration in lichens. In Air Pollution and Lichens, ed. Ferry, B. W., Baddeley, M. S. and Hawksworth, D. L., pp. 299–313. Toronto: University of Toronto Press.Google Scholar
Badger, M. R., Pfanz, H., Büdel, B., Heber, U. and Lange, O. L. (1993). Evidence for the functioning of photosynthetic CO2 concentration mechanisms in lichens containing green algal and cyanobacterial photobionts. Planta, 191, 57–70.CrossRefGoogle Scholar
Bailey, R. H. (1966). Studies on the dispersal of lichen soredia. Journal of the Linnean Society, Botany, 59, 479–490.CrossRefGoogle Scholar
Bailey, R. H. (1976). Ecological aspects of dispersal and establishment in lichens. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 215–247. New York: Academic Press.Google Scholar
Balaguer, L. and Manrique, E. (1995). Factors which determine lichen response to chronic fumigations with sulphur dioxide. Cryptogamic Botany, 5, 215–219.Google Scholar
Balaguer, L., Manrique, E., los Rios, A., et al. (1999). Long-term responses of the green algal lichen Parmelia caperata to natural CO2 enrichment. Oecologia, 119, 166–174.CrossRefGoogle ScholarPubMed
Balaguer, L., Valladares, F., Ascaso, C., et al. (1996). Potential effects of rising tropospheric concentrations of CO2 and O3 on green-algal lichens. New Phytologist, 132, 641–652.CrossRefGoogle Scholar
Bargagli, R., Iosco, F. P. and Barghigiani, C. (1987). Assessment of mercury dispersal in an abandoned mining area by soil and lichen analysis. Water, Air, and Soil Pollution, 36, 219–225.CrossRefGoogle Scholar
Barghigiani, C., Bargagli, R., Siegel, B. Z. and Siegel, S. M. (1990). A comparative study of mercury distribution on the Aeolian volcanoes, Vulcano and Stromboli. Water, Air, and Soil Pollution, 53, 179–188.CrossRefGoogle Scholar
Barkman, J. J. (1958). Phytosociology and Ecology of Cryptogamic Epiphytes. Assen: Van Gorcum.Google Scholar
Barreno, E. (1991). Phytogeography of terricolous lichens in the Iberian Peninsula and the Canary Islands. Botanika Chronika, 10, 199–210.Google Scholar
Barreno, E., Grube, M., Bois, L., et al. (1998). Forum discussion. Lichens: a special case in biogeographical analysis. International Lichenologial Newsletter, 31, 18–24.Google Scholar
Barták, M., Solhaug, K. A., Vráblíková, H. and Gaulaa, Y. (2006). Curling during desiccation protects the foliose lichen Lobaria pulmonaria against photoinhibition. Oecologia, 149, 553–560.CrossRefGoogle ScholarPubMed
Bartók, K. (1999). Pesticide usage and epiphytic lichen diversity in Romanian orchards. Lichenologist, 31, 21–25.Google Scholar
Bates, J. W., Bell, J. N. B. and Farmer, A. M. (1990). Epiphyte recolonisation of oaks along a gradient of air pollution in south-east England, 1979–1990. Environmental Pollution, 68, 81–99.CrossRefGoogle Scholar
Bates, J. W., Bell, J. N. B. and Massara, A. C. (2001). Loss of Lecanora conizaeoides and other fluctuations of epiphytes on oak in S. E. England over 21 years with declining SO2 concentrations. Atmospheric Environment, 35, 2557–2568.CrossRefGoogle Scholar
Bauer, H. (1984). Net photosynthetic CO2 compensation concentrations of some lichens. Zeitschrift für Pflanzenphysiologie, 114, 45–50.CrossRefGoogle Scholar
Beard, K. H. and DePriest, P. T. (1996). Genetic variation within and among mats of the reindeer lichen, Cladina subtenuis. Lichenologist, 28, 171–182.CrossRefGoogle Scholar
Beck, A. (1999). Photobiont inventory of a lichen community growing on heavy-metal-rich rock. Lichenologist, 31, 501–510.CrossRefGoogle Scholar
Beck, A. (2002). Photobionts: diversity and selectivity in lichen symbioses. International Lichenological Newsletter, 35, 18–24.Google Scholar
Beck, A. and Koop, H. U. (2001). Analysis of the photobiont population in lichens using a single-cell manipulator. Symbiosis, 31, 57–67.Google Scholar
Beck, A., Friedl, T. and Rambold, G., (1998). Selectivity of photobiont choice in a defined lichen community: inferences from cultural and molecular studies. New Phytologist, 139, 709–720.CrossRefGoogle Scholar
Beck, A., Kasalicky, T. and Rambold, G. (2002). Myco-photobiontal selection in a Mediterranean cryptogam community with Fulgensia fulgida. New Phytologist, 153, 317–326.CrossRefGoogle Scholar
Beckelhimer, S. L. and Weaks, T. E. (1986). Effects of water transported sediment on corticolous lichen communities. Lichenologist, 18, 339–347.CrossRefGoogle Scholar
Becker, V. E. (1980). Nitrogen fixing lichens in forests of the southern Appalachian Mountains of North Carolina. Bryologist, 83, 29–39.CrossRefGoogle Scholar
Becker, V. E., Reeder, J. and Stetler, R. (1977). Biomass and habitat of nitrogen fixing lichens in an oak forest in the North Carolina Piedmont. Bryologist, 80, 93–99.CrossRefGoogle Scholar
Beckett, A. (1981). Ascospore formation. In The Fungal Spore: Morphogenetic Controls, ed. Turian, G. and Hohl, H. R., pp. 107–129. London: Academic Press.Google Scholar
Beckett, P. J., Boileau, L. J. R., Padovan, D., Richardson, D. H. S. and Nieboer, E. (1982). Lichens and mosses as monitors of industrial activity associated with uranium and lead accumulation patterns. Environmental Pollution (Series B), 4, 91–107.CrossRefGoogle Scholar
Beckett, R. P. (1995). Some aspects of the water relations of lichens from habitats of contrasting water status studied using thermocouple psychrometry. Annals of Botany, 76, 211–217.CrossRefGoogle Scholar
Beckett, R. P. (1997). Pressure-volume analysis of a range of poikilohydric plants implies the existence of negative turgor in vegetative cells. Annals of Botany, 79, 145–152.CrossRefGoogle Scholar
Beckett, R. P. and Brown, D. H. (1983). Natural and experimentally-induced zinc and copper resistance in the lichen genus Peltigera. Annals of Botany, 52, 43–50.CrossRefGoogle Scholar
Beckett, R. P. and Brown, D. H. (1984 a). The control of cadmium uptake in the lichen genus Peltigera. Journal of Experimental Botany, 35, 1071–1082.CrossRefGoogle Scholar
Beckett, R. P. and Brown, D. H. (1984 b). The relationship between cadmium uptake and heavy metal uptake tolerance in the lichen genus Peltigera. New Phytologist, 97, 301–311.CrossRefGoogle Scholar
Beckett, R. P. and Minibayeva, F. V. (2007). Rapid breakdown of exogenous extracellular hydrogen peroxide by lichens. Physiologia Plantarum, 129, 588–596.CrossRefGoogle Scholar
Beckett, R. P., Marschall, M. and Laufer, Z. (2005 a). Hardening enhances photoprotection in the moss Atrichum androgynum during rehydration by increasing fast rather than slow-relaxing quenching. Journal of Bryology, 27, 7–12.CrossRefGoogle Scholar
Beckett, R. P., Mayaba, N., Minibayeva, F. V. and Alyabyev, A. J. (2005 b). Hardening by partial dehydration and ABA increase desiccation tolerance in the cyanobacterial lichen Peltigera polydactylon. Annals of Botany, 96, 109–115.CrossRefGoogle ScholarPubMed
Beckett, R. P., Minibayeva, F. V., Vylegzhanina, N. V. and Tolpysheva, T. (2003). High rates of extracellular superoxide reduction by lichens in the Suborder Peltigerineae correlate with indices of high metabolic activity. Plant, Cell and Environment, 26, 1827–1837.CrossRefGoogle Scholar
Bedeneau, M. (1982). Reproduction in vitro des effets de la pollution par le dioxyde de soufre sur quelques lichens. Annales des Sciences Forestieres, 39, 165–178.CrossRefGoogle Scholar
Bedford, D. J., Schweizer, E., Hopwood, D. A. and Khosla, C. (1995). Expression of a functional fungal polyketide synthase in the bacterium Streptomyces coelicolor A3(2). Journal of Bacteriology, 177, 4544–4548.CrossRefGoogle Scholar
Begora, M. and Fahselt, D. (2001). Photolability of secondary compounds in some lichen species. Symbiosis, 31, 3–22.Google Scholar
Belandria, G., Asta, J. and Nurit, F. (1989). Effects of sulphur dioxide and fluoride on ascospore germination of several lichens. Lichenologist, 21, 79–86.CrossRefGoogle Scholar
Belnap, J. (2001). Factors influencing nitrogen fixation and nitrogen release in biological soil crusts. In Biological Soil Crusts: Structure, Function, and Management, ed. Belnap, J. and Lange, O. L., pp. 241–261. Berlin: Springer.Google Scholar
Belnap, J. (2002). Nitrogen fixation in biological soil crusts from southeast Utah, USA. Biological Fertility of Soils, 35, 128–135.CrossRefGoogle Scholar
Belnap, J. and Lange, O. L. (eds.) (2003). Biological Soil Crusts: Structure, Function, and Management: Ecological Studies 150. Berlin: Springer.CrossRefGoogle Scholar
Belnap, J. and Lange, O. L. (2005 a). Lichens and microfungi in biological soil crusts: community structure, physiology, and ecological functions. In The Fungal Community: Its Organization and Role in the Ecosystem, Vol. 3, ed. Dighton, J., White, J. F. and Oudemans, P., pp. 117–138. Boca Raton: CRC Press.CrossRefGoogle Scholar
Belnap, J. and Lange, O. L. (2005 b). Biological soil crusts and global changes: what does the future hold? In The Fungal Community: Its Organization and Role in the Ecosystem, Vol. 3, ed. Dighton, J., White, J. F. and Oudemans, P., pp. 697–712. Boca Raton: CRC Press.CrossRefGoogle Scholar
Benedict, J. B. (1991). Experiments on lichen growth. II. Effects of a seasonal snow cover. Arctic and Alpine Research, 23, 189–199.CrossRefGoogle Scholar
Benner, J. W. and Vitousek, P. M. (2007). Development of a diverse epiphyte community in response to phosphorous fertilization. Ecological Letters, 10, 628–636.CrossRefGoogle Scholar
Benner, J. W., Conroy, S., Lunch, C. K.Toyoda, N. and Vitousek, P. M. (2007). Phosphorus fertilization increases the abundance and nitrogenase activity of the cyanolichen Pseudocyphellaria crocata in Hawaiian montane forests. Biotropica, 39, 400–405.CrossRefGoogle Scholar
Bennett, J. P. and Wetmore, C. M. (1999). Geothermal elements in lichens of Yellowstone National Park, USA. Environmental and Experimental Botany, 42, 191–200.CrossRefGoogle Scholar
Berger, F. and Aptroot, A. (2003). Further contributions to the flora of lichens and lichenicolous fungi of the Azores. Arquipélago, 19A, 1–12.Google Scholar
Bergman, D. E. and Ebinger, J. E. (1990). Cyanogenesis in the lichen genus Dermatocarpon. Castanea, 55, 207–210.Google Scholar
Beschel, R. E. (1961). Dating rock surfaces by lichen growth and its application to glaciology and physiography (lichenometry). In Geology of the Arctic, Vol. 2, ed. Raasch, G. O., pp. 1044–1062. Toronto: University of Toronto Press.Google Scholar
Bewley, J. D. (1979). Physiological aspects of desiccation tolerance. Annual Review of Plant Physiology, 30, 195–238.CrossRefGoogle Scholar
Bewley, J. D. and Krochko, J. E. (1982). Desiccation tolerance. In Physiological Plant Ecology. Vol. II: Water Relations and Carbon Assimilation, ed. Lange, O. L., Nobel, P. S., Osmond, C. B. and Ziegler, H., pp. 325–378. Encyclopedia of Plant Physiology 12B. Berlin: Springer.CrossRefGoogle Scholar
Bilger, W., Rimke, S., Schreiber, U. and Lange, O. L. (1989). Inhibition of energy-transfer to photosystem II in lichens by dehydration: different properties of reversibility with green and blue-green phycobionts. Journal of Plant Physiology, 134, 261–268.CrossRefGoogle Scholar
Bingle, L. E., Simpson, T. J. and Lazarus, C. M. (1999). Ketosynthase domain probes identify two subclasses of fungal polyketide synthase genes. Fungal Genetics and Biology, 26, 209–223.CrossRefGoogle ScholarPubMed
Bischoff, H. W. and Bold, H. C. (1963). Phycological studies. IV. Some soil algae from Enchanted Rock and related algal species. University of Texas Publication, 6318, 1–95.Google Scholar
Bjelland, T. and Ekman, S. (2005). Fungal diversity in rock beneath a crustose lichen as revealed by molecular markers. Microbial Ecology, 49, 598–603.CrossRefGoogle ScholarPubMed
Bjerke, J. W. (2005). Synopsis of the lichen genus Menegazzia (Parmeliaceae, lichenized Ascomycotina) in South America. Mycotaxon, 91, 423–454.Google Scholar
Bjerke, J. W., Elvebakk, A., Dominguez, B. and Dahlbäck, A. (2005). Seasonal trends in usnic acid concentrations of Arctic, alpine and Patagonian populations of the lichen Flavocetraria nivalis. Phytochemistry, 66, 337–344.CrossRefGoogle ScholarPubMed
Bjerke, J. W., Lerfall, K. and Elvebakk, A. (2002). Effects of ultraviolet radiation and PAR on the content of usnic and divaricatic acids in two arctic-alpine lichens. Photochemical and Photobiological Sciences, 1, 678–685.CrossRefGoogle ScholarPubMed
Bjerke, J. W., Zielke, M. and Solheim, B. (2003). Long-term impacts of simulated climatic change on secondary metabolism, thallus structure and nitrogen fixation activity in two cyanolichens from the Arctic. New Phytologist, 159, 361–367.CrossRefGoogle Scholar
Björkman, O. (1981). Responses to different quantum flux densities. In Physiological Plant Ecology. Vol. I: Responses to the Physical Environment, ed. Lange, O. L., Nobel, P. S., Osmond, C. B. and Ziegler, H., pp. 57–108. Encyclopedia of Plant Physiology 12A. Berlin: Springer.CrossRefGoogle Scholar
Björkman, O., Boardman, N. K., Anderson, J. A., et al. (1972). Effect of light intensity during growth of Atriplex patula on the capacity of photosynthetic reactions, chloroplast components and structure. Carnegie Institution Washington Yearbook, 71, 115–135.Google Scholar
Black, M. and Pritchard, H. W. (2002). Desiccation and Survival in Plants: Drying Without Dying. Oxon: CABI Publishing.CrossRefGoogle Scholar
Blaha, J., Baloch, E. and Grube, M. (2006). High photobiont diversity associated with the euryoecious lichen-forming ascomycete Lecanora rupicola (Lecanoraceae, Ascomycota). Biological Journal of the Linnean Society, 88, 283–293.CrossRefGoogle Scholar
Blanco, O., Crespo, A., Divakar, P. K., Elix, J. A. and Lumbsch, H. T. (2005). Molecular phylogeny of parmotremoid lichens (Ascomycota, Parmeliaceae). Mycologia, 97, 150–159.CrossRefGoogle Scholar
Blanco, O., Crespo, A., Elix, J. A., Hawksworth, D. L. and Lumbsch, H. T. (2004). A molecular phylogeny and a new classification of parmelioid lichens containing Xanthoparmelia-type lichenan (Ascomycota: Lecanorales). Taxon, 53, 959–975.CrossRefGoogle Scholar
Blanco, O., Crespo, A., Ree, R. H. and Lumbsch, H. T. (2006). Major clades of parmelioid lichens (Parmeliaceae, Ascomycota) and the evolution of their morphological and chemical diversity. Molecular Phylogenetics and Evolution, 39, 52–69.CrossRefGoogle ScholarPubMed
Blum, O. B. (1973). Water relations. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 381–400. New York: Academic Press.Google Scholar
Boardman, N. K. (1977). Comparative photosynthesis of sun and shade plants. Annual Review of Plant Physiology, 28, 355–377.CrossRefGoogle Scholar
Boardman, N. K., Anderson, J. M., Thorne, S. E. and Björkman, O. (1972). Photochemical reactions of chloroplasts and components of the photosynthetic electron transport chain in two rainforest species. Carnegie Institution Washington Yearbook, 71, 107–114.Google Scholar
Boileau, L. J. R., Beckett, P. J., Lavoie, P., Richardson, D. H. S. and Nieboer, E. (1982). Lichens and mosses as monitors of industrial activity associated with uranium mining in northern Ontario, Canada. I. Field procedures, chemical analyses and interspecies comparisons. Experimental Pollution (Series B), 4, 69–84.CrossRefGoogle Scholar
Boison, G., Mergel, A., Jolkver, H. and Bothe, H. (2004). Bacterial life and dinitrogen fixation at a gypsum rock. Applied and Environmental Microbiology, 70, 7070–7077.CrossRefGoogle Scholar
Boissière, M.-C. (1982). Cytochemical ultrastructure of Peltigera canina: some factors relating to its symbiosis. Lichenologist, 14, 1–28.CrossRefGoogle Scholar
Boissière, M.-C. (1987). Ultrastructural relationship between the composition and the structure of the cell wall of the mycobiont of two lichens. Bibliotheca Lichenologica, 25, 117–123.Google Scholar
Bold, H. C. and Wynne, M. J. (1985). Introduction to the Algae. Stucture and Reproduction. 2nd edn. Englewood Cliffs: Prentice Hall.Google Scholar
Boonpragob, K., Nash, T. H., III, and Fox, C. A. (1989). Seasonal deposition patterns of acidic ions and ammonium to the lichen Ramalina menziesii Tayl. in southern California. Environmental and Experimental Botany, 29, 187–197.CrossRefGoogle Scholar
Borecký, J. and Vercesi, A. (2005). Plant uncoupling mitochondrial protein and alternative oxidase: energy metabolism and stress. Bioscience Reports, 25, 271–286.CrossRefGoogle Scholar
Boreham, S. (1992). A study of corticolous lichens on London plane Platanus × hybrida trees in West Ham Park, London. London Naturalist, 71, 61–71.Google Scholar
Bothe, H. and Loos, E. (1972). Effect of far red light and inhibitors on nitrogen fixation and photosynthesis in the blue-green alga Anabaena cylindrica. Archiv für Microbiologie, 86, 241–254.CrossRefGoogle Scholar
Bothe, H., Distler, E. and Eisbrenner, G. (1978). Hydrogen metabolism in blue-green algae. Biochimie, 60, 277–289.CrossRefGoogle ScholarPubMed
Bottomley, P. J. and Stewart, W. D. P. (1977). ATP and nitrogenase activity in nitrogen-fixing heterocystous blue-green algae. New Phytologist, 79, 625–638.CrossRefGoogle Scholar
Boucher, V. L. and Nash, T. H., III (1990 a). Growth pattern in Ramalina menziesii in California: coastal vs. inland populations. Bryologist, 93, 295–302.CrossRefGoogle Scholar
Boucher, V. L. and Nash, T. H., III (1990 b). The role of the fruticose lichen Ramalina menziesii in the annual turnover of biomass and macronutrients in a blue oak woodland. Botanical Gazette, 151, 114–118.CrossRefGoogle Scholar
Boustie, J. and Grube, M. (2005). Lichens – a promising source of bioactive secondary metabolites. Plant Genetic Resources, 3, 273–287.CrossRefGoogle Scholar
Bowker, M. A., Belnap, J., Davidson, D. W. and Phillips, S. L. (2005). Evidence for micronutrient limitation of biological soil crusts: importance to arid-lands restoration. Ecological Applications, 15, 1941–1951.CrossRefGoogle Scholar
Branquinho, C., Brown, D. H. and Catarino, F. (1997). The cellular location of Cu in lichens and its effects on membrane integrity and chlorophyll fluorescence. Environmental and Experimental Botany, 38, 165–179.CrossRefGoogle Scholar
Brightman, F. H. and Seaward, M. R. D. (1977). Lichens of man-made substrates. In Lichen Ecology, ed. Seaward, M. R. D., pp. 253–293. London: Academic Press.Google Scholar
Broady, P. A. and Ingerfeld, M. (1993). Three new species and a new record of chaetophoracean (Chlorophyta) algae from terrestrial habitats in Antarctica. European Journal of Phycology, 28, 25–31.CrossRefGoogle Scholar
Brochmann, C., Gabrielsen, T. M., Nordal, I., Landvik, J. Y. and Elven, R. (2003). Glacial survival or tabula rasa? The history of North Atlantic biota revisited. Taxon, 52, 417–450.CrossRefGoogle Scholar
Brock, T. D. (1978). Thermophilic Microorganisms and Life at High Temperatures. New York: Springer.CrossRefGoogle Scholar
Brodo, I. M. (1973). Substrate ecology. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 401–441. New York: Academic Press.Google Scholar
Brodo, I. M. (1978). Changing concepts regarding chemical diversity in lichens. Lichenologist, 10, 1–11.CrossRefGoogle Scholar
Brodo, I. M. and Richardson, D. H. S. (1978). Chimeroid associations in the genus Peltigera. Lichenologist, 10, 157–170.CrossRefGoogle Scholar
Brodo, I. M., Sharnoff, S. D. and Sharnoff, S. (2001). Lichens of North America. New Haven: Yale University Press.Google Scholar
Brooks, D. R. (2004). Reticulations in historical biogeography: the triumph of time over space in evolution. In Frontiers of Biogeography: New Directions in the Geography of Nature, ed. Lomolino, M. V. and Heaney, L. R., pp. 125–144. Sunderland: Sinauer Associates.Google Scholar
Brouwer, R. (1962). Distribution of dry matter in the plant. Netherland Journal of Agricultural Sciences, 10, 399–408.Google Scholar
Brown, D. H. (1972). The effect of Kuwait crude oil and a solvent emulsifier on the metabolism of the marine lichen, Lichina pygmaea. Marine Biology, 12, 309–315.CrossRefGoogle Scholar
Brown, D. H. (1992). Impact of agriculture on bryophytes and lichens. In Bryophytes and Lichens in a Changing Environment, ed. Bates, J. W. and Farmer, A. M., pp. 259–283. Oxford: Clarendon Press.Google Scholar
Brown, D. H. and Beckett, R. P. (1984). Uptake and effect of cations on lichen metabolism. Lichenologist, 16, 173–188.CrossRefGoogle Scholar
Brown, D. H. and Beckett, R. P. (1985). The role of the cell wall in the intracellular uptake of cations by lichens. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 247–258. New York: Plenum Press.CrossRefGoogle Scholar
Brown, D. H., Ascaso, C. and Rapsch, S. (1987). Ultrastructural changes in the pyrenoid of the lichen Parmelia sulcata stored under controlled conditions. Protoplasma, 136, 136–144.CrossRefGoogle Scholar
Brown, D. H., MacFarlane, J. D. and Kershaw, K. A. (1983). Physiological-environmental interactions in lichens. XVI. A re-examination of resaturation respiration phenomena. New Phytologist, 93, 237–246.CrossRefGoogle Scholar
Brown, D. H., Standell, C. J. and Miller, J. E. (1995). Effects of agricultural chemicals on lichens. Cryptogamic Botany, 5, 220–223.Google Scholar
Brown, M. J., Jarman, S. K. and Kantvilas, G. (1994). Conservation and reservation of non-vascular plants in Tasmania, with special reference to lichens. Biodiversity and Conservation, 3, 263–278.CrossRefGoogle Scholar
Brown, R. M. Jr. and Bold, H. C. (1964). Comparative studies of the algal genera Tetracystis and Chlorococcum. Phycological Studies V. University of Texas Publications, 6417, 1–213.Google Scholar
Brunauer, G. and Stocker-Wörgötter, E. (2005). Culture of lichen fungi for future production of biologically active compounds. Symbiosis, 38, 187–201.Google Scholar
Brunauer, G., Grube, M., Muggia, L. and Stocker-Wörgötter, E. (2006). Gene bank of PKS from the mycobiont of Xanthoria elegans. Published in NCBI Database.
Brunner, U. and Honegger, R. (1985). Chemical and ultrastructural studies on the distribution of sporopollenin-like biopolymers in 6 genera of lichen phycobionts. Canadian Journal of Botany, 63, 2221–2230.CrossRefGoogle Scholar
Bruns, T. D., White, T. J. and Taylor, J. W. (1991). Fungal molecular systematics. Annual Review of Ecology and Systematics, 22, 525–564.CrossRefGoogle Scholar
Bruteig, I. E. (1993). The epiphytic lichen Hypogymnia physodes as biomonitor of atmospheric nitrogen and sulphur deposition in Norway. Environmental Monitoring and Assessment, 26, 27–47.CrossRefGoogle ScholarPubMed
Bryant, J. P., Chapin, F. S. III and Klein, D. R. (1983). Carbon nutrient balance of boreal plants in relation to vertebrate herbivory. Oikos, 40, 357–368.CrossRefGoogle Scholar
Bubrick, P. and Galun, M. (1980 a). Proteins from the lichen Xanthoria parietina which bind to phycobiont cell walls. Correlation between binding patterns and cell wall cytochemistry. Protoplasma, 104, 167–173.CrossRefGoogle Scholar
Bubrick, P. and Galun, M. (1980 b). Symbiosis in lichens: differences in cell wall properties of freshly isolated and cultured phycobionts. FEMS Microbiology Letters, 7, 311–313.CrossRefGoogle Scholar
Bubrick, P., Galun, M. and Frensdorff, A. (1981). Proteins from the lichen Xanthoria parietina which bind to phycobiont cell walls. Localization in the intact lichen and cultured mycobiont. Protoplasma, 105, 207–211.CrossRefGoogle Scholar
Bubrick, P., Galun, M. and Frensdorff, A. (1984). Observations on free-living Trebouxia de Puymaly and Pseudotrebouxia Archibald, and evidence that both symbionts from Xanthoria parietina (L.) Th. Fr. can be found free-living in nature. New Phytologist, 97, 455–462.CrossRefGoogle Scholar
Buchauer, M. J. (1973). Contamination of soil and vegetation near a zinc smelter by zinc, cadmium, copper, and lead. Environmental Science and Technology, 7, 131–135.CrossRefGoogle Scholar
Büdel, B. (1982). Phycobionten der Lichinaceen. Diplom-Thesis. Marburg: Universität Marburg.
Büdel, B. (1987). Zur Biologie und Systematik der Flechtengattungen Heppia und Peltula im südlichen Afrika. Bibliotheca Lichenologica, 23, 1–105.Google Scholar
Büdel, B. (1990). Anatomical adaptations to the semiarid/arid environment in the lichen genus Peltula. Bibliotheca Lichenologica, 38, 47–61.Google Scholar
Büdel, B. (1992). Taxonomy of lichenized procaryotic blue-green algae. In Algae and Symbioses, ed. Reisser, W., pp. 301–324. Bristol: Biopress Limited.Google Scholar
Büdel, B. and Henssen, A. (1983). Chroococcidiopsis (Cyanophyceae), a phycobiont in the lichen family Lichinaceae. Phycologia, 22, 367–375.CrossRefGoogle Scholar
Büdel, B. and Henssen, A. (1988). Zwei neue Peltula-Arten von Südafrika. International Journal of Mycology and Lichenology, 2, 235–249.Google Scholar
Büdel, B. and Lange, O. L. (1991). Water status of green and blue-green phycobionts in lichen thalli after hydration by water vapor uptake: do they become turgid?Botanica Acta, 104, 361–366.CrossRefGoogle Scholar
Büdel, B. and Lange, O. L. (1994). The role of cortical and epicortical layers in the lichen genus Peltula. Cryptogamic Botany, 4, 262–269.Google Scholar
Büdel, B. and Wessels, D. C. (1986). Parmelia hueana Gyeln., a vagrant lichen from the Namib Desert, SWA/Namibia. I. Anatomical and reproductive adaptations. Dinteria, 18, 3–15.Google Scholar
Bull, W. B. (1996). Dating San Andreas fault earthquakes with lichenometry. Geology, 24, 111–114.2.3.CO;2>CrossRefGoogle Scholar
Bull, W. B. and Brandon, M. T. (1998). Lichen dating of earthquake-generated regional rockfall events, Southern Alps, New Zealand. Geological Society of America Bulletin, 110, 60–84.2.3.CO;2>CrossRefGoogle Scholar
Bull, W. B., King, J., Kong, F., Moutoux, T. and Phillips, W. M. (1994). Lichen dating of coseismic landslide hazards in alpine mountains. Geomorphology, 10, 253–264.CrossRefGoogle Scholar
Bunce, H. W. F. (1996). Methods of monitoring smelter emission effects on a temperate rain forest. Fluoride, 29, 241–251.Google Scholar
Bungartz, F., Garvie, L. A. J. and Nash, T. H., III (2004). Anatomy of the endolithic Sonoran Desert lichen Verrucaria rubrocincta Breuss: implications for biodeterioration and biomineralization. Lichenologist, 36, 55–73.CrossRefGoogle Scholar
Burkholder, P. R. and Evans, A. W. (1945). Further studies on the antibiotic activity of Lichens. Bulletin of the Torrey Botanical Club, 72, 157–164.CrossRefGoogle Scholar
Burkholder, P. R., Evans, A. W., McVeigh, I. and Thornton, H. K. (1944). Antibiotic activity of Lichens. Proceedings of the National Academy of Sciences, USA, 30, 250–255.CrossRefGoogle ScholarPubMed
Buschbom, J. and Barker, D. (2006). Evolutionary history of vegetative reproduction in Porpidia s.l. (lichen-forming Ascomycota). Systematic Biology, 55, 417–484.CrossRefGoogle Scholar
Buschbom, J. and Mueller, G. M. (2004). Resolving evolutionary relationships in the lichen-forming genus Porpidia and related allies (Porpidiaceae, Ascomycota). Molecular Phylogenetics and Evolution, 32, 66–82.CrossRefGoogle Scholar
Buschbom, J. and Mueller, G. M. (2005). Testing “species pair” hypotheses: evolutionary processes in the lichen-forming species complex Porpidia flavocoerulescens and Porpidia melinodes. Molecular Biology and Evolution, 23, 574–586.CrossRefGoogle ScholarPubMed
Butin, H. (1954). Physiologisch-ökologische Untersuchungen über den Wasserhaushalt und die Photosynthese bei Flechten. Biologisches Zentralblatt, 73, 459–502.Google Scholar
Butler, M. J. and Day, A. W. (1998). Fungal melanins: a review. Canadian Journal of Microbiology, 44, 1115–1136.CrossRefGoogle Scholar
Bychek-Guschina, I. A., Kotlova, E. R. and Heipieper, H. (1999). Effects of sulfur dioxide on lichen lipids and fatty acids. Biochemistry (Moscow), 64, 61–65.Google ScholarPubMed
Calatayud, A., Sanz, M.-J., Calvo, E., Barreno, E. and Valle-Tascon, S. (1996). Chlorophyll a fluorescence and chlorophyll content in Parmelia quercina thalli from a polluted region of northern Castellón (Spain). Lichenologist, 28, 49–65.CrossRefGoogle Scholar
Calatayud, A., Tempe, P. J. and Barreno, E. (2000). Chlorophyll a fluorescence emission, xanthophylls cycle activity, and net photosynthetic responses to ozone in some foliose and fruticose lichen species. Photosynthetica, 38, 281–286.CrossRefGoogle Scholar
Caldwell, C. F., Turano, F. J. and McMahon, M. B. (1998). Identification of two cytosolic ascorbate peroxidase cDNAs from soybean leaves and characterization of their products by functional expression in E. coli. Planta, 204, 120–126.Google ScholarPubMed
Calvelo, S. and Liberatore, S. (2001). Checklist of Argentinian lichens (version 2). Online: www.biologie.uni-hamburg.de/checklists.argen_12.htm.
Campbell, D. (1996). Complementary chromatic adaptation alters photosynthetic strategies in the cyanobacterium Calothrix. Microbiology, 142, 1255–1263.CrossRefGoogle Scholar
Campbell, D., Hurry, V., Clarke, A. K., Gustafsson, P. and Öquist, G. (1998). Chlorophyll fluorescence analysis of cyanobacterial photosynthesis and acclimation. Microbiology and Molecular Biology Reviews, 62, 667–683.Google ScholarPubMed
Cane, D. E., Walsh, C. T. and Khosla, C. (1998). Harnessing the biosynthetic code: combinations, permutations, and mutations. Science, 282, 63–68.CrossRefGoogle ScholarPubMed
Cardinale, M., Puglia, A. M. and Grube, M. (2006). Molecular analysis of lichen-associated bacterial communities. FEMS Microbiology Ecology, 57, 484–495.CrossRefGoogle ScholarPubMed
Carlberg, G. E., Ofstad, E. B., Drangsholt, H. and Steinnes, E. (1983). Atmospheric deposition of organic micropollutants in Norway studied by means of moss and lichen analysis. Chemosphere, 12, 341–356.CrossRefGoogle Scholar
Carter, N. E. A. and Viles, H. A. (2003). Experimental investigations into the interactions between moisture, rock surface temperatures and an epilithic lichen cover in the bioprotection of limestone. Building and Environment, 38, 1225–1234.CrossRefGoogle Scholar
Carter, N. E. A. and Viles, H. A. (2005). Bioprotection explored: the story of a little known earth surface process. Geomorphology, 67, 273–281.CrossRefGoogle Scholar
Case, J. W. and Krouse, H. R. (1980). Variations in sulphur content and stable sulphur isotope composition of vegetation near a SO2 source at Fox Creek, Alberta, Canada. Oecologia, 44, 248–257.CrossRefGoogle Scholar
Casely, A. F. and Dugmore, A. J. (2004). Climate change and ‘anomalous’ glacier fluctuations: the southwest outlets of Myrdalsjokull, Iceland. Boreas, 33, 108–122.CrossRefGoogle Scholar
Casselman, K. D. (2001). Lichen Dyes: The New Source Book. Mineola, NY: Dover Publications.Google Scholar
Castenholz, R. W. and Waterbury, J. B. (1989). Group I. Cyanobacteria. In Bergey's Manual of Systematic Bacteriology, Vol. 3, ed. Staley, J. T., Bryant, P., Pfennig, N and Holt, J. G., pp. 1710–1806. Baltimore: Williams and Wilkins.Google Scholar
Cavalier-Smith, T. (1987). The origin of fungi and pseudofungi. In Evolutionary Biology of the Fungi, ed. Rayner, A. D. M., Brasier, C. N. and Moore, D., pp. 339–353. Cambridge: Cambridge University Press.Google Scholar
Chamberlain, A. C. (1970). Interception and retention of radioactive aerosols by vegetation. Atmospheric Environment, 4, 57–78.CrossRefGoogle ScholarPubMed
Chapin, F. S. III (1991). Integrated responses of plants to stress. BioScience, 41, 29–36.CrossRefGoogle Scholar
Chapin, F. S. III, Bloom, A. J., Field, C. B. and Waring, R. H. (1987). Plant responses to multiple environmental factors. BioScience, 37, 49–57.CrossRefGoogle Scholar
Chapman, R. L. (1984). An assessment of the current state of our knowledge of the Trentepohliaceae. In Systematics of the Green Algae. ed. Irvine, D. E. G. and John, D. M., pp. 233–250. London: Academic Press.Google Scholar
Chen, G.-X., Kazmir, J. and Cheniae, G. M. (1992). Photoinhibition of hydroxylamine-extracted photosystem II membranes: studies of the mechanism. Biochemistry, 31, 11 072–11 083.CrossRefGoogle ScholarPubMed
Chen, S., Wu, D. and Wu, J. (1989). Using lichen communities as SO2 pollution monitors. Journal of Nanjing Normal University (Natural Science), 12, 77–82.Google Scholar
Chooi, Y. H., Stalker, D., Louwhoff, S. and Lawrie, A. (2006). The search for a polyketide synthase gene producing beta-orsellinic acid and methylphloroacetophenone as precursors of depsidones and usnic acid in the lichen Chondropsis semiviridis. Poster presentation, International Mycological Congress, IMC 8, Cairns, Australia.
Cislaghi, C. and Nimis, P. L. (1997). Lichens, air pollution and lung cancer. Nature, 387, 463–464.CrossRefGoogle ScholarPubMed
Clarke, A. K., Campbell, D., Gustafsson, P. and Öquist, G. (1995). Dynamic responses of photosystem II and phycobilisomes to changing light in the cyanobacterium Synechococcus sp. PCC 7942. Planta, 197, 553–562.CrossRefGoogle Scholar
Clauzade, G. and Roux, C. (1976). Les Champignons Lichénicoles non Lichénisés. Montpellier: Université des Sciences et Techniques du Languedoc.Google Scholar
Clayden, S. R. (1992). Chemical divergence of eastern North American and European populations of Arctoparmelia centrifuga and their sympatric usnic acid-deficient chemotypes. Bryologist, 95, 1–4.CrossRefGoogle Scholar
Clayden, S. R. (1997). Intraspecific interactions and parasitism in an association of Rhizocarpon lecanorinum and R. geographicum. Lichenologist, 29, 533–545.CrossRefGoogle Scholar
Clerc, P. (2006). Synopsis of Usnea (lichenized Ascomycetes) from the Azores with additional information on the species in Macaronesia. Lichenologist 38, 191–212.CrossRefGoogle Scholar
Codogno, M., Poelt, J. and Puntillo, D. (1989). Umbilicaria freyi spec. nova und der Formenkreis von Umbilicaria hirsuta in Europa. Plant Systematics and Evolution, 165, 55–69.CrossRefGoogle Scholar
Cohn, F. (1853). Untersuchungen über die Entwicklungsgeschichte microskopischer Algen und Pilze. Novorum actorum academiae caesareae leopoldinae-carolinae naturae curiosorum, 24, 101–256.Google Scholar
Coley, P. D. (1988). Effects of plant growth rate and leaf lifetime on the amount and type of anti-herbivore defense. Oecologia, 74, 531–536.CrossRefGoogle ScholarPubMed
Collins, C. R. and Farrar, J. F. (1978). Structural resistances to mass transfer in the lichen Xanthoria parietina. New Phytologist, 31, 71–78.CrossRefGoogle Scholar
Common, R. S. (1991). The distribution and taxonomic significance of lichenan and isolichenan in the Parmeliaceae (lichenized Ascomycotina), as determined by iodine reactions. I. Introduction and methods. II. The genus Alectoria and associated taxa. Mycotaxon, 41, 67–112.Google Scholar
Cook, L. G. and Crisp, M. D. (2005). Directional asymmetry of long-distance dispersal and colonization could mislead reconstructions of biogeography. Journal of Biogeography, 32, 741–754.CrossRefGoogle Scholar
Cook, L. M. (2003). The rise and fall of the carbonaria form of the peppered moth. Quarterly Review of Biology, 78, 399–417.CrossRefGoogle ScholarPubMed
Cowan, D. A., Green, T. G. A. and Wilson, A. T. (1979 a). Lichen metabolism. 1. The use of tritium labelled water in studies of anhydrobiotic metabolism in Ramalina celastri and Peltigera polydactyla. New Phytologist, 82, 489–503.CrossRefGoogle Scholar
Cowan, D. A., Green, T. G. A. and Wilson, A. T. (1979 b). Lichen metabolism. 2. Aspects of light and dark physiology. New Phytologist, 83, 761–769.CrossRefGoogle Scholar
Cowan, I. R., Lange, O. L. and Green, T. G. A. (1992). Carbon-dioxide exchange in lichens: determination of transport and carboxylation characteristics. Planta, 187, 282–294.CrossRefGoogle ScholarPubMed
Cowie, R. H. and Holland, B. S. (2006). Dispersal is fundamental to biogeography and the evolution of biodiversity on oceanic islands. Journal of Biogeography, 33, 193–198.CrossRefGoogle Scholar
Cox, C. B. and Moore, P. D. (2005). Biogeography: An Ecological and Evolutionary Approach. Oxford: Blackwell Publishing.Google Scholar
Cox, R. J., Hitchman, T. S., Byron, K. J., et al. (1997). Post-translational modification of heterologously expressed Streptomyces Type II polyketide synthase acyl carrier proteins. FEBS Letters, 405, 267–272.CrossRefGoogle ScholarPubMed
Coxson, D. S. (1988). Recovery of net photosynthesis and dark respiration on rehydration of the lichen, Cladina mitis, and the influence of prior exposure to sulphur dioxide while desiccated. New Phytologist, 108, 483–487.CrossRefGoogle Scholar
Coxson, D. S. (1991). Impedance measurement of thallus moisture content in lichens. Lichenologist, 23, 77–84.CrossRefGoogle Scholar
Coxson, D. S. and Curteanu, M. (2002). Decomposition of hair lichens (Alectoria sarmentosa and Bryoria spp.) under snowpack in montane forest, Cariboo Mountains, British Columbia. Lichenologist, 34, 395–402.CrossRefGoogle Scholar
Coxson, D. S. and Nadkarni, N. M. (1995). Ecological roles of epiphytes in nutrient cycles of forest ecosystems. In Forest Canopies, ed. Lowman, M. D. and Nadkarni, N. M., pp. 495–543. London: Academic Press.Google Scholar
Coxson, D. S., Webber, M. R. and Kershaw, K. A. (1984). The thermal operating environment of corticolous and pendulous tree lichens. Bryologist, 87, 197–202.CrossRefGoogle Scholar
Craw, R. C., Grehan, J. R. and Heads, M. J. (1999). Panbiogeography: Tracking the History of Life. Oxford Biogeography Series. Oxford: Oxford University Press, 11, 1–238.Google Scholar
Crespo, A., Bridge, P. D., Hawksworth, D. L., Grube, M and Cubero, O. F. (1999). Comparison of rRNA genotypic variability in the lichen-forming fungus Parmelia sulcata from long established and recolonizing sites following sulfur dioxide amelioration. Plant Systematics and Evolution, 217, 177–183.CrossRefGoogle Scholar
Crespo, A., Lumbsch, H. T., Matsson, J.-E., et al. (2007). Testing morphology-based hypotheses of phylogenetic relationships in Parmeliaceae (Ascomycota) using three ribosomal markers and the nuclear RPB-1 gene. Molecular Phylogenetics and Evolution, 44, 812–824.CrossRefGoogle Scholar
Crews, T. E., Kurina, L. M. and Vitousek, P. M. (2001). Organic matter and nitrogen accumulation and nitrogen fixation during early ecosystem development in Hawaii. Biogeochemistry, 52, 259–279.CrossRefGoogle Scholar
Crisci, J. V., Katinas, L. and Posadas, P. (2003). Historical Biogeography: An Introduction. Cambridge: Harvard University Press.Google Scholar
Crisci, J. V., Sala, O. E., Katinas, L. and Posadas, P. (2006). Bridging historical and ecological approaches to biogeography. Australian Systematic Botany, 19, 1–10.CrossRefGoogle Scholar
Crittenden, P. D. (1975). Nitrogen fixation on the glacial drift of Iceland. New Phytologist, 74, 41–49.CrossRefGoogle Scholar
Crittenden, P. D. (1983). The role of lichens in the nitrogen economy of subarctic woodlands: nitrogen loss from the nitrogen-fixing lichen Stereocaulon paschale during rainfall. In Nitrogen as an Ecological Factor, ed. Boddy, L., Marchant, R. and Read, D. J., pp. 43–68. Oxford: Blackwell Scientific Publications.Google Scholar
Crittenden, P. D. (1989). Nitrogen relations of mat-forming lichens. In Nitrogen, Phosphorus and Sulphur Utilization by Fungi, ed. Boddy, L., Marchant, R. and Read, D. J., pp. 243–268. Cambridge: Cambridge University Press.Google Scholar
Crittenden, P. D. (1996). The effect of oxygen deprivation on inorganic nitrogen uptake in an Antarctic macrolichen. Lichenologist, 28, 347–354.CrossRefGoogle Scholar
Crittenden, P. D. (1998). Nutrient exchange in an Antarctic macrolichen during summer snowfall snow melt events. New Phytologist, 139, 697–707.CrossRefGoogle Scholar
Crittenden, P. D. and Kershaw, K. A. (1978). A procedure for the simultaneous measurement of net CO2-exchange and nitrogenase activity in lichens. New Phytologist, 80, 393–401.CrossRefGoogle Scholar
Crittenden, P. D. and Kershaw, K. A. (1979). Studies on lichen-dominated systems. XXII. The environmental control of nitrogenase activity in Stereocaulon paschale in spruce-lichen woodland. Canadian Journal of Botany, 53, 236–254.CrossRefGoogle Scholar
Crittenden, P. D., Kalucka, I. and Oliver, E. (1994). Does nitrogen supply limit the growth of lichens?Cryptogamic Botany, 4, 143–155.Google Scholar
Crittenden, P. D., Llimona, X. and Sancho, L. (2004). Nitrogenase activity in Thyrea spp. – preliminary results. In Book of Abstracts of the 5th IAL Symposium, Lichens in Focus, ed. Randlane, T. and Saag, A., p. 44. Tartu: Tartu University Press.Google Scholar
Crowe, J. H., Crowe, L. M. and Chapman, D. (1984). Preservation of membranes in anhydrobiotic organisms: the role of trehalose. Science, 223, 701–703.CrossRefGoogle Scholar
Culberson, C. F. (1972). Improved conditions and new data for the identification of lichen products by a standardized thin-layer chromatographic method. Journal of Chromatography, 72, 113–125.CrossRefGoogle ScholarPubMed
Culberson, C. F. (1986). Biogenetic relationships of the lichen substances in the framework of systematics. Bryologist, 89, 91–98.CrossRefGoogle Scholar
Culberson, C. F. and Ammann, K. (1979). Standardmethode zur Dünnschichtchromatographie von Flechtensubstanzen. Herzogia, 5, 1–24.Google Scholar
Culberson, C. F. and Culberson, W. L. (1976). Chemosyndromic variation in lichens. Systematic Botany, 1, 325–339.CrossRefGoogle Scholar
Culberson, C. F. and Elix, J. A. (1989). Lichen substances. In Methods in Plant Biochemistry, Vol. 1: Plant Phenolics, ed. Dey, P. M. and Harborne, J. B., pp. 509–535. London: Academic Press.Google Scholar
Culberson, C. F. and Johnson, A. (1976). A standardized two-dimensional thin-layer chromatographic method for lichen products. Journal of Chromatography, 128, 253–259.CrossRefGoogle Scholar
Culberson, C. F., Culberson, W. L. and Johnson, A. (1981). A standardized TLC analysis of β-orcinol depsidones. Bryologist, 84, 16–29.CrossRefGoogle Scholar
Culberson, C. F., Culberson, W. L. and Johnson, A. (1985). Orcinol-type depsides and depsidones in the lichens of the Cladonia chlorophaea group (Ascomycotina, Cladoniaceae). Bryologist, 88, 380–387.CrossRefGoogle Scholar
Culberson, C. F., Culberson, W. L. and Johnson, A. (1988). Gene flow in lichens. American Journal of Botany, 75, 1135–1139.CrossRefGoogle Scholar
Culberson, C. F., Hale, M. E. Jr., Tønsberg, T. and Johnson, A. (1984). New depsides from the lichens Dimelaena oreina and Fuscidea viridis. Mycologia, 76, 148–160.CrossRefGoogle Scholar
Culberson, C. F., Nash, T. H. III and Johnson, A. (1979). 3-α-Hydroxybarbatic acid, a new depside in chemosyndromes of some Xanthoparmeliae with β-orcinol depsides. Bryologist, 82, 154–161.CrossRefGoogle Scholar
Culberson, W. L. (1967). Analysis of chemical and morphological variation in the Ramalina siliquosa species complex. Brittonia, 19, 333–52.CrossRefGoogle Scholar
Culberson, W. L. (1986). Chemistry and sibling speciation in the lichen-forming fungi: ecological and biological considerations. Bryologist, 89, 123–131.CrossRefGoogle Scholar
Culberson, W. L. and Culberson, C. F. (1967). Habitat selection by chemically differentiated races of lichens. Science, 158, 1195–1197.CrossRefGoogle ScholarPubMed
Culberson, W. L. and Culberson, C. F. (1968). The lichen genera Cetrelia and Platismatia (Parmeliaceae). Contributions from the United States National Herbarium, 34, 449–558.Google Scholar
Culberson, W. L. and Culberson, C. F. (1994). Secondary metabolites as a tool in ascomycete systematics: lichenized fungi. In Ascomycetes Systematics: Problems and Perspectives in the Nineties, ed. Hawksworth, D. L.. pp. 155–163. New York: Plenum Press.CrossRefGoogle Scholar
Culberson, W. L., Culberson, C. F. and Johnson, A. (1977). Pseudevernia furfuracea – olivetorina relationships and ecology. Mycologia, 69, 604–614.CrossRefGoogle Scholar
Culberson, W. L., Culberson, C. F. and Johnson, A. (1993). Speciation in the lichens of the Ramalina siliquosa complex (Ascomycotina, Ramalinaceae): gene flow and reproductive isolation. American Journal of Botany, 80, 1472–1481.CrossRefGoogle Scholar
Curtis, C. J., Emmett, B. A., Grant, H., et al. (2005). Nitrogen saturation in UK moorlands: the critical role of bryophytes and lichens in determining retention of atmospheric N deposition. Journal of Applied Ecology, 42, 507–517.CrossRefGoogle Scholar
Czehura, S. J. (1977). A lichen indicator of copper mineralization, Lights Creek District, Plumas County, California. Economic Geology, 72, 796–803.CrossRefGoogle Scholar
Dahl, E. (1998). The Phytogeography of Northern Europe (British Isles, Fennoscandia and Adjacent Areas). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Dahl, E. and Krog, H. (1973). Macrolichens of Denmark, Finland, Norway and Sweden. Oslo: Universitetsforlaget.Google Scholar
Dahlkild, A., Kallersjo, M., Lohtander, K. and Tehler, A. (2001). Photobiont diversity in the Physciaceae (Lecanorales). Bryologist, 104, 527–536.CrossRefGoogle Scholar
Dahlman, L. and Palmqvist, K. (2003). Growth in two foliose tripartite lichens Nephroma arcticum and Peltigera aphthosa – empirical modeling of external versus internal factors. Functional Ecology, 17, 821–831.CrossRefGoogle Scholar
Dahlman, L., Näsholm, T. and Palmqvist, K. (2002). Growth, nitrogen uptake, and resource allocation in the two tripartite lichens Nephroma arcticum and Peltigera aphthosa during nitrogen stress. New Phytologist, 153, 307–315.CrossRefGoogle Scholar
Dahlman, L., Persson, J., Näsholm, T. and Palmqvist, K. (2003). Carbon and nitrogen distribution in the green algal lichens Hypogymnia physodes and Platismatia glauca in relation to nutrient supply. Planta, 217, 41–48.Google ScholarPubMed
David, J. C. (1987). Studies on the genus Epigloea. Systema Ascomycetum, 6, 217–221.Google Scholar
David, K. A. and Fay, P. (1977). Effects of long term treatment with C2H2 on N2-fixing microorganisms. Applied Environmental Microbiology, 34, 640–646.Google Scholar
Davis, W. C., Gries, C. and Nash, T. H. III (2000). The ecophysiological response of the aquatic lichen Hydrothyria venosa to nitrate in terms of weight and photosynthesis over long periods of time. Bibliotheca Lichenologica, 75, 201–208.Google Scholar
Davison, J. (1988). Plant beneficial bacteria. Nature Bio/Technology, 6, 282–286.CrossRefGoogle Scholar
Deason, T. R. and Bold, H. C. (1960). Phycological studies. I. Exploratory studies of Texas soil algae. University of Texas Publications, 6022, 1–70.Google Scholar
Debuchy, R. and Turgeon, B. (2006). Mating-type structure, evolution and function in Euascomycetes. In Growth, Differentiation and Sexuality, The Mycota, Vol. 1, ed. Kües, U. and Fischer, R., pp. 293–323. Heidelberg: Springer.CrossRefGoogle Scholar
De Kok, L. J. and Stulen, I. (1993). Role of glutathione in plants under oxidative stress. In Sulfur Nutrition and Assimilation in Higher Plants, ed. Kok, L. J., Stulen, I., Rennenberg, H., Brunold, C. and Rauser, W. E., pp. 295–313. The Hague: SPB Academic Publishing.Google Scholar
Ríos, los A. and Grube, M. (2000). Host-parasite interfaces of some lichenicolous fungi in the Dacampiaceae (Dothideales, Ascomycota). Mycological Research, 104, 1348–1353.CrossRefGoogle Scholar
Ríos, los A., Wierzchos, J., Sancho, L. G., Green, T. G. A. and Ascaso, C. (2005). Ecology of endolithic lichens colonizing granite in continental Antarctica. Lichenologist, 37, 383–395.CrossRefGoogle Scholar
Deltoro, V. I., Gimeno, C., Calatayud, A. and Barreno, E. (1999). Effects of SO2 fumigations on photosynthetic CO2 gas exchange, chlorophyll a fluorescence emission and antioxidant enzymes in the lichens Evernia prunastri and Ramalina farinacea. Physiologia Plantarum, 105, 648–654.CrossRefGoogle Scholar
Dembitsky, V. M. and Tolstikov, G. A. (2005). Organic Metabolites of Lichens. Novosibirsk: Publishing House of SB RAS.Google Scholar
Demmig-Adams, B. (2006). Linking the xanthophyll cycle with thermal energy dissipation. Photosynthesis Research, 76, 73–80.CrossRefGoogle Scholar
Demmig-Adams, B., Máguas, C., Adams, W. W. III, et al. (1990). Effect of high light on the efficiency of photochemical energy conversion in a variety of lichen species with green and blue-green phycobionts. Planta, 180, 400–409.CrossRefGoogle Scholar
Denison, W. C. (1973). Life in tall trees. Scientific American, 228, 74–80.CrossRefGoogle Scholar
Denison, W. C. (1979). Lobaria oregana, a nitrogen-fixing lichen in old-growth Douglas fir forests. In Symbiotic Nitrogen Fixation in the Management of Temperate Forests, ed. Gordon, J. C., Wheeler, C. T. and Perry, D. A., pp. 266–275. Corvallis: Oregon State University.Google Scholar
DePriest, P. T. (1993 a). Variation in the Cladonia chlorophaea complex. I. Morphological and chemical variation in Southern Appalachian populations. Bryologist, 96, 555–563.CrossRefGoogle Scholar
DePriest, P. T. (1993 b). Variation in southern Appalachian populations of the Cladonia chlorophaea complex. II. ribosomal DNA variation. Bryologist, 97, 117–126.CrossRefGoogle Scholar
DePriest, P. T. (1993 c). Small subunit rDNA variation in a population of lichen fungi due to optional group-I introns. Gene, 134, 314–325.CrossRefGoogle Scholar
DePriest, P. T. (2004). Early molecular investigations of lichen-forming symbionts: 1986–2001. Annual Review of Microbiology, 58, 273–301.CrossRefGoogle ScholarPubMed
DePriest, P. T., Ivanova, N. V., Fahselt, D., Alstrup, V. and Gargas, A. (2000). Sequences of psychrophilic fungi amplified from glacier-preserved ascolichens. Canadian Journal of Botany, 78, 1450–1459.CrossRefGoogle Scholar
Queiroz, A. (2005). The resurrection of oceanic dispersal in historical biogeography. Trends in Ecology and Evolution, 20, 68–73.CrossRefGoogle ScholarPubMed
des Abbayes, H. (1953). Travaux sur les lichens parus de 1939 à 1952. Bulletin Sociétié Botanique de France, 100, 83–123.CrossRefGoogle Scholar
Vera, J.-P., Horneck, G., Rettberg, P. and Ott, S. (2003). The potential of the lichen symbiosis to cope with extreme conditions of outer space. I. Influence of UV radiation and space vacuum on the vitality of lichen symbiosis and germination capacity. International Journal of Astrobiology, 1, 285–293.CrossRefGoogle Scholar
Vera, J.-P., Horneck, G., Rettberg, P. and Ott, S. (2004). The potential of the lichen symbiosis to cope with extreme conditions of outer space. II. Germination capacity of lichen ascospores in response to simulated space conditions. Advances in Space Research, 33, 1236–1243.CrossRefGoogle ScholarPubMed
Dibben, M. J. (1971). Whole-lichen culture in a phytotron. Lichenologist, 5, 1–10.CrossRefGoogle Scholar
Diederich, P. (1996). The lichenicolous heterobasidiomycetes. Bibliotheca Lichenologica, 61, 1–198.Google Scholar
Diederich, P. (2003). Review. Bryologist, 106, 629–630.Google Scholar
Dietz, S., Büdel, B., Lange, O. L. and Bilger, W. (2000). Transmittance of light through the cortex of lichens from contrasting habitats. In Aspects in Cryptogamic Research. Contributions in Honour of Ludger Kappen, ed. Schroeter, B., Schlensog, M. and Green, T. G. A., pp. 171–182. Berlin-Stuttgart: Gebrüder Borntraeger Verlagsbuchhandlung.Google Scholar
Divakar, P. K., Crespo, A., Blanco, O. and Lumbsch, H. T. (2006). Phylogenetic significance of morphological characters in the tropical Hypotrachyna clade of parmelioid lichens (Parmeliaceae. Ascomycota). Molecular Phylogenetics and Evolution, 40, 448–458.CrossRefGoogle Scholar
Döbbeler, P. (1984). Symbiosen zwischen Gallertalgen und Gallertpilzen der Gattung Epigloea (Ascomycetes). Beihefte zur Nova Hedwigia, 79, 203–239.Google Scholar
Döbbeler, P. and Poelt, J. (1981). Arthropyrenia endobrya spec nov., eine hepaticole Flechte mit intrazellulärem Thallus aus Brasilien. Plant Systematics and Evolution, 138, 275–281.CrossRefGoogle Scholar
Döring, H. and Lumbsch, H. T. (1998). Ascoma ontogeny: is this character set of any use in the systematics of lichenized ascomycetes? Lichenologist, 30, 489–500.CrossRefGoogle Scholar
Döring, H. and Wedin, M. (2004). Infraspecific variation within Stereocaulon species complexes – genetic markers, individuals, populations and species. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 26. Tartu: Tartu University Press.Google Scholar
Drew, E. A. (1966). Some aspects of the carbohydrate metabolism of lichens. Ph.D. thesis. Oxford: University of Oxford.
Duchesne, L. C. and Larson, W. (1989). Cellulose and the evolution of plant life. BioScience, 4, 238–241.CrossRefGoogle Scholar
Duguay, K. J. and Klironomos, J. M. (2000). Direct and indirect effects of enhanced UV-B radiation on the decomposing and competitive abilities of saprobic fungi. Applied Soil Ecology, 14, 157–164.CrossRefGoogle Scholar
Durrell, L. W. and Newsom, I. E. (1939). Colorado's Poisonous and Injurious Plants. Fort Collins: Colorado Experiment Station.Google Scholar
Dyer, P. S., Murtagh, G. J. and Crittenden, P. D. (2001). Use of RAPD-PCR DNA fingerprinting and vegetative incompatibility tests to investigate genetic variation within lichen-forming fungi. Symbiosis, 31, 213–229.Google Scholar
Easton, R. M. (1994). Lichens and rocks: a review. Geoscience Canada, 21, 59–76.Google Scholar
Ebach, M. C. and Humphries, C. J. (2003). Ontology of biogeography. Journal of Biogeography, 30, 959–962.CrossRefGoogle Scholar
Edmands, S. (1999). Heterosis and outbreeding depression in interpopulation crosses spanning a wide range of divergence. Evolution, 53, 1757–1768.CrossRefGoogle Scholar
Edwards, T. C. Jr., Cutler, D. R., Geiser, L., Alegria, J. and McKenzie, D. (2004). Assessing rarity of species with low detectability: lichens in Pacific Northwest forests. Ecological Applications, 14, 414–424.CrossRefGoogle Scholar
Egea, J. M. (1996). Catalogue of lichenized and lichenicolous fungi of Morocco. Bocconea, 6, 19–114.Google Scholar
Egea, J. M. and Torrente, P. (1993). The lichen genus Bactrospora. Lichenologist, 25, 211–255.CrossRefGoogle Scholar
Egea, J. M. and Torrente, P. (1994). El género de hongos liquenizados Lecanactis (Ascomycotina). Bibliotheca Lichenologica 54, 1–205.Google Scholar
Ekman, S. (1997). The genus Cliostomum revisited. Symbolae Botanicae Upsalienses, 32, 17–28.Google Scholar
Ekman, S. (2001). Molecular phylogeny of the Bacidiaceae (Lecanorales, lichenized Ascomycota). Mycological Research, 105, 783–797.CrossRefGoogle Scholar
Ekman, S. and Fröberg, S. (1988). Taxonomical problems in Aspicilia contorta and A. hoffmannii: an effect of hybridization?International Journal of Mycology and Lichenology, 3, 215–225.Google Scholar
Ekman, S. and Jørgensen, P. M. (2002). Towards a molecular phylogeny for the lichen family Pannariaceae (Lecanorales, Ascomycota). Canadian Journal of Botany, 80, 625–634.CrossRefGoogle Scholar
Ekman, S. and Tønsberg, T. (2002). Most species of Lepraria and Leproloma form a monophyletic group closely related to Stereocaulon. Mycological Research, 106, 1262–1276.CrossRefGoogle Scholar
Ekman, S. and Wedin, M. (2000). The phylogeny of the families Lecanoraceae and Bacidiaceae (lichenized Ascomycota) inferred from nuclear SSU rDNA sequences. Plant Biology, 2, 350–360.CrossRefGoogle Scholar
Elix, J. A. (1991). The lichen genus Relicina in Australasia. In Tropical Lichens: Their Systematics, Conservation and Ecology, ed. Galloway, D. J.. Systematics Association Special Volume 43, pp. 17–34. Oxford: Clarendon Press.Google Scholar
Elix, J. A. (1993). Progress in the generic delimitation of Parmelia sensu lato lichens (Ascomycotina: Parmeliaceae) and a synoptic key to the Parmeliaceae. Bryologist, 96, 359–383.CrossRefGoogle Scholar
Elix, J. A. (1994). Xanthoparmelia. Flora of Australia, 55, 201–308.
Elix, J. A. (1999). Detection and identification of secondary lichen substances: contributions by the Uppsala school of lichen chemistry. Symbolae Botanicae Upsalienses, 32, 103–121.Google Scholar
Elix, J. A. and Ernst-Russell, K. D. (1993). A Catalogue of Standardized Thin Layer Chromatographic Data and Biosynthetic Relationships for Lichen Substances, 2nd edn. Canberra: Australian National University.Google Scholar
Elix, J. A, Giralt, M. and Wardlaw, J. H. (2003). New chloro-depsides from the lichen Dimelaena radiata. Bibliotheca Lichenologica, 86, 1–7.Google Scholar
Elix, J. A., Johnston, J. and Parker, J. L. (1988). A computer program for the rapid identification of lichen substances. Mycotaxon, 31, 89–99.Google Scholar
Ellis, C. J. and Coppins, B. J. (2007). 19th century woodland structure controls stand-scale epiphyte diversity in present-day Scotland. Diversity and Distributions, 13, 84–91.CrossRefGoogle Scholar
Ellis, C. J., Crittenden, P. D., Scrimgeour, C. M. and Ashcroft, C. J. (2005). Translocation of 15N indicates nitrogen recycling in the mat-forming lichen Cladonia portentosa. New Phytologist, 168, 423–434.CrossRefGoogle ScholarPubMed
Elstner, E. F. and Oßwald, W. (1994). Mechanisms of oxygen activation during plant stress. Proceeding of the Royal Society of Edinburgh, 102, 131–154.Google Scholar
Elvebakk, A. and Bjerke, J. W. (2006 a). The Skibotn area in North Norway – an example of very high lichen species richness far to the north. Mycotaxon, 96, 141–146.Google Scholar
Elvebakk, A. and Bjerke, J. W. (2006b). The Skibotn area in North Norway – an example of very high lichen species richness far to the north: a supplement with an annotated list of species. Online: www.mycotaxon.com/resources/weblists.html.
Engelbert, R. and Vonarburg, C. (1995). Lichen diversity and ozone impact in rural areas of central Switzerland. Cryptogamic Botany, 5, 252–263.Google Scholar
Engelmann, M. D. (1966). Energetics, terrestrial field studies, and animal productivity. Advances in Ecological Research, 3, 73–115.CrossRefGoogle Scholar
Englund, B. (1977). The physiology of the lichen Peltigera aphthosa, with special reference to the blue-green phycobiont (Nostoc sp.). Physiologia Plantarum, 41, 298–304.CrossRefGoogle Scholar
Englund, B. (1978). Effects of environmental factors on acetylene reduction by intact thallus and excised cephalodia of Peltigera aphthosa Willd. Ecological Bulletin (Stockholm), 26, 234–246.Google Scholar
Englund, B. and Myerson, H. (1974). In situ measurement of nitrogen fixation at low temperatures. Oikos, 25, 183–187.CrossRefGoogle Scholar
Enriquez, S., Duarte, C. M., Sand-Jensen, K. and Nielsen, S. L. (1996). Broad-scale comparison of photosynthetic rates across phototrophic organisms. Oecologia, 108, 197–206.CrossRefGoogle ScholarPubMed
Eriksson, O. E. (2005). Ascomyceternas ursprung och evolution – Protolichenes-hypotesen. Svensk Mykologisk Tidskrift, 26, 22–29.Google Scholar
Eriksson, O. E. (ed.) (2006 a). Notes on ascomycete systematics Nr. 4324. Myconet, 12, 83–101. Online: www.fieldmuseum.org/myconet/Google Scholar
Eriksson, O. E. (ed.) (2006 b). Outline of Ascomycota, Myconet, 12, 1–82. Online: www.fieldmuseum.org/myconet/.Google Scholar
Ernst, A., Kirschenlohr, H., Diez, J. and Boger, P. (1984). Glycogen content and nitrogenase activity in Anabaena variabilis. Archiv für Microbiologie, 140, 120–125.CrossRefGoogle Scholar
Ertl, L. (1951). Über die Lichtverhältnisse in Laubflecten. Planta, 39, 245–270.CrossRefGoogle Scholar
Ertz, D., Christnach, C., Wedin, M. and Diederich, P. (2005). A world monograph of the genus Plectocarpon (Roccellaceae, Arthoniales). Bibliotheca Lichenologica, 91, 1–155.Google Scholar
Esseen, P.-A. and Renhorn, K. E. (1998). Mass loss of epiphytic lichen litter in a boreal forest. Annals Botanica Fennica, 35, 211–217.Google Scholar
Esseen, P.-A., Ehnström, B., Ericson, L. and Sjöberg, K. (1997). Boreal forests. Ecological Bulletins, 45, 16–47.Google Scholar
Esser, K. (1976). Kryptogamen: Blaualgen, Algen, Pilze, Flechten. Berlin: Springer.CrossRefGoogle Scholar
Esslinger, T. L. (1977). A chemosystematic revision of the brown Parmeliae. Journal of the Hattori Botanical Laboratory, 42, 1–211.Google Scholar
Esslinger, T. L. (1989). Systematics of Oropogon (Alectoriaceae) in the New World. Systematic Botany Monographs, 28, 1–111.CrossRefGoogle Scholar
Ettl, H. and Gärtner, G. (1984). Über die Bedeutung der Cytologie für die Algentaxonomie, dargestellt an Trebouxia (Chlorellales, Chlorophyceae). Plant Systematics and Evolution, 148, 135–147.CrossRefGoogle Scholar
Evans, C. A. and Hutchinson, T. C. (1996). Mercury accumulation in transplanted moss and lichens at high elevation sites in Quebec. Water, Air, and Soil Pollution, 90, 475–488.CrossRefGoogle Scholar
Evans, D. J. A., Archer, S. and Wilson, D. J. H. (1999). A comparison of the lichenometric and Schmidt hammer dating techniques based on data from the proglacial areas of some Icelandic glaciers. Quaternary Science Reviews, 18, 13–41.CrossRefGoogle Scholar
Evans, J. R. (1983). Nitrogen and photosynthesis in the flag leaf of wheat (Triticum aestivum L.). Plant Physiology, 72, 297–302.CrossRefGoogle Scholar
Evans, J. R. (1989). Photosynthesis and nitrogen relationships of C-3 plants. Oecologia, 78, 9–19.CrossRefGoogle Scholar
Evans, R. D. and Johansen, J. R. (1999). Microbiotic crusts and ecosystem processes. Critical Reviews in Plant Sciences, 18, 183–225.CrossRefGoogle Scholar
Evans, R. D., Belnap, J. Garcia-Pichel, F. and Phillips, S. L. (2001). Global change and the future of biological soil crusts. In Biological Soil Crusts: Structure, Function and Management, ed. Belnap, J. and Lange, O. L., pp. 417–429. Berlin: Springer.Google Scholar
Eversman, S. (1978). Effects of low-level SO2 on Usnea hirta and Parmelia chlorochroa. Bryologist, 81, 368–377.CrossRefGoogle Scholar
Eversman, S. and Sigal, L. L. (1987). Effects of SO2, O3, and SO2 and O3 in combination on photosynthesis and ultrastructure of two lichen species. Canadian Journal of Botany, 65, 1806–1818.CrossRefGoogle Scholar
Eversman, S., Johnson, C. and Gustafson, D. (1987). Vertical distribution of epiphytic lichens on three tree species in Yellowstone National Park. Bryologist, 90, 212–216.CrossRefGoogle Scholar
Fahselt, D. (1984). Interthalline variability in levels of lichen products within stands of Cladina stellaris. Bryologist, 87, 50–56.CrossRefGoogle Scholar
Fahselt, D. (1985). Multiple enzyme forms in lichens. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 129–143. New York: Plenum Publishing.CrossRefGoogle Scholar
Fahselt, D. (1987). Electrophoretic analysis of esterase and alkaline phosphatase enzyme forms in single spore cultures of Cladonia cristatella. Lichenologist, 19, 71–75.CrossRefGoogle Scholar
Fahselt, D. (1989). Enzyme polymorphism in sexual and asexual umbilicate lichens from Sverdrup Pass, Ellesmere Island, Canada. Lichenologist, 21, 279–285.CrossRefGoogle Scholar
Fahselt, D. (1991). Enzyme similarity as an indicator of evolutionary divergence: Stereocaulon saxatile H. Magn. Symbiosis, 11, 119–130.Google Scholar
Fahselt, D. (1992). Geothermal effects on multiple enzyme forms in the lichen Cladonia mitis. Lichenologist, 24, 181–192.Google Scholar
Fahselt, D. (1994 a). Secondary biochemistry of lichens. Symbiosis, 16, 117–165.Google Scholar
Fahselt, D. (1994 b). Carbon metabolism in lichens. Symbiosis, 17, 127–182.Google Scholar
Fahselt, D. (1995). Lichen sexuality from the perspective of multiple enzyme forms. Cryptogamic Botany, 5, 137–143.Google Scholar
Fahselt, D. (2001). Analysing lichen enzymes by isoelectricfocussing. In Protocols in Lichenology, ed. Kranner, I., Varma, A. and Beckett, P., pp. 307–331. Berlin: Springer.Google Scholar
Fahselt, D. and Alstrup, V. (1997). High performance liquid chromatography of phenolics in recent and subfossil lichens. Canadian Journal Botany, 75, 1148–1154.CrossRefGoogle Scholar
Fahselt, D. and Hageman, C. (1994). Rhizine and upper thallus isozymes in umbilicate lichens. Symbiosis, 16, 95–103.Google Scholar
Fahselt, D. and Krol, M. (1989). Biochemical comparison of two ecologically distinctive forms of Xanthoria elegans in the Canadian High Arctic. Lichenologist, 21, 135–145.CrossRefGoogle Scholar
Fahselt, D., Krol, M., Hüner, N. and Tønsberg, T. (2000). Pigmentation of Cladonia infected by the lichenicolous fungus Arthrorhaphis aeroguinosa. Lichenologist, 32, 300–303.CrossRefGoogle Scholar
Fahselt, D., Madzia, S. and Alstrup, V. (2001). Scanning electron microscopy of invasive fungi in lichens. Bryologist, 104, 24–39.CrossRefGoogle Scholar
Fahselt, D., Tavares, S. and Mazdia, S. (1997). Isozyme variation in lichens in relation to mine dust exposure. In Progress and Problems in Lichenology in the Nineties, ed. Türk, R. and Zorer, R.. Bibliotheca Lichenologica, 68, 111–127.Google Scholar
Farkas, E. E. and Sipman, H. J. M. (1993). Bibliography and checklist of foliicolous lichenized fungi up to 1992. Tropical Bryology, 7, 93–148.Google Scholar
Farkas, E. and Sipman, H. J. M. (1997). Checklist of foliicolous lichenized fungi – after Farkas and Sipman (1993), with additions to 1996. Abstracta Botanica, 21, 173–206.Google Scholar
Farrar, J. F. (1976 a) Ecological physiology of the lichen Hypogymnia physodes. I. Some effects of constant water saturation. New Phytologist, 77, 93–103.CrossRefGoogle Scholar
Farrar, J. F. (1976 b). Ecological physiology of the lichen Hypogymnia physodes. II. Effects of wetting and drying cycles and the concept of physiological buffering. New Phytologist, 77, 105–113.CrossRefGoogle Scholar
Farrar, J. F. (1976 c). The lichen as an ecosystem: observation and experiment. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 385–406. London: Academic Press.Google Scholar
Farrar, J. F. (1976 d). The uptake and metabolism of phosphate by the lichen Hypogymnia physodes. New Phytologist, 77, 127–134.CrossRefGoogle Scholar
Farrar, J. F. (1988). Physiological buffering. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 101–105. Boca Raton: CRC Press.Google Scholar
Farrar, J. F. and Smith, D. C. (1976). Ecological physiology of the lichen Hypogymnia physodes. III. The importance of the rewetting phase. New Phytologist, 77, 115–125.CrossRefGoogle Scholar
Feige, G. B. (1973). Untersuchungen zur Ökologie und Physiologie der marinen Blaualgenflechte Lichina pygmaea Ag. II. Die Reversibilität der Osmoregulation. Zeitschrift für Pflanzenphyiologie, 68, 415–421.CrossRefGoogle Scholar
Feige, G. B. and Jensen, M. (1992). Basic carbon and nitrogen metabolism of lichens. In Algae and Symbioses, ed. Reisser, W., pp. 277–299. Bristol: Biopress Limited.Google Scholar
Feige, G. B., Lumbsch, H. T., Huneck, S. and Elix, J. A. (1993). The identification of lichen substances by a standardized high-performance liquid chromatographic method. Journal of Chromatography, 646, 417–427.CrossRefGoogle Scholar
Fenn, M. E., Baron, J. S., Allen, E. B., et al. (2003). Ecological effects of nitrogen deposition in the western United States. BioScience, 53, 404–420.CrossRefGoogle Scholar
Ferber, T. (2002). The age and origin of talus cones in the light of lichenometric research. The Skalnisty and Zielony talus cones, High Tatra Mountains, Poland. Studia Geomorphologica Carpatho-Balcanica, 36, 77–90.Google Scholar
Ferraro, L. I., Lücking, R. and Sérusiaux, E. (2001). A world monograph of the lichen genus Gyalectidium (Gomphillaceae). Botanical Journal of the Linnean Society, 137, 311–345.CrossRefGoogle Scholar
Ferry, B. W. and Baddeley, M. S. (1976). Sulphur dioxide uptake in lichens. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 407–418. London: Academic Press.Google Scholar
Feuerer, T. (ed.) (2006). Checklists of lichens and lichenicolous fungi. Version 1 June 2006. Online: www.checklists.de.
Fewer, D., Friedl, T. and Büdel, B. (2002). Chroococcidiopsis and heterocyst-differentiating cyanobacteria are each other's closest living relatives. Molecular Phylogenetics and Evolution, 41, 498–506.Google Scholar
Fiechter, E. (1990). Thallusdifferenzierung und intrathalline Sekundärstoffverteilung bei Parmeliaceae (Lecanorales, lichenisierte Ascomyceten). Inauguraldissertation. Zürich: Universität Zürich.
Fields, R. D. (1988). Physiological responses of lichens to air pollutant fumigations. In Lichens, Bryophytes and Air Quality, ed. Nash, T. H. III and Wirth, V., pp. 175–200. Bibliotheca Lichenologica 30. Berlin-Stuttgart: J. Cramer.Google Scholar
Fields, R. D. and St. Clair, L. L. (1984). The effects of SO2 on photosynthesis and carbohydrate transfer in the two lichens: Collema polycarpon and Parmelia chlorochroa. American Journal of Botany, 71, 986–998.CrossRefGoogle Scholar
Fletcher, A. (1980). Marine and maritime lichens of rocky shores: their ecology, physiology and biological interactions. In The Shore Environment, Vol. 2: Ecosystems, ed. Price, J. H., Irvine, D. E. G. and Farnham, W. F., pp. 789–842. London: Academic Press.Google Scholar
Fletcher, J. (2002). Coordination of cell proliferation and cell fate decisions in the angiosperm shoot apical meristem. BioEssays, 24, 27–37.CrossRefGoogle ScholarPubMed
Fogg, G. E., Fay, P. and Walsby, A. E. (1973). The Bluegreen Algae. London: Academic Press.Google Scholar
Follmann, G. (2002). South America as diversity centre of the lichen family Roccellaceae (Arthoniales). Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg, 30–32, 61–77.Google Scholar
Forman, R. T. T. (1975). Canopy lichens with blue-green algae: a nitrogen source in a Columbian rain forest. Ecology, 56, 1176–1184.CrossRefGoogle Scholar
Forman, R. T. T. and Dowden, D. L. (1977). Nitrogen fixing lichen roles from desert to alpine in the Sangre de Cristo Mountains, New Mexico. Bryologist, 80, 561–70.CrossRefGoogle Scholar
Foyer, C. and Halliwell, B. (1976). The presence of glutathione and glutathione reductase in chloroplasts: a proposed role in ascorbic acid metabolism. Planta, 133, 21–25.CrossRefGoogle ScholarPubMed
Frank, H. A., Young, A., Britton, G. and Cogdell, R. J. (1999). The Photochemistry of Carotenoids. Berlin: Springer.Google Scholar
Freedman, B., Zobens, V., Hutchinson, T. C. and Gizyn, W. I. (1990). Intense, natural pollution affects Arctic tundra vegetation at the Smoking Hills, Canada. Ecology, 71, 492–503.CrossRefGoogle Scholar
Fremstad, E., Paal, J. and Möls, T. (2005). Impacts of increased nitrogen supply on Norwegian lichen-rich alpine communities: a 10-year experiment. Journal of Ecology, 93, 471–481.CrossRefGoogle Scholar
Friedl, T. (1987). Thallus development and phycobionts of the parasitic lichen Diploschistes muscorum. Lichenologist, 19, 183–191.CrossRefGoogle Scholar
Friedl, T. (1989 a). Comparative ultrastructure of pyrenoids in Trebouxia (Microthamniales, Chlorophyta). Plant Systematics and Evolution, 164, 145–159.CrossRefGoogle Scholar
Friedl, T. (1989b). Systematik und Biologie von Trebouxia (Microthamniales, Chlorophyta) als Phycobiont der Parmeliaceae (lichenisierte Ascomyceten). Ph.D. thesis. Bayreuth: Universtät Bayreuth.
Friedl, T. (1993). New aspects of the reproduction by autospores in the lichen alga Trebouxia (Microthamniales, Chlorophyta). Archiv für Protistenkunde, 143, 153–161.CrossRefGoogle Scholar
Friedl, T. and Gärtner, G. (1988). Trebouxia (Pleurastrales, Chlorophyta) as a phycobiont in the lichen genus Diploschistes. Archiv für Protistenkunde, 135, 147–158.CrossRefGoogle Scholar
Friedl, T. and Zeltner, C. (1994). Assessing the relationships of some coccoid green lichen algae and the Microthamniales (Chlorophyta) with 18S rRNA gene sequence comparisons. Journal of Phycology, 30, 500–506.CrossRefGoogle Scholar
Friedmann, E. I. (1982). Endolithic microorganisms in the Antarctic cold desert. Science, 215, 1045–1053.CrossRefGoogle ScholarPubMed
Friedmann, E. I. and Sun, H. J. (2005). Communities adjust their temperature optima by shifting producer-to-consumer ratio, shown in lichens as models. I. Hypothesis. Microbial Ecology, 49, 523–527.CrossRefGoogle ScholarPubMed
Fritz-Sheridan, R. P. (1985). Impact of simulated acid rain on nitrogenase activity in Peltigera aphthosa and P. polydactyla. Lichenologist, 17, 27–31.CrossRefGoogle Scholar
Fritz-Sheridan, R. P. and Coxson, D. S. (1988 a). Nitrogen fixation on the tropical volcano, La Soufrière (Guadeloupe): nitrogen fixation, photosynthesis and respiration during the prevailing cloud/shroud climate by Stereocaulon virgatum. Lichenologist, 20, 41–61.CrossRefGoogle Scholar
Fritz-Sheridan, R. P. and Coxson, D. S. (1988 b). Nitrogen fixation on the tropical volcano, La Soufrière (Guadeloupe): the interaction of temperature, moisture, and light with net photosynthesis and nitrogenase activity in Stereocaulon virgatum and response to periods of insolation shock. Lichenologist, 20, 63–81.CrossRefGoogle Scholar
Fröberg, L., Berg, C. O., Baur, A. and Baur, B. (2001). Viability of lichen photobionts after passing through the digestive tract of a land snail. Lichenologist, 33, 543–545.CrossRefGoogle Scholar
Fujii, I., Watanabe, A., Sankawa, U. and Ebizuka, Y. (2001). Identification of Claisen cyclase domain in fungal polyketide synthase WA, a naphthapyrone synthase of Aspergillus nidulans. Chemical Biology, 8, 187–197.CrossRefGoogle Scholar
Gagunashvili, A. and Andrésson, O. (2006). Heterologous expression of a lichen polyketide synthase in filamentous fungi. EUKETIDES Meeting, Turku, Finland.
Gaio-Oliveira, G., Dahlman, L., Máguas, C. and Palmqvist, K. (2004 a). Growth in relation to microclimatic conditions and physiological characteristics of four Lobaria pulmonaria populations in two contrasting habitats. Ecography, 27, 13–28.CrossRefGoogle Scholar
Gaio-Oliveira, G., Dahlman, L., Martins-Loução, M. A., Máguas, C. and Palmqvist, K. (2005 a). Nitrogen uptake in relation to excess supply and its effects on the lichens Evernia prunastri (L.) Ach and Xanthoria parietina (L.) Th. Fr. Planta, 220, 794–803.CrossRefGoogle Scholar
Gaio-Oliveira, G., Dahlman, L., Palmqvist, K. and Máguas, C. (2004 b). Ammonium uptake in the nitrophytic lichen Xanthoria parietina and its effects on vitality and balance between symbionts. Lichenologist, 36, 75–86.CrossRefGoogle Scholar
Gaio-Oliveira, G., Dahlman, L., Palmqvist, K. and Máguas, C. (2005 b). Responses of the lichen Xanthoria parietina (L.) Th. Fr. to varying thallus nitrogen concentrations. Lichenologist, 37, 171–179.CrossRefGoogle Scholar
Gaio-Oliveira, G., Moen, J., Danell, O. and Palmqvist, K. (2006). Effect of simulated reindeer grazing on the re-growth capacity of mat-forming lichens. Basic and Applied Ecology, 7, 109–121.CrossRefGoogle Scholar
Galley, C. and Linder, H. P. (2006). Geographical affinities of the Cape flora, South Africa. Journal of Biogeography, 33, 236–250.CrossRefGoogle Scholar
Galloway, D. J. (1991 a). Chemical evolution in the order Peltigerales: triterpenoids. Symbiosis, 11, 327–344.Google Scholar
Galloway, D. J. (1991 b). Phytogeography of Southern Hemisphere lichens. In Quantitative Approaches to Phytogeography, ed. Nimis, P. L. and Crovello, T. J., pp. 233–262. Dordrecht: Kluwer Academic.Google Scholar
Galloway, D. J. (1993). Global environmental change: lichens and chemistry. Bibliotheca Lichenologica, 53, 87–95.Google Scholar
Galloway, D. J. (1994 a). Pseudocyphellaria lacerata new to the Faeroe Islands. Lichenologist, 26, 391–393.CrossRefGoogle Scholar
Galloway, D. J. (1994 b). Studies on the lichen genus Sticta (Schreber) Ach. I. Southern South American species. Lichenologist, 26, 223–282.CrossRefGoogle Scholar
Galloway, D. J. (1995 a). The extra-European lichen collections of Archibald Menzies MD, FLS (1754–1842). Edinburgh Journal of Botany, 52, 95–139.CrossRefGoogle Scholar
Galloway, D. J. (1995 b). Studies on the lichen genus Sticta (Schreber) Ach. III. Notes on species described by Bory de St-Vincent, William Hooker, and Delise, between 1804 and 1825. Nova Hedwigia, 61, 147–188.Google Scholar
Galloway, D. J. (1996). Lichen biogeography. In Lichen Biology, ed. Nash, T. H. III, pp. 199–216. Cambridge: Cambridge University Press.Google Scholar
Galloway, D. J. (1997). Studies on the lichen genus Sticta (Schreber) Ach. IV. New Zealand species. Lichenologist, 29, 105–168.CrossRefGoogle Scholar
Galloway, D. J. (1999). Notes on the lichen genus Leptogium (Collemataceae, Ascomycota) in New Zealand. Nova Hedwigia, 69, 317–355.Google Scholar
Galloway, D. J. (2001). Sticta. Flora of Australia, 58A, 78–97.
Galloway, D. J. (2002). Taxonomic notes on the lichen genus Placopsis (Agyriaceae: Ascomycotina) on southern South America, with a key to species. Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg 30–32, 70–107.Google Scholar
Galloway, D. J. (2003). Additional lichen records from New Zealand 40. Buellia aethalea (Ach.) Th. Fr., Catillaria contristans (Nyl.) Zahlbr., Frutidella caesioatra (Schaer.) Kalb, Placynthium rosulans (Th.Fr.) Zahlbr. and Pseudocyphellaria mallota. Australasian Lichenology, 53, 20–29.Google Scholar
Galloway, D. J. (2004 a). Placopsis hertelii (Agyriaceae, Ascomycota) endemic to New Zealand, with descriptions of four additional new species of Placopsis (Nyl.) Linds, from New Zealand. Bibliotheca Lichenologica, 88, 147–161.Google Scholar
Galloway, D. J. (2004 b). New lichen taxa and names in the New Zealand mycobiota. New Zealand Journal of Botany, 42, 105–120.CrossRefGoogle Scholar
Galloway, D. J. (2007). Flora of New Zealand Lichens, Including Lichen-forming and Lichenicolous Fungi, 2nd edn. Lincoln: Manaaki Whenua.Google Scholar
Galloway, D. J. (2008). Southern Hemisphere lichens. New Zealand Journal of Botany (In review,)Google Scholar
Galloway, D. J. and Aptroot, A. (1995). Bipolar lichens: a review. Cryptogamic Botany, 5, 184–191.Google Scholar
Galloway, D. J. and Quilhot, W. (1998) [“1999”]. Checklist of Chilean lichen-forming and lichenicolous fungi. Gayana (Botanica), 55, 111–185.Google Scholar
Galloway, D. J. and Thomas, M. A. (2004). Sticta. In Lichen Flora of the Greater Sonoran Desert Region, Vol. 2, ed. Nash, T. H. III, Ryan, B. D., Diederich, P., Gries, C. and Bungartz, F., pp. 513–524. Tempe: Lichens Unlimited.Google Scholar
Galloway, D. J., Hafellner, J. and Elix, J. A. (2005). Stirtoniella, a new genus for Catillaria kelica (Lecanorales: Ramalinaceae). Lichenologist, 37, 261–271.CrossRefGoogle Scholar
Galloway, D. J., Kantvilas, G. and Elix, J. A. (2001). Pseudocyphellaria. Flora of Australia, 58A, 47–77.
Galun, M. (1988 a). Handbook of Lichenology, Vols. 1, 2 and 3, Boca Raton: CRC Press.Google Scholar
Galun, M. (1988 b). Carbon metabolism. In Handbook of Lichenology, Vol. I, ed. Galun, M., pp. 147–156. Boca Raton: CRC Press.Google Scholar
Galun, M. and Bubrick, P. (1984). Physiological interactions between the partners of the lichen symbiosis. In Cellular Interactions: Encyclopedia of Plant Physiology, ed. Linskens, H. F. and Heslop-Harrison, J., pp. 362–401. Berlin: Springer.CrossRefGoogle Scholar
Galun, M. and Mukhtar, A. (1996). Checklist of the lichens of Israel. Bocconea, 6, 149–171.Google Scholar
Galun, M. and Ronen, R. (1988). Interactions of lichens and pollutants. In CRC Handbook of Lichenology, Vol. III, ed. Galun, M., pp. 55–72. Boca Raton: CRC Press.Google Scholar
Galun, M. and Shomer-Ilan, A. (1988). Secondary metabolic products. In CRC Handbook of Lichenology, Vol. III, ed. Galun, M., pp. 3–8. Boca Raton: CRC Press.Google Scholar
Ganderton, P. and Coker, P. (2005). Environmental Biogeography. Harlow, Essex: Pearson Education Ltd.Google Scholar
Garbaye, J. (1994). Helper bacteria: a new dimension to the mycorrhizal symbiosis. New Phytologist, 128, 197–210.CrossRefGoogle Scholar
Garcia-Molina, F., Hiner, A. N., Fenoll, L. G., et al. (2005). Mushroom tyrosinase: catalase activity, inhibition, and suicide inactivation. Journal of Agricultural and Food Chemistry, 53, 3702–3709.CrossRefGoogle ScholarPubMed
Gargas, A., DePriest, P. T., Grube, M. and Tehler, A. (1995). Multiple origins of lichen symbiosis in fungi suggested by SSU rDNA phylogeny. Science, 268, 1492–1495.CrossRefGoogle ScholarPubMed
Garrett, R. M. (1972). Electrostatic charges on freshly discharged lichen ascospores. Lichenologist, 5, 311–313.CrossRefGoogle Scholar
Gärtner, G. (1985). Die Gattung Trebouxia PUYMALY (Chlorellales, Chlorophyceae). Archiv für Hydrobiologie, Supplementband, 74, (Algological Studies, 41), 495–548.Google Scholar
Gärtner, G. (1992). Taxonomy of symbiotic eukaryotic algae. In Algae and Symbioses, ed. Reisser, W., pp. 325–338. Bristol: Biopress Limited.Google Scholar
Garty, J. (2001). Biomonitoring atmospheric heavy metals with lichens: theory and application. Critical Review in Plant Sciences, 20, 309–371.CrossRefGoogle Scholar
Garty, J., Cohen, Y., Kloog, N. and Karnieli, A. (1997). Effect of air pollution on cell membrane integrity, spectral reflectance and metal and sulfur concentrations in lichens. Environmental Toxicology and Chemistry, 16, 1396–1402.CrossRefGoogle Scholar
Garty, J., Galun, M. and Kessel, M. (1979). Localization of heavy metals and other elements accumulated in the lichen thallus. New Phytologist, 82, 159–168.CrossRefGoogle Scholar
Garty, J.Karary, Y., Harel, J. and Lurie, S. (1993). The impact of air pollution on the integrity of cell membranes and chlorophyll in the lichen Ramalina duriaei (De Not.) Bagl. transplanted to industrial sites in Israel. Archives of Environmental Contamination and Toxicology, 24, 455–460.CrossRefGoogle Scholar
Garty, J., Perry, A. S. and Mozel, J. (1982). Accumulation of polychlorinated biphenyls (PCBs) in the transplanted lichen Ramalina duriaei in air quality biomonitoring experiments. Nordic Journal of Botany, 2, 583–586.CrossRefGoogle Scholar
Gaugh, H. G. Jr. (1982). Multivariate Analysis in Community Ecology. Cambridge: Cambridge University Press.Google Scholar
Gauslaa, Y. (2006). Trade-off between reproduction and growth in the foliose old forest lichen Lobaria pulmonaria. Basic and Applied Ecology, 7, 455–460.CrossRefGoogle Scholar
Gauslaa, Y. and McEvoy, M. (2005). Seasonal changes in solar radiation drive acclimation of the sun-screening compound parietin in the lichen Xanthoria parietina. Basic and Applied Ecology, 6, 75–82.CrossRefGoogle Scholar
Gauslaa, Y. and Solhaug, K. A. (1998). The significance of thallus size for the water economy of the cyanobacterial old-forest lichen Degelia plumbea. Oecologia, 116, 76–84.CrossRefGoogle ScholarPubMed
Gauslaa, Y. and Solhaug, K. A. (1999). High-light damage in air-dry thalli of the old forest lichen Lobaria pulmonaria – interactions of irradiance, exposure duration and high temperature. Journal of Experimental Botany, 50, 697–705.Google Scholar
Gauslaa, Y. and Solhaug, K. A. (2001). Fungal melanins as a sun screen for symbiotic green algae in the lichen Lobaria pulmonaria. Oecologia, 126, 462–471.CrossRefGoogle ScholarPubMed
Gauslaa, Y. and Ustvedt, E. M. (2003). Is parietin a UV-B or a blue-light screening pigment in the lichen Xanthoria parietina?Photochemical and Photobiological Sciences, 2, 424–432.CrossRefGoogle Scholar
Gauslaa, Y., Holien, H., Ohlson, M. and Solhøy, T. (2006 a). Does snail grazing affect growth of the old forest lichen Lobaria pulmonaria?Lichenologist, 38, 587–593.CrossRefGoogle Scholar
Gauslaa, Y., Lie, M., Solhaug, K. and Ohlson, M. (2006 b). Growth and ecophysiological acclimation of the foliose lichen Lobaria pulmonaria in forests with contrasting light climates. Oecologia, 147, 406–416.CrossRefGoogle ScholarPubMed
Gauslaa, Y., Ohlson, M., Solhaug, K. A., Bilger, W. and Nybakken, L. (2001). Aspect dependent high-irradiance damage to two transplanted foliose forest lichens Lobaria pulmonaria and Parmelia sulcata. Canadian Journal of Forest Research, 31, 1639–1649.CrossRefGoogle Scholar
Gaya, E., Lutzoni, F., Zoller, S. and Navarro-Rosinés, P. (2003). Phylogenetic study of Fulgensia and allied Caloplaca and Xanthoria species (Teloschistaceae, lichen-forming Ascomycota). American Journal of Botany, 90, 1095–1103.CrossRefGoogle Scholar
Geebelen, W. and Hoffmann, M. (2001). Evaluation of bio-indication methods using epiphytes by correlating with SO2-pollution parameters. Lichenologist, 33, 249–260.CrossRefGoogle Scholar
Gehrig, H., Schüssler, A. and Kluge, M. (1996). Geosiphon pyriforme, a fungus forming endocytobiosis with Nostoc (cyanobacteria), is an ancestral member of the Glomales: evidence by SSU rRNA analysis. Journal of Molecular Evolution, 43, 71–81.CrossRefGoogle ScholarPubMed
Geitler, L. (1932). Cyanophyceae von Europa unter Berücksichtigung der anderen Kontinente. In Rabenhorst's Kryptogamenflora von Deutschland, Österreich und der Schweiz, 2nd edn., Vol. 14, ed. Kolkwitz, R., pp. 1–1196. Leipzig: Akademische Verlagsgesellschaft.Google Scholar
Geitler, L. (1934). Beiträge zur Kenntnis der Flechtensymbiose. IV, V. Archiv für Protistenkunde 82, 51–85.Google Scholar
Geitler, L. (1937). Beiträge zur Kenntnis der Flechtensymbiose. VI. Die Verbindung von Pilz und Alge bei den Pyrenopsidaceen Synalissa, Thyrea, Peccania und Psorotichia. Archiv für Protistenkunde, 88, 161–179.Google Scholar
Gerson, U. and Seaward, M. R. D. (1977). Lichen-invertebrate associations. In Lichen Ecology, ed. Seaward, M. R. D., pp. 69–119. London: Academic Press.Google Scholar
Geyer, M., Feuerer, T. and Feige, G. B. (1984). Chemie und Systematik in der Flechtengattung Rhizocarpon: Hochdruckflüssigkeitschromatographie (HPLC) der Flechten-Sekundärstoffe der Rhizocarpon superficiale-Gruppe. Plant Systematics and Evolution, 145, 41–54.CrossRefGoogle Scholar
Gignac, D. and Dale, M. R. T. (2005). Effects of fragment size and habitat heterogeneity on cryptogam diversity in the low-boreal forest of western Canada. Bryologist, 108, 50–66.CrossRefGoogle Scholar
Gilbert, O. L. (1970). Further studies on the effect of sulphur dioxide on lichens and bryophytes. New Phytologist, 69, 605–627.CrossRefGoogle Scholar
Gilbert, O. L. (1971). The effect of airborne fluorides on lichens. Lichenologist, 5, 26–32.CrossRefGoogle Scholar
Gilbert, O. L. (1985). Environmental effects of airborne fluorides from aluminium smelting at Invergordon, Scotland 1971–1983. Environmental Pollution, Series A, 39, 293–302.CrossRefGoogle Scholar
Gilbert, O. L. (1992). Lichen reinvasion with declining air pollution. In Bryophytes and Lichens in a Changing Environment, ed. Bates, J. W. and Farmer, A. M., pp. 159–177. Oxford: Clarendon Press.Google Scholar
Gillespie, J. H. (1991). The Causes of Molecular Evolution. Oxford: Oxford University Press.Google Scholar
Gjerde, I., Sætersdal, M., Rolstad, J., et al. (2005). Productivity–diversity relationships for plants, bryophytes, lichens and polypore fungi in six northern forest landscapes. Ecography, 28, 705–720.CrossRefGoogle Scholar
Gob, F., Oetit, F., Bravard, J. P., Ozer, A. and Gob, A. (2003). Lichenometric application to historical and subrecent dynamics and sediment transport of a Corsican stream (Figarella River, France). Quaternary Science Reviews, 22, 2111–2124.CrossRefGoogle Scholar
Goebel, K. (1926). Die Wasseraufnahme der Flechten. Berichte der deutschen botanischen Gesellschaft, 44, 158–161.Google Scholar
Goffinet, B. and Bayer, R. J. (1997). Characterization of mycobionts of photomorph pairs in the Peltigerineae (lichenized ascomycetes) based on internal transcribed spacer sequences of the nuclear ribosomal DNA. Fungal Genetics and Biology, 21, 228–237.CrossRefGoogle ScholarPubMed
Goldner, W. R., Hoffman, F. M. and Medve, R. J. (1986). Allelopathic effects of Cladonia cristatella on ectomycorrhizal fungi common to bituminous strip-mine spoils. Canadian Journal of Botany, 64, 1586–1590.CrossRefGoogle Scholar
Golm, G. T., Hill, P. S. and Wells, H. (1993). Life expectancy in a Tulsa cemetery: growth and population structure of the lichen Xanthoparmelia cumberlandia. American Midland Naturalist, 129, 373–383.CrossRefGoogle Scholar
Gombert, S., Asta, J. and Seaward, M. R. D. (2003). Correlation between the nitrogen concentration of two epiphytic lichens and the traffic density in an urban area. Environmental Pollution, 123, 281–290.CrossRefGoogle Scholar
Gordy, V. R. and Hendrix, D. L. (1982). Respiratory response of the lichens Ramalina stenospora Mull. Arg. and Ramalina complanata (Sw.) Ach. to azide, cyanide, salicylhydroxamic acid and bisulfate during thallus hydration. Bryologist, 85, 361–374.CrossRefGoogle Scholar
Gorin, P. A. J., Baron, M. and Iacomini, M. (1988). Storage products of lichens. In CRC Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 9–24. Boca Raton: CRC Press.Google Scholar
Gough, L. P., Severson, R. C. and Jackson, L. L. (1988). Determining baseline element composition of lichens. I. Parmelia sulcata at Theodore Roosevelt National Park, North Dakota. Water, Air, and Soil Pollution, 38, 157–167.Google Scholar
Goward, T. (1994). Notes on old growth-dependent epiphytic macrolichens in inland British Columbia, Canada. Acta Botanica Fennica, 150, 31–38.Google Scholar
Goward, T. (1995). Nephroma occultum and the maintenance of lichen diversity in British Columbia. Mitteilungen der Eidgenössischen Forschungsanstalt für Wald, Schnee und Landschaft, 70, 93–101.Google Scholar
Goward, T. (1996). Lichens of British Columbia: Rare Species and Priorities for Inventory. Victoria: Province of British Columbia, Ministry of Forests Research Program. Working Paper 08/1996, pp. i–viii + 1–34.
Goward, T. and Ahti, T. (1997). Notes on the distributional ecology of the Cladoniaceae (lichenized ascomycetes) in temperate and boreal western North America. Journal of the Hattori Botanical Laboratory, 82, 143–155.Google Scholar
Goyal, R. and Seaward, M. R. D. (1982). Metal uptake in terricolous lichens. III. Translocation in the thallus of Peltigera canina. New Phytologist, 90, 85–98.CrossRefGoogle Scholar
Grace, B., Gillespie, T. J. and Puckett, K. J. (1985 a). Sulphur dioxide threshold concentration values for Cladina rangiferina in the Mackenzie Valley, N. W. T. Canadian Journal of Botany, 63, 806–812.CrossRefGoogle Scholar
Grace, B., Gillespie, T. J. and Puckett, K. J. (1985 b). Uptake of gaseous sulphur dioxide by the lichen Cladina rangiferina. Canadian Journal of Botany, 63, 797–805.CrossRefGoogle Scholar
Grace, J. (1997). Toward models of resource allocation by plants. In Plant Resource Allocation, ed. Bazzaz, F. A. and Grace, J., pp. 279–291. San Diego: Academic Press.Google Scholar
Gradstein, S. and Lücking, R. (1997). Synthesis of the Symposium (on Foliicolous Cryptogams) and priorities for future research. Abstracta Botanica, 21, 207–214.Google Scholar
Grant, B. S., Owen, D. F. and Clarke, C. A. (1996). Parallel rise and fall of melanic peppered moths in America and Britain. Journal of Heredity, 87, 351–357.CrossRefGoogle Scholar
Green, T. G. A. and Lange, O. L. (1995). Photosynthesis in poikilohydric plants: a comparison of lichens and bryophytes. In Ecophysiology of Photosynthesis, ed. Schulze, E. D. and Caldwell, M. M., pp. 71–101. Berlin: Springer.Google Scholar
Green, T. G. A., Büdel, B., Heber, U., et al. (1993). Differences in photosynthetic performance between cyanobacterial and green algal components of lichen photosymbiodemes measured in the field. New Phytologist, 125, 723–731.CrossRefGoogle Scholar
Green, T. G. A, Büdel, B., Meyer, A., Zellner, H. and Lange, O. L. (1997). Temperate rainforest lichens in New Zealand: light response of photosynthesis. New Zealand Journal of Botany, 35, 493–504.CrossRefGoogle Scholar
Green, T. G. A., Horstmann, J., Bonnett, H., Wilkins, A. and Silvester, W. B. (1980). Nitrogen fixation by members of the Stictaceae (Lichens) of New Zealand. New Phytologist, 84, 339–348.CrossRefGoogle Scholar
Green, T. G. A., Meyer, A., Büdel, B., Zellner, H. and Lange, O. L. (1995). Diel patterns of CO2-exchange for six lichens from a temperate rain forest in New Zealand. Symbiosis, 18, 251–273.Google Scholar
Green, T. G. A., Schlensog, M., Sancho, L. G., et al. (2002). The photobiont (cyanobacterial or green algal) determines the pattern of photosynthetic activity within a lichen photosymbiodeme: evidence obtained from in situ measurements of chlorophyll a fluorescence. Oecologia, 130, 191–198.CrossRefGoogle Scholar
Green, T. G. A., Schroeter, B., Kappen, L., Seppelt, R. D. and Maseyk, K. (1998). An assessment of the relationship between chlorophyll a fluorescence and CO2 gas exchange from field measurements on a moss and lichen. Planta, 206, 611–618.CrossRefGoogle Scholar
Green, T. G. A., Schroeter, B. and Sancho, L. G. (1999). Plant life in Antarctica. In Handbook of Functional Plant Ecology, ed. Pugnaire, F. I. and Valladares, F., pp. 495–543. New York: Marcel Dekker, Inc.Google Scholar
Green, T. G. A., Schroeter, B. and Sancho, L. G. (2007). Plant life in Antarctica. In Functional Plant Ecology, 2nd edn., ed. Pugnaire, F. I. and Valladares, F., pp. 389–433. New York: Marcel Dekker.Google Scholar
Green, T. G. A., Snelgar, W. P. and Wilkins, A. L. (1985). Photosynthesis, water relations and thallus structure of Stictaceae lichens. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 57–75. New York: Plenum Press.CrossRefGoogle Scholar
Gregory, K. J. (1975). Lichens and the determination of river channel capacity. Earth Surface Processes, 1, 273–285.CrossRefGoogle Scholar
Grehan, J. R. (2001). Panbiogeography from tracks to ocean basins: evolving perspectives. Journal of Biogeography, 28, 413–429.CrossRefGoogle Scholar
Gries, C., Nash, T. H. III and Kesselmeier, J. (1994). Exchange of reduced sulfur gases between lichens and the atmosphere. Biogeochemistry, 23, 25–39.CrossRefGoogle Scholar
Gries, C., Romagni, J. G., Nash, T. H. III, Kuhn, U. and Kesselmeier, J. (1997 a). The relation of H2S release to SO2. New Phytologist, 136, 703–711.CrossRefGoogle Scholar
Gries, C., Sanz, M.-J. and Nash, T. H. III (1995). The effect of SO2 fumigation on CO2 gas exchange, chlorophyll fluorescence and chlorophyll degradation in different lichen species from western North America. Cryptogamic Botany, 5, 239–246.Google Scholar
Gries, C., Sanz, M.-J., Romagni, J. G., et al. (1997 b). The uptake of gaseous sulphur dioxide by non-gelatinous lichens. New Phytologist, 135, 595–602.CrossRefGoogle Scholar
Griffith, M. and Yaish, M. W. F. (2004). Antifreeze proteins in overwintering plants: a tale of two activities. Trends in Plant Science, 9, 399–405.CrossRefGoogle ScholarPubMed
Grime, J. P. (1979). Plant Strategies and Vegetation Processes. Chichester: Wiley.Google Scholar
Grime, J. P., Hodgson, J. P. and Hunt, R. (1988). Comparative Plant Ecology: a Functional Approach to Common British Species. London: Unwin Hyman.CrossRefGoogle Scholar
Grodzinska, K., Godzik, B. and Bienkowski, P. (1999). Cladina stellaris (Opiz) Brodo as a bioindicator of atmospheric deposition on the Kola Peninsula, Russia. Polar Research, 18, 105–110.CrossRefGoogle Scholar
Groombridge, B., ed. (1992). Global Biodiversity: Status of the Earth's Living Resources. London: Chapman and Hall.CrossRefGoogle Scholar
Grube, M. (1998). Classification and phylogeny in the Arthoniales (lichenized Ascomycetes). Bryologist, 101, 377–391.CrossRefGoogle Scholar
Grube, M. and Blaha, J. (2003). On the pylogeny of some polyketide synthase genes in the lichenized genus Lecanora. Mycological Research, 107, 1419–1426.CrossRefGoogle Scholar
Grube, M. and Los Ríos, A. (2001). Observations on Biatoropsis usnearum, a lichenicolous heterobasidiomycete, and other gall-forming fungi, using different microscopical techniques. Mycological Research, 105, 1116–1122.CrossRefGoogle Scholar
Grube, M. and Hafellner, F. (1990). Studien an flechtenbewohnenden Pilzen der Sammelgattung Didymella (Ascomycetes, Dothideales). Nova Hedwigia, 51, 283–360.Google Scholar
Grube, M. and Kantvilas, G. (2006). Siphula represents a remarkable case of morphological congruence in sterile lichens. Lichenologist, 38, 241–249.CrossRefGoogle Scholar
Grube, M. and Kroken, S. (2000). Molecular approaches and the concept of species and species complexes in lichenized fungi. Mycological Research, 104, 1284–1294.CrossRefGoogle Scholar
Grube, M., Baloch, E. and Lumbsch, H. T. (2004). The phylogeny of Porinaceae (Ostropomycetidae) suggests a neotenic origin of perithecia in Lecanoromycetes. Mycological Research, 108, 1111–1118.CrossRefGoogle ScholarPubMed
Guenther, J. E. and Melis, A. (1990). Dynamics of photosystem II heterogeneity in Dunaliella salina (green alga). Photosynthesis Research, 23, 195–203.CrossRefGoogle Scholar
Gugger, M. F. and Hoffmann, L. (2004). Polyphyly of true branching cyanobacteria (Stigonematales). International Journal of Systematic and Evolutionary Microbiology, 54, 349–357.CrossRefGoogle Scholar
Gunn, J., Keller, W., Negusanti, J., et al. (1995). Ecosystem recovery after emission reductions: Sudbury, Canada. Water, Air, and Soil Pollution, 85, 1783–1788.CrossRefGoogle Scholar
Gunther, A. J. (1988). Effects of simulated acid rain on nitrogenase activity in the lichen genus Peltigera under field and laboratory conditions. Water, Air, and Soil Pollution, 38, 379–385.Google Scholar
Gunther, A. J. (1989). Nitrogen fixation by lichens in a subarctic Alaskan watershed. Bryologist, 92, 202–208.CrossRefGoogle Scholar
Gupta, V. (2005). Application of lichenometry to slided materials in the Higher Himalayan landslide zone. Current Science, 89, 1032–1036.Google Scholar
Haas, D. and Keel, C. (2003). Regulation of antibiotic production in root-colonizing Pseudomonas spp. and relevance for biological control of plant disease. Annual Review of Phytopathology, 41, 117–153.CrossRefGoogle ScholarPubMed
Hafellner, J. (1984). Studien in Richtung einer naturlicheren Gliederung der Sammelfamilien Lecanoraceae und Lecideaceae. Beitrage zur Lichenologie. Festschrift J. Poelt. Beihefte zur Nova Hedwigia, 79, 241–371.Google Scholar
Hafellner, J. (1995) A new checklist of lichens and lichenicolous fungi of insular Laurimacaronesia including a lichenological bibliography for the area. Fristchiana 5, 3–132.Google Scholar
Hafellner, J., Hertel, H., Rambold, G. and Timdal, E. (1993). A New Outline of the Lecanorales. Graz: Privately published by the authors.Google Scholar
Häffner, E., Lomský, B., Hynek, V.,et al. (2001). Air pollution and lichen physiology. Physiological responses of different lichens in a transplant experiment following an SO2-gradient. Water, Air, and Soil Pollution, 131, 185–201.CrossRefGoogle Scholar
Hageman, C. M. (1989). Enzyme electromorph variation in the lichen family Umbilicariaceae. Ph.D. thesis. London, Ontario: University of Western Ontario.
Hageman, C. M. and Fahselt, D. (1986). A comparison of isozyme patterns of morphological variants in the lichen Umbilicaria muhlenbergii (Ach.) Tuck. Bryologist, 89, 285–290.CrossRefGoogle Scholar
Hageman, C. M. and Fahselt, D. (1990). Enzyme electromorph variation in the lichen family Umbilicariaceae: within-stand polymorphism in umbilicate lichens of eastern Canada. Canadian Journal of Botany, 68, 2636–2643.CrossRefGoogle Scholar
Hageman, C. M. and Fahselt, D. (1992). Geographical distance and enzyme polymorphisms in the lichen Umbilicaria mammulata. Bryologist, 93, 316–323.CrossRefGoogle Scholar
Hahn, S. C., Tenhunen, J. D., Popp, P. W., Meyer, A. and Lange, O. L. (1993). Upland tundra in the foothills of the Brooks Range, Alaska: diurnal CO2 exchange patterns of characteristic lichen species. Flora, 188, 125–143.CrossRefGoogle Scholar
Hale, M. E. (1973). Growth. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 473–492. New York: Academic Press.Google Scholar
Hale, M. E. (1974). The Biology of Lichens. 2nd edn. London: Edward Arnold.Google Scholar
Hale, M. E. (1983). The Biology of Lichens. 3rd edn. London: Edward Arnold.Google Scholar
Hale, M. E. (1984). The lichen line and high water levels in a freshwater stream in Florida. Bryologist, 87, 261–265.CrossRefGoogle Scholar
Hale, M. E. (1990). A synopsis of the lichen genus Xanthoparmelia (Vainio) Hale (Ascomycotina, Parmeliaceae). Smithsonian Contributions to Botany, 74, 1–250.CrossRefGoogle Scholar
Hallbom, L. and Bergman, B. (1979). Influence of certain herbicides and a forest fertilizer on the nitrogen fixation by the lichen Peltigera praetextata. Oecologia, 40, 19–27.CrossRefGoogle Scholar
Hällgren, J.-E. and Huss, K. (1975). Effects of SO2 on photosynthesis and nitrogen fixation. Physiologia Plantarum, 34, 171–176.CrossRefGoogle Scholar
Hallingbäck, T. (1991). Blue-green algae and cyanophilic lichens are threatened by air pollution and fertilization. Svensk Botanisk Tidskrift, 85, 87–104.Google Scholar
Hallingbäck, T. and Kellner, O. (1992). Effects of simulated nitrogen rich and acid rain on the nitrogen-fixing lichen Peltigera aphthosa (L.) Willd. New Phytologist, 120, 99–103.CrossRefGoogle Scholar
Halliwell, B. (2006). Reactive species and antioxidants. Redox biology is a fundamental theme of aerobic life. Plant Physiology, 141, 312–322.CrossRefGoogle ScholarPubMed
Halliwell, B. and Gutteridge, J. M. C. (1999). Free Radicals in Biology and Medicine. Oxford: Oxford University Press.Google Scholar
Hamada, N., Miyawaki, H. and Yamada, A. (1995). Distribution pattern of air pollution and epiphytic lichens in the Osaka Plain (Japan). Journal of Plant Research, 108, 483–491.CrossRefGoogle Scholar
Hamada, N., Tanahashi, T., Miyagawa, H. and Miyawaki, H. (2001). Characteristics of secondary metabolites from isolated lichen mycobionts. Symbiosis, 31, 23–33.Google Scholar
Hammer, S. (2003). Notocladonia, a new genus in the Cladoniaceae. Bryologist, 106, 162–167.CrossRefGoogle Scholar
Hardy, R. W. F., Burns, R. C. and Holsten, R. D. (1973). Applications of the C2H2 reduction assay for measurement of N2 fixation. Soil Biology and Biochemistry, 5, 47–81.CrossRefGoogle Scholar
Harper, K. T. and Marble, J. R. (1988). A role of nonvascular plants in management of semiarid rangelands. In Vegetation Science Applications for Rangeland Analysis and Management, ed. Tueller, P. T., pp. 135–169. London: Kluwer Academic.CrossRefGoogle Scholar
Harris, G. B. (1971). The ecology of corticolous lichens. I. The zonation on oak and birch in South Devon. Journal of Ecology, 59, 431–439.CrossRefGoogle Scholar
Harrison, S. and Winchester, V. (2000). Nineteenth- and twentieth-century glacier fluctuations and climatic implications in the Arco and Colonia valleys, Hielo Patagónico Norte, Chile. Arctic, Antarctic, and Alpine Research, 32, 55–63.CrossRefGoogle Scholar
Hasegawa, M., Kishino, H. and Yano, K. (1985). Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. Journal of Molecular Evolution, 22, 160–174.CrossRefGoogle ScholarPubMed
Hasenhüttl, G. and Poelt, J. (1978). Über die Brutkörner bei der Flechtengattung Umbilicaria. Berichte der deutschen botanischen Gesellschaft, 91, 275–296.Google Scholar
Hasse, H. E. (1913). The lichen flora of southern California. Contributions from the United States National Herbarium, 17, 1–132.Google Scholar
Hauck, M. and Paul, A. (2005). Manganese as a site factor for epiphytic lichens. Lichenologist, 37, 409–423.CrossRefGoogle Scholar
Hauck, M. and Spribille, T. (2005). The significance of precipitation and substrate chemistry for epiphytic lichen diversity in spruce-fir forests of the Salish Mountains, northwestern Montana. Flora, 200, 547–562.CrossRefGoogle Scholar
Hauck, M. and Zoller, T. (2003). Copper sensitivity of soredia of the epiphytic lichen Hypogymnia physodes. Lichenologist, 35, 271–274.CrossRefGoogle Scholar
Hauck, M., Hesse, V., Jung, R., Zöller, T. and Runge, M. (2001). Long-distance transported sulphur as a limiting factor for the abundance of Lecanora conizaeoides in montane spruce forests. Lichenologist, 33, 267–269.CrossRefGoogle Scholar
Hauck, M., Hesse, V. and Runge, M. (2002 a). Correlations between the Mn/Ca ratio in stemflow and epiphytic lichen abundance in a dieback-affected spruce forest of the Harz Mountains, Germany. Flora, 197, 361–369.CrossRefGoogle Scholar
Hauck, M., Mulack, C. and Paul, A. (2002 b). Manganese uptake in the epiphytic lichens Hypogymnia physodes and Lecanora conizaeoides. Environmental and Experimental Botany, 48, 107–117.CrossRefGoogle Scholar
Hauck, M., Paul, A., Mulack, C., Fritz, E. and Runge, M. (2002 c). Effects of manganese on the viability of vegetative diaspores of the epiphytic lichen Hypogymnia physodes. Environmental and Experimental Botany, 47, 127–142.CrossRefGoogle Scholar
Hauck, M., Paul, A. and Spribille, T. (2006). Uptake and toxicity of manganese in epiphytic cyanolichens. Environmental and Experimental Botany, 56, 216–224.CrossRefGoogle Scholar
Hawksworth, D. L. (1971). Lichens as a litmus for air pollution: a historical review. International Journal of Environmental Studies, 1, 281–296.CrossRefGoogle Scholar
Hawksworth, D. L. (1982). Secondary fungi in lichen symbioses: parasites, saprophytes and parasymbionts. Journal of the Hattori Botanical Laboratory, 52, 357–366.Google Scholar
Hawksworth, D. L. (1983). A key to the lichen-forming, parasitic, parasymbiotic and saprophytic fungi occurring on lichens in the British Isles. Lichenologist, 15, 1–44.CrossRefGoogle Scholar
Hawksworth, D. L. (1985). Problems and prospects in the systematics of the Ascomycotina. Proceedings of the Indian Academy of Sciences (Plant Sciences), 94, 319–39.Google Scholar
Hawksworth, D. L. (1988 a). The fungal partner. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 35–38. Boca Raton: CRC Press.Google Scholar
Hawksworth, D. L. (1988 b). The variety of fungal-algal symbioses, their evolutionary significance, and the nature of lichens. Botanical Journal of the Linnean Society, 96, 3–20.CrossRefGoogle Scholar
Hawksworth, D. L. (1988 c). Effects of algae and lichen-forming fungi on tropical crops. In Perspectives of Mycopathology, ed. Agnihotri, V. P., Sarbhoy, K. A. and Kumar, D., pp. 76–83. New Delhi: Malhorta Publishing House.Google Scholar
Hawksworth, D. L. (1988 d). Conidiomata, conidiogenesis, and conidia. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 181–193. Boca Raton: CRC Press.Google Scholar
Hawksworth, D. L. (1991). The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycological Research, 95, 641–655.CrossRefGoogle Scholar
Hawksworth, D. L. (2002). Bioindication: calibrated scales and their utility. In Monitoring with Lichens – Monitoring Lichens, Nato Science Series IV: Earth and Environmental Sciences, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 11–20. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Hawksworth, D. L. (2003). The lichenicolous fungi of Great Britain and Ireland: an overview and annotated checklist. Lichenologist, 35, 191–232.CrossRefGoogle Scholar
Hawksworth, D. L. and Hill, D. J. (1984). The Lichen-forming Fungi. Glasgow: Blackie.CrossRefGoogle Scholar
Hawksworth, D. L. and Honegger, R. (1994). The lichen thallus: a symbiotic phenotype of nutritionally specialized fungi and its response to gall producers. In Plant Galls: Organisms, Interactions, Populations, ed. Williams, M. A. J., pp. 77–98. Oxford: Clarendon Press.Google Scholar
Hawksworth, D. L. and McManus, P. M. (1989). Lichen recolonization in London under conditions of rapidly falling sulphur dioxide levels, and the concept of zone skipping. Botanical Journal of the Linnean Society, 100, 99–109.CrossRefGoogle Scholar
Hawksworth, D. L. and McManus, P. M. (1992). Changes in the lichen flora on trees in Epping Forest through periods of increasing and then ameliorating sulphur dioxide air pollution. Essex Naturalist, 11, 92–101.Google Scholar
Hawksworth, D. L. and Rose, F. (1970). Qualitative scale for estimating sulphur dioxide air pollution in England and Wales using epiphytic lichens. Nature, 227, 145–148.CrossRefGoogle ScholarPubMed
Hawksworth, D. L., Kirk, P. M., Sutton, B. C. and Pegler, D. N. (1995). Ainsworth and Bisby's Dictionary of the Fungi. 8th edn. Wallingford: CAB International.Google Scholar
Hawksworth, D. L., Lawton, R. M., Martin, P. G. and Stanley-Price, K. (1984). Nutritive value of Ramalina duriaei grazed by gazelles in Oman. Lichenologist, 16, 93–94.CrossRefGoogle Scholar
Hawksworth, D. L., Sutton, B. C. and Ainsworth, D. C. (1983). Ainsworth and Bisby's Dictionary of the Fungi. 7th edn. Kew: Commonwealth Mycological Institute.Google Scholar
Hayward, G. D. and Rosentreter, R. (1994). Lichens as nesting material for northern flying squirrels in the northern Rocky Mountains. Journal of Mammalogy, 75, 663–673.CrossRefGoogle Scholar
Heads, M. (1997). Regional patterns of biodiversity in New Zealand: one degree grid analysis of plant and animal distributions. Journal of the Royal Society of New Zealand, 27, 337–354.CrossRefGoogle Scholar
Heads, M. (1998). Biogeographic disjunction along the Alpine Fault, New Zealand. Biological Journal of the Linnean Society, 63, 161–176.CrossRefGoogle Scholar
Heads, M. (2001). Birds of paradise, biogeography and ecology in New Guinea: a review. Journal of Biogeography, 28, 893–927.CrossRefGoogle Scholar
Heads, M. (2004). What is a node?Journal of Biogeography, 31, 1883–1891.CrossRefGoogle Scholar
Heads, M. (2005). Towards a panbiogeography of the seas. Biological Journal of the Linnean Society, 84, 675–723.CrossRefGoogle Scholar
Heads, M. and Craw, R. (2004). The alpine fault biogeographic hypothesis revisited. Cladistics, 20, 184–190.CrossRefGoogle Scholar
Heber, U., Bilger, W. and Shuvalov, V. A. (2006 a). Thermal energy dissipation in reaction centers and in the antenna of photosystem II protects desiccated poikilohydric mosses against photo-oxidation. Journal of Experimental Botany, 57, 2993–3006.CrossRefGoogle Scholar
Heber, U., Lange, O. L. and Shuvalov, V. A. (2006 b). Conservation and dissipation of light energy as complementary processes: homoiohydric and poikilohydric autotrophs. Journal of Experimental Botany, 57, 1211–1223.CrossRefGoogle ScholarPubMed
Heibel, E., Lumbsch, H. T. and Schmitt, I. (1999). Genetic variation of Usnea filipendula (Parmeliaceae) populations in western Germany investigated by RAPDs suggest reinvasion from various sources. American Journal of Botany, 86, 753–757.CrossRefGoogle ScholarPubMed
Heiđmarsson, S., Mattson, J. E., Moberg, R., et al. (1997). Classification of lichen photomorphs. Taxon, 46, 519–520.CrossRefGoogle Scholar
Helms, G. (2003). Taxonomy and symbiosis in associations of Physciaceae and Trebouxia. Ph.D. thesis. Göttingen: Albrecht-von-Haller Institute, University of Göttingen.
Helms, G., Friedl, T., Rambold, G. and Mayrhofer, H. (2001). Identification of photobionts from the lichen family Physciaceae using algal-specific ITS rDNA sequencing. Lichenologist, 33, 73–86.CrossRefGoogle Scholar
Henderson, A. (1999). Lichen dyes. An historical perspective. Lees Museums and Galleries Review, 2, 30–34.Google Scholar
Henderson-Sellers, A. and Seaward, M. R. D. (1979). Monitoring lichen reinvasion of ameliorating environments. Environmental Pollution, 19, 207–213.CrossRefGoogle Scholar
Henriksson, E. and Pearson, L. C. (1981). Nitrogen fixation rate and chlorophyll content of the lichen Peltigera canina exposed to sulfur dioxide. American Journal of Botany, 68, 680–684.CrossRefGoogle Scholar
Henson, B. J., Hesselbrock, S. M., Watson, L. E., and Barnum, S. R. (2004). Molecular phylogeny of the heterocystous cyanobacteria (subsections IV and V) based on nifD. International Journal of Systematic and Evolutionary Microbiology, 54, 493–497.CrossRefGoogle ScholarPubMed
Henssen, A. (1963). Eine Revision der Flechtenfamilien Lichinaceae und Ephebaceae. Symbolae Bototanicae Upsala, 18, 1–123.Google Scholar
Henssen, A., in cooperation with Keuck, G., Renner, B. and Vobis, G. (1981). The lecanoralean centrum. In Ascomycete Systematics: The Lutrellian Concept, ed. Reynolds, D. R., pp. 138–234. New York: Springer.CrossRefGoogle Scholar
Henssen, A. (1986). The genus Paulia (Lichinaceae). Lichenologist, 18, 201–229.CrossRefGoogle Scholar
Henssen, A. (1995). The new lichen family Gloeoheppiaceae and its genera Gloeoheppia, Pseudopeltula and Gudella (Lichinales). Lichenologist, 27, 261–290.CrossRef
Henssen, A. and Jahns, H. M. (1973[1974]). Lichenes. Eine Einführung in die Flechtenkunde. Stuttgart: Thieme.Google Scholar
Henssen, A. and Tretiach, M. (1995). Paulia glomerata, a new epilithic species from Europe, and additional notes on some other Paulia species. Nova Hedwigia, 60, 297–309.Google Scholar
Henssen, A., Büdel, B. and Titze, A. (1987). Euopsis and Harpidium, genera of the Lichinaceae (Lichenes) with rostrate asci. Botanica Acta, 101, 49–55.Google Scholar
Henzler, T. and Steudle, E. (2000). Transport and metabolic degradation of hydrogen peroxide in Chara coralline: model calculations and measurements with the pressure probe suggest transport of H2O2 across water channels. Journal of Experimental Botany, 51, 2053–2066.CrossRefGoogle Scholar
Herrera-Campos, M. A., Lücking, R., Perez, R. E., et al. (2004). The foliicolous lichen flora of Mexico. V. Biogeographical affinities, altitudinal preferences, and an updated checklist of 293 species. Lichenologist, 36, 309–327.CrossRefGoogle Scholar
Hersoug, L. G. (1983). Lichen protein affinity towards walls of cultured and freshly isolated phycobionts and its relationship to cell wall cytochemistry. FEMS Microbiology Letters, 20, 417–420.CrossRefGoogle Scholar
Herzig, R. and Urech, M. (1991). Flechten als Bioindikatoren: Intergriertes biologisches Messsystem der Luftverschmutzung für das Schweizer Mittelland. Bibliotheca Lichenologica, 43, 1–283.Google Scholar
Hesbacher, S., Fröbery, L., Baur, A., Baur, B. and Proksch, P. (1996). Chemical variation within and between individuals of the lichenized ascomycete Tephromela atra. Biochemical Systematics and Ecology, 24, 603–609.CrossRefGoogle Scholar
Hestmark, G. (1990). Thalloconidia in the genus Umbilicaria. Nordic Journal of Botany, 9, 547–574.CrossRefGoogle Scholar
Hestmark, G. (1992). Sex, size competition and escape – strategies of reproduction and dispersal in Lasallia pustulata (Umbilicariaceae, Ascomycetes). Oecologia, 92, 305–312.CrossRefGoogle Scholar
Hestmark, G. (1997). Species diversity and reproductive strategies in the family Umbilicariaceae on high equatorial mountains – with remarks on global patterns. Bibliotheca Lichenologica, 68, 195–202.Google Scholar
Hestmark, G. (2000). The ecophysiology of lichen population biology. Bibliotheca Lichenologica, 75, 397–403.Google Scholar
Hestmark, G., Schroeter, B. and Kappen, L. (1997). Intrathalline and size-dependent patterns of activity in Lasallia pustulata and their possible consequences for competitive interactions. Functional Ecology, 11, 318–322.CrossRefGoogle Scholar
Heywood, V. H. (ed.) (1995). Global Biodiversity Assessment. Cambridge: Cambridge University Press.Google Scholar
Hildreth, K. C. and Ahmadjian, V. (1981). A study of Trebouxia and Pseudotrebouxia isolated from different lichens. Lichenologist, 13, 65–86.CrossRefGoogle Scholar
Hill, D. J. (1971). Experimental study of the effect of sulfite on lichens with reference to atmospheric pollution. New Phytologist, 70, 831–836.CrossRefGoogle Scholar
Hill, D. J. (1976). The physiology of lichen symbiosis. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 457–496. London: Academic Press.Google Scholar
Hill, D. J. (1985). Changes in photobiont dimensions and numbers during co-development of lichen symbionts. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 303–317. New York: Plenum Press.CrossRefGoogle Scholar
Hill, D. J. (1989). The control of the cell cycle in microbial symbionts. New Phytologist, 112, 175–184.CrossRefGoogle Scholar
Hill, D. J. and Smith, D. C. (1972). Lichen physiology XII. The “inhibition technique”. New Phytologist, 71, 15–30.CrossRefGoogle Scholar
Hilmo, O. (1994). Distribution and succession of epiphytic lichens on Picea abies in a boreal forest, central Norway. Lichenologist, 26, 149–169.CrossRefGoogle Scholar
Hilmo, O. and Holien, H. (2002). Epiphytic lichen response to the edge environment in a boreal Picea abies forest in central Norway. Bryologist, 105, 48–56.CrossRefGoogle Scholar
Hilmo, O. and Ott, S. (2002). Juvenile development of the cyanolichen Lobaria scrobiculata and the green algal lichen Platismatia glauca and Platismatia norvegica in a boreal Picea abies forest. Plant Biology, 4, 273–280.CrossRefGoogle Scholar
Hilmo, O. and Sastad, S. M. (2001). Colonization of old-forest lichens in a young and an old boreal Picea abies forest: an experimental approach. Biological Conservation, 102, 251–259.CrossRefGoogle Scholar
Hilmo, O., Holien, H. and Hytteborn, H. (2005). Logging strategy influences colonization of common chlorolichens on branches of Picea abies. Ecological Applications, 15, 983–996.CrossRefGoogle Scholar
Hirt, R. P., Logsdon, J., Doolittle, W. F. and Embley, T. M. (1999). Microsporidia are related to Fungi: evidence from the largest subunit of RNA polymerase II and other proteins. Proceedings of the National Academy of Sciences, USA, 96, 580–585.CrossRefGoogle ScholarPubMed
Hiserodt, R. D., Swijter, D. F. H. and Mussinan, C. J. (2000). Identification of atranorin and related potential allergens in oakmoss absolute by high-performance liquid chromatography-tandem mass spectrometry using negative ion atmospheric pressure chemical ionization. Journal of Chromatography A, 888, 103–111.CrossRefGoogle ScholarPubMed
Hitch, C. J. B. and Millbank, J. W. (1975). Nitrogen metabolism in lichens. VI. The blue-green phycobiont content, heterocyst frequency and nitrogenase activity in Peltigera species. New Phytologist, 74, 473–476.CrossRefGoogle Scholar
Hitch, C. J. B. and Millbank, J. W. (1976). Nitrogen metabolism in lichens. VII. Nitrogenase activity and heterocyst frequency in lichens with blue-green phycobionts. New Phytologist, 75, 239–244.CrossRefGoogle Scholar
Hocking, D., Kuchar, P., Plambeck, J. A. and Smith, R. A. (1978). The impact of gold smelter emissions on vegetation and soils of a sub-arctic forest–tundra transition ecosystem. Air Pollution Control Association Journal, 28, 133–137.CrossRefGoogle Scholar
Högnabba, F. (2006). Molecular phylogeny of the genus Stereocaulon (Stereocaulaceae, lichenized Ascomycetes). Mycological Research, 110, 1080–1092.CrossRefGoogle Scholar
Holopainen, T. (1983). Ultrastructural changes in epiphytic lichen Bryoria capillaris and Hypogymnia physodes in central Finland. Annales Botanici Fennici, 19, 39–52.Google Scholar
Holopainen, T. (1984). Types and distribution of ultrastructural symptoms in epiphytic lichens in several urban and industrial environments in Finland. Annales Botanici Fennici, 21, 219–229.Google Scholar
Holopainen, T. and Kärenlampi, L. (1984). Injuries to lichen ultrastructure caused by sulphur dioxide fumigations. New Phytologist, 98, 285–294.CrossRefGoogle Scholar
Holopainen, T. and Kärenlampi, L. (1985). Characteristic ultrastructural symptoms caused in lichens by experimental exposure to nitrogen compounds and fluorides. Annales Botanici Fennici, 22, 333–342.Google Scholar
Holopainen, T. and Kauppi, M. (1989). A comparison of light, fluorescence and electron microscopic observations in assessing the SO2 injury of lichens under different moisture conditions. Lichenologist, 21, 119–134.CrossRefGoogle Scholar
Holt, N. E., Tigmantas, D., Valkunas, L., et al. (2005). Carotenoid cation formation and the regulation of photosynthetic light harvesting. Science, 307, 433–436.CrossRefGoogle ScholarPubMed
Holub, S. M. and Lajtha, K. (2003). Mass loss and nitrogen dynamics during the decomposition of a 15N-labeled N2-fixing epiphytic lichen, Lobaria oregana. Canadian Journal of Botany, 81, 698–705.CrossRefGoogle Scholar
Holub, S. M. and Lajtha, K. (2004). The fate and retention of organic and inorganic 15N-nitrogen in an old-growth forest soil in Western Oregon. Ecosystems, 7, 368–380.CrossRefGoogle Scholar
Honegger, R. (1978 a). Ascocarpontogenie, Ascusstruktur und -funktion bei Vertretern der Gattung Rhizocarpon. Berichte der deutschen botanischen Gesellschaft, 91, 579–594.Google Scholar
Honegger, R. (1978 b). The ascus apex in lichenized fungi. I. The Lecanora-, Peltigera- and Teloschistes-types. Lichenologist, 10, 47–67.CrossRefGoogle Scholar
Honegger, R. (1980). The ascus apex in lichenized fungi. II. The Rhizocarpon-type. Lichenologist, 12, 157–172.CrossRefGoogle Scholar
Honegger, R. (1982 a). The ascus apex in lichenized fungi. III. The Pertusaria–type. Lichenologist, 14, 205–217.CrossRefGoogle Scholar
Honegger, R. (1982 b). Ascus structure and function, ascospore delimitation, and phycobiont cell wall types associated with the Lecanorales (lichenized ascomycetes). Journal of the Hattori Botanical Laboratory, 52, 417–429.Google Scholar
Honegger, R. (1984 a). Scanning electron microscopy of the contact site of conidia and trichogynes in Cladonia furcata. Lichenologist, 16, 11–19.CrossRefGoogle Scholar
Honegger, R. (1984 b). Ultrastructural studies on conidiomata, conidiophores, and conidiogenous cells in six lichen-forming Ascomycetes. Canadian Journal of Botany, 62, 2081–2093.CrossRefGoogle Scholar
Honegger, R. (1985). Ascus structure and ascospore formation in the lichen-forming Chaenotheca chrysocephala (Caliciales). Sydowia, Annales Mycologici Series II, 38, 146–157.Google Scholar
Honegger, R. (1986 a). Ultrastructural studies in lichens. I. Haustorial types and their frequencies in a range of lichens with trebouxioid phycobionts. New Phytologist, 103, 785–795.CrossRefGoogle Scholar
Honegger, R. (1986 b). Ultrastructural studies in lichens. II. Mycobiont and photobiont cell wall surface layers and adhering crystalline lichen products in four Parmeliaceae. New Phytologist, 103, 797–808.CrossRefGoogle Scholar
Honegger, R. (1991 a). Functional aspects of the lichen symbiosis. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 553–578.CrossRefGoogle Scholar
Honegger, R. (1991 b). Fungal evolution: Symbioses and morphogenesis. In Symbiosis, a Source of Evolutionary Innovation, ed. Margulis, L. and Fester, R., pp. 319–340. Cambridge: Massachusetts Institute of Technology Press.Google ScholarPubMed
Honegger, R. (1991 c). Haustoria-like structures and hydrophobic cell wall surface layers in lichens. In Electron Microscopy of Plant Pathogens, ed. Mendgen, K. and Lesemann, D. E., pp. 277–290. Berlin: Springer.CrossRefGoogle Scholar
Honegger, R. (1992). Lichens: mycobiont-photobiont relationships. In Algae and Symbioses. Plants, Animals, Fungi, Viruses, Interactions Explored, ed. Reisser, W., pp. 255–275. Bristol: Biopress.Google Scholar
Honegger, R. (1993). Developmental biology of lichens. New Phytologist, 125, 659–677.CrossRefGoogle Scholar
Honegger, R. (1995). Experimental studies with foliose macrolichens: fungal responses to spatial disturbance at the organismic level and to spatial problems at the cellular level. Canadian Journal of Botany, 73, 569–578.CrossRefGoogle Scholar
Honegger, R. (1997). Metabolic interactions at the mycobiont-photobiont interface in lichens. In Plant Relationships, The Mycota, Vol. V, Part A, ed. Carroll, G. C. and Tudzynski, P., pp. 209–221. Berlin: Springer.Google Scholar
Honegger, R. (1998). The lichen symbiosis – what is so spectacular about it?Lichenologist, 30, 193–212.CrossRefGoogle Scholar
Honegger, R. (2000). Great discoveries in bryology and lichenology – Simon Schwendener (1829–1919) and the Dual Hypothesis of Lichens. Bryologist, 103, 307–313.CrossRefGoogle Scholar
Honegger, R. (2001). The symbiotic phenotype of lichen-forming ascomycetes. In Fungal Associations, Vol. IX: The Mycota, ed. Hock, B., pp. 165–188. Berlin: Springer.CrossRefGoogle Scholar
Honegger, R. (2006). Water relations in lichens. In Fungi in the Environment, ed. Gadd, G. M., Watkinson, S. C. and Dyer, P., pp. 185–200. Cambridge: Cambridge University Press.Google Scholar
Honegger, R. and Bartnicki-Garcia, S. (1991). Cell wall structure and composition of cultured mycobionts from the lichen Cladonia macrophylla, Cladonia caespiticia, and Physcia stellaris (Lecanorales, Ascomycetes). Mycological Research, 95, 905–914.CrossRefGoogle Scholar
Honegger, R. and Haisch, A. (2001). Immunocytochemical location of the (1 → 3) (1 → 4)-beta-glucan lichenin in the lichen-forming ascomycete Cetraria islandica (Icelandic moss). New Phytologist, 150, 739–746.CrossRefGoogle Scholar
Honegger, R. and Zippler, U. (2007). Mating systems in representatives of the Parmeliaceae, Ramalinaceae and Physciaceae (Lecanoromycetes, lichen-forming ascomycetes). Mycological Research, 11, 424–432.CrossRefGoogle Scholar
Honegger, R., Peter, M. and Scherrer, S. (1996). Drought-stress induced structural alterations at the mycobiont-photobiont interface in a range of foliose macrolichens. Protoplasma, 190, 221–232.CrossRefGoogle Scholar
Honegger, R., Zippler, U., Gansner, H. and Scherrer, S. (2004). Mating systems in the genus Xanthoria (lichen-forming ascomycetes). Mycological Research, 108, 480–488.CrossRefGoogle Scholar
Hopwood, D. A. (1997). Genetic contribution to understanding polyketide synthases. Chemical Reviews, 97, 2465–2498.CrossRefGoogle Scholar
Horstmann, J. L., Denison, W. C. and Silvester, W. B. (1982). 15N2 fixation and molybdenum enhancement of acetylene reduction by Lobaria spp. New Phytologist, 92, 235–241.CrossRefGoogle Scholar
Houdijk, A. L. F. M. and Roelofs, J. G. M. (1991). Deposition of acidifying and eutrophicating substances in Dutch forest. Acta Botanica Neerlandica, 40, 245–255.CrossRefGoogle Scholar
Huebert, D. B., L'Hirondelle, S. J. and Addison, P. A. (1985). The effects of sulphur dioxide on net CO2 assimilation in the lichen Evernia mesomorpha Nyl. New Phytologist, 100, 643–651.CrossRefGoogle Scholar
Huelsenbeck, J. P. and Ronquist, F. (2001). Bayesian inference of phylogenetic trees. Bioinformatics, 17, 754–755.CrossRefGoogle ScholarPubMed
Humphries, C. J. and Ebach, M. C. (2004). Biogeography on a dynamic earth. In Frontiers of Biogeography: New Directions in the Geography of Nature, ed. Lomolino, M. V. and Heaney, L. R., pp. 67–86. Sunderland: Sinauer Associates.Google Scholar
Humphries, C. J. and Parenti, L. R. (1999). Cladistic Biogeography: Interpreting Patterns of Plant and Animal Distributions. 2nd edn, Oxford Biogeography Series 12. Oxford: Oxford University Press.Google Scholar
Humphries, C. J., Williams, P. H. and Vane-Wright, R. I. (1995). Measuring biodiversity value for conservation. Annual Reviews of Ecology and Systematics, 26, 93–111.CrossRefGoogle Scholar
Huneck, S. (1999). The significance of lichens and their metabolites. Die Naturwissenschaften, 86, 559–570.CrossRefGoogle ScholarPubMed
Huneck, S. (2001). New results on the chemistry of lichen substances. In Progress in the Chemistry of Organic Products, ed. Herz, W., Falk, H., Kirby, G. W. and Moore, R. E., pp. 1–276. New York: Springer.Google Scholar
Huneck, S. and Schreiber, K. (1972). Wachstumsregulatorische Eigenschaften von Flechten- und Moos-Inhaltstoffen. Phytochemistry, 11, 2429–2434.CrossRefGoogle Scholar
Huneck, S. and Yoshimura, I. (1996). Identification of Lichen Substances. Springer: Berlin.CrossRefGoogle Scholar
Huneck, S., Bothe, H.-K. and Richter, W. (1990). Über den Metallgehalt von Flechten von Kupferschieferhalden der Umgebung von Mansfeld. Herzogia, 8, 295–304.Google Scholar
Huovinen, K., Hiltunen, R. and Schantz, M. (1985). A high performance liquid chromatographic method for the analyses of lichen compounds from the genera Cladina and Cladonia. Acta Pharmaceutica Fennica, 94, 99–112.Google Scholar
Huss, V. A. R. and Sogin, M. L. (1990). Phylogenetic position of some Chlorella species within the Chlorococcales based upon complete small-subunit ribosomal RNA sequences. Journal of Molecular Evolution, 31, 432–442.CrossRefGoogle ScholarPubMed
Huss-Danell, K. (1977). Nitrogen fixation by Stereocaulon paschale under field conditions. Canadian Journal of Botany, 55, 585–592.CrossRefGoogle Scholar
Huss-Danell, K. (1978). Seasonal variation in the capacity for nitrogenase activity in the lichen Stereocaulon paschale. New Phytologist, 81, 89–98.CrossRefGoogle Scholar
Huss-Danell, K. (1979). The cephalodia and their nitrogenase activity in the lichen Stereocaulon paschale. Zeitschrift für Pflanzenphysiologie, 95, 431–440.CrossRefGoogle Scholar
Hyvärinen, M. and Crittenden, P. D. (1998 a). Relationships between atmospheric nitrogen inputs and the vertical nitrogen and phosphorus concentration gradients in the lichen Cladonia portentosa. New Phytologist, 140, 519–530.CrossRefGoogle Scholar
Hyvärinen, M. and Crittenden, P. D. (1998 b). Growth of the cushion-forming lichen, Cladonia portentosa, at nitrogen-polluted and unpolluted heathland sites. Environmental and Experimental Botany, 40, 67–76.CrossRefGoogle Scholar
Hyvärinen, M. and Crittenden, P. D. (2000). 33P translocation in the thallus of the mat-forming lichen Cladonia portentosa. New Phytologist, 145, 281–288.CrossRefGoogle Scholar
Hyvärinen, M., Härdling, R. and Tuomi, J. (2002). Cyanobacterial lichen symbiosis: the fungal partner as an optimal harvester. Oikos, 98, 498–504.CrossRefGoogle Scholar
Hyvärinen, M., Roitto, M., Ohtonen, R. and Markkola, A. (2000). Impact of wet deposited nickel on the cation content of a mat-forming lichen Cladina stellaris. Environmental and Experimental Botany, 43, 211–218.CrossRefGoogle ScholarPubMed
Hyvärinen, M., Walter, B. and Koopmann, R. (2003). Impact of fertilization on phenol content and growth rate of Cladina stellaris: a test of the carbon-nutrient balance hypothesis. Oecologia, 134, 176–181.CrossRefGoogle ScholarPubMed
Ihda, T. A., Nakano, T., Yoshimura, I. and Iwatsuki, Z. (1993). Phycobionts isolated from Japanese species of Anzia (lichenes). Archiv für Protistenkunde, 143, 163–172.CrossRefGoogle Scholar
Ihlen, P. G. and Ekman, S. (2002). Outline of phylogeny and character evolution in Rhizocarpon (Rhizocarpaceae, lichenized Ascomycota) based on nuclear ITS and mitochondrial SSU ribosomal DNA sequences. Biological Journal of the Linnean Society, London, 77, 535–546.CrossRefGoogle Scholar
Ingólfsdóttir, K. (2002). Molecules of interest: usnic acid. Phytochemistry, 61, 729–736.CrossRefGoogle Scholar
Innes, J. L. (1983). Lichenometric dating of debris-flow deposits in the Scottish Highlands. Earth Surface Processes and Landforms, 8, 579–588.CrossRefGoogle Scholar
Innes, J. L. (1985). Lichenometry. Progress in Physical Geography, 9, 187–254.CrossRefGoogle Scholar
Innes, J. L. (1988). The use of lichens in dating. In Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 75–91. Boca Raton: CRC Press.Google Scholar
Insarov, G. and Schroeter, B. (2002). Lichen monitoring and climate change. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 183–201. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Jaag, O. and Thomas, E. (1934). Neue Untersuchungen über die Flechte Epigloea bactrospora Zukal. Berichte der schweierischen botanischen Gesellschaft, 34, 77–89.Google Scholar
Jahns, H. M. (1970). Untersuchungen zur Entwicklungsgeschichte der Cladoniaceen, unter besonderer Berücksichtigung des Podetien-Problems. Nova Hedwigia, 20, 1–177.Google Scholar
Jahns, H. M. (1984). Morphology, reproduction and water relations – a system of morphogenetic interactions in Parmelia saxatilis. Nova Hedwigia, 79, 715–37.Google Scholar
Jahns, H. M. (1987). New trends in developmental morphology of the thallus. Bibliotheca Lichenologica, 25, 17–33.Google Scholar
Jahns, H. M. (1988). The lichen thallus. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 95–143. Boca Raton: CRC Press.Google Scholar
Jahns, H. M., Tuiz-Dubiel, A. and Blank, L. (1976). Hygroskopische Bewegungen der Sorale von Hypogymnia physodes. Herzogia, 4, 15–23.Google Scholar
James, P. W. and Henssen, A. (1976). The morphological and taxonomic significance of cephalodia. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 27–77. London: Academic Press.Google Scholar
James, P. W., Hawksworth, D. L. and Rose, F. (1977). Lichen communities in the British Isles: a preliminary conspectus. In Lichen Ecology, ed. Seaward, M. R. D., pp. 295–413. London: Academic Press.Google Scholar
Jancey, R. (1966). The application of numerical methods of data analysis to the genus Phyllota (Benth.) in New South Wales. Australian Journal of Botany, 14, 131–149.CrossRefGoogle Scholar
Janex-Favre, M. C. and Ghaleb, M. I. (1986). L'ontogenie et la structure des apothecies du Xanthoria parietina (L.) Beltr. (discolichen). Cryptogamie, Bryologie et Lichenologie, 7, 457–478.Google Scholar
Jansova, I. and Soldan, Z. (2006). The habitat factors that affect the composition of bryophyte and lichen communities on fallen logs. Preslia, 78, 67–86.Google Scholar
Jarmuszkiewicz, W. (2001). Uncoupling proteins in mitochondria of plants and some microorganisms. Acta Biochimica Polonica, 48, 145–155.Google ScholarPubMed
Jennings, D. (1995). The Physiology of Fungal Nutrition. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jensen, M. (2002). Measurement of chlorophyll fluorescence in lichens. In Protocols in Lichenology, ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 135–151. Berlin: Springer.CrossRefGoogle Scholar
Jensen, M. and Kricke, R. (2002). Chlorophyll fluorescence measurements in the field: assessment of the vitality of large numbers of lichen thalli. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 327–332. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Jensen, M., Linke, K., Dickhäuser, A. and Feige, G. B. (1999). The effect of agronomic photosystem-II herbicides on lichens. Lichenologist, 31, 95–103.Google Scholar
John, D. M., Whitton, B. A. and Brook, A. J. (2002). The Freshwater Algal Flora of the British Isles. Cambridge: Cambridge University Press.Google Scholar
John, E. and Dale, M. R. T. (1991). Determinants of spatial pattern in saxicolous lichen communities. Lichenologist, 23, 227–236.CrossRefGoogle Scholar
John, V. (1996). Preliminary catalogue of lichenized and lichenicolous fungi of Mediterranean Turkey. Bocconea, 6, 173–216.Google Scholar
Johnson, L. R. and John, D. M. (1990). Observations on Dilabifilium (class Chlorophyta, order Chaetophorales sensu strictu) and allied genera. British Phycological Journal, 25, 53–61.CrossRefGoogle Scholar
Johnson, P. N. and Galloway, D. J. (2002). Lichens and Their Conservation Needs in New Zealand. Landcare Research Contract Report: LC0102/132. Dunedin: Landcare Research.Google Scholar
Johnston, J. (2001). Siphulella. Flora of Australia, 58A, 22–23.
Jones, D. (1988). Lichens and pedogenesis. In CRC Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 109–124. Boca Raton: CRC Press.Google Scholar
Jones, D., Wilson, M. J. and Laundon, J. R. (1982). Observations on the location and form of lead in Stereocaulon vesuvianum. Lichenologist, 14, 281–286.CrossRefGoogle Scholar
Jørgensen, P. M. (1996). The oceanic element in the Scandinavian lichen flora revisited. Symbolae Botanicae Upsalienses, 31, 297–317.Google Scholar
Jørgensen, P. M. (1997). Further notes on hairy Leptogium species. Symbolae Botanicae Upsalienses, 32, 113–130.Google Scholar
Jørgensen, P. M. (1998). What shall we do with the blue-green counterparts?Lichenologist, 30, 351–356.CrossRefGoogle Scholar
Jørgensen, P. M. (2000 a). On the sorediate counterparts of the lichen Fuscopannaria leucosticta. Bryologist, 103, 104–107.CrossRefGoogle Scholar
Jørgensen, P. M. (2000 b). New or interesting Parmeliella species from the Andes and Central America. Lichenologist, 32, 139–147.CrossRefGoogle Scholar
Jørgensen, P. M. (2001). The lichen genus Erioderma (Pannariaceae) in China and Japan. Annales Botanici Fennici, 38, 259–264.Google Scholar
Jørgensen, P. M. (2005). A new Atlantic species in Fuscopannaria, with a key to its European species. Lichenologist, 37, 221–225.CrossRefGoogle Scholar
Jørgensen, P. M. and Arvidsson, L. (2002). The lichen genus Erioderma (Pannariaceae) in Ecuador and neighbouring countries. Nordic Journal of Botany, 22, 87–114.CrossRefGoogle Scholar
Jørgensen, P. M. and Jahns, H. M. (1987). Muhria, a remarkable new lichen genus from Scandinavia. Notes of the Royal Botanical Garden Edinburgh, 44, 581–599.Google Scholar
Jovan, S. and Carlberg, T. (2007). Nitrogen content of Letharia vulpina tissue from forests of the Sierra Nevada, California: geographic patterns and relationships to ammonia estimates and climate. Environmental Monitoring and Assessment, 129, 243–251.CrossRefGoogle ScholarPubMed
Jovan, S. and McCune, B. (2004). Regional variation in epiphytic macrolichen communities in northern and central California forests. Bryologist, 107, 328–339.CrossRefGoogle Scholar
Jovan, S. and McCune, B. (2005). Air-quality bioindication in the greater Central Valley of California, with epiphytic macrolichen communities. Ecological Applications, 15, 1712–1726.CrossRefGoogle Scholar
Jovan, S. and McCune, B. (2006). Using epiphytic macrolichen communities for biomonitoring ammonia in forests of the Greater Sierra Nevada, California. Water, Air, and Soil Pollution, 170, 69–93.CrossRefGoogle Scholar
Kaasalainen, S. and Rautiainen, M. (2005). Hot spot reflectance signatures of common boreal lichens. Journal of Geophysical Research, 110, 15–24.CrossRefGoogle Scholar
Kaiser, M. A. and Debbrecht, F. J. (1977). Qualitative and quantitative analysis of gas chromatography. In Modern Practice of Gas Chromatography, ed. Grob, R., pp. 151–211. New York: John Wiley.Google Scholar
Kalb, K. (1987). Brasilianische Flechten. 1. Die Gattung Pyxine. Bibliotheca Lichenologica, 24, 1–89.Google Scholar
Kallio, P. (1974). Nitrogen fixation in subarctic lichens. Oikos, 25, 194–198.CrossRefGoogle Scholar
Kallio, P., Suhonen, S. and Kallio, S. (1972). The ecology of nitrogen fixation in Nephroma arcticum and Solorina crocea. Reports from the Kevo Subarctic Research Station, 9, 7–14.Google Scholar
Kallio, S. (1978). On the effect of forest fertilizers on nitrogenase activity in two subarctic lichens. In Environmental Role of Nitrogen-fixing Blue-green Algae and Asymbiotic Bacteria, ed. Granhall, U., pp. 217–224. Stockholm: Swedish Natural Science Research Council.Google Scholar
Kandler, O. (1987). Lichen and conifer recolonization in Munich's cleaner air. In Symposium of the Commission of the European Communities on “Effects of Air Pollution on Terrestrial and Aquatic Ecosystems” 18–22 May 1987, ed. Mathy, P., pp. 1–7. Brussels: European Commission.Google Scholar
Kantvilas, G. (2000). Conservation of Tasmanian lichens. Mitteilungen der Eidgenössischen Forschungsanstalt für Wald, Schnee und Landschaft, 75, 357–367.Google Scholar
Kantvilas, G. (2002). Studies on the lichen genus Siphula Fr. Bibliotheca Lichenologica, 82, 37–53.Google Scholar
Kantvilas, G. (2004). Steinia. Flora of Australia 56A, 1–3.
Kantvilas, G. and Jarman, S. J. (1999). Lichens of rainforest in Tasmania and south-eastern Australia. Flora of Australia Supplementary Series, 9, 1–212.Google Scholar
Kantvilas, G. and Jarman, S. J. (2006). Recovery of lichens after logging: preliminary results from Tasmania's wet forests. Lichenologist, 38, 383–394.CrossRefGoogle Scholar
Kantvilas, G. and McCarthy, P. M. (2004). Hueidea. Flora of Australia 56A, 182–183.
Kantvilas, G., Elix, J. A. and Jarman, S. J. (2002). Tasmanian lichens, identification, distribution and conservation status. I. Parmeliaceae. Flora of Australia Supplementary Series 15, 1–274.Google Scholar
Kappen, L. (1974). Response to extreme environments. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 311–380. New York: Academic Press.Google Scholar
Kappen, L. (1985). Vegetation and ecology of ice-free areas of northern Victoria Land, Antarctica. Polar Biology, 4, 213–225.CrossRefGoogle Scholar
Kappen, L. (1988). Ecophysiological relationships in different climatic regions. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 37–99. Boca Raton: CRC Press.Google Scholar
Kappen, L. (1993). Lichens in the antarctic region. In Antarctic Microbiology, ed. Friedmann, E. I., pp. 433–490. New York: Wiley-Liss.Google Scholar
Kappen, L. (2000). Some aspects of the great success of lichens in Antarctica. Antarctic Science, 12, 314–324.CrossRefGoogle Scholar
Kappen, L. (2004). The diversity of lichens in Antarctica, a review and comments. Bibliotheca Lichenologica, 88, 331–343.Google Scholar
Kappen, L. and Valladares, F. (1999). Opportunistic growth and desiccation tolerance: the ecological success of poikilohydrous autotrophs. In Handbook of Functional Plant Ecology, ed. Pugnaire, F. I. and Valldares, F., pp. 121–194. New York: Marcel Dekker.Google Scholar
Kappen, L., Breuer, M. and Bölter, M. (1991). Ecological and physiological investigations in continental Antarctic cryptogams. 3. Photosynthetic production of Usnea sphacelata: diurnal courses, models, and the effect of photoinhibition. Polar Biology, 11, 393–401.CrossRefGoogle Scholar
Kappen, L., Schroeter, B., Green, T. G. A. and Seppelt, R. D. (1998 a). Chlorophyll a fluorescence and CO2 exchange on Umbilicaria aprina under extreme light stress in the cold. Oecologia, 113, 325–331.CrossRefGoogle ScholarPubMed
Kappen, L., Schroeter, B., Green, T. G. A. and Seppelt, R. D. (1998 b). Microclimatic conditions, meltwater moistening, and the distributional pattern of Buellia frigida on rock in a southern continental Antarctic habitat. Polar Biology, 19, 101–106.CrossRefGoogle Scholar
Kappen, L., Sommerkorn, M. and Schroeter, B. (1995). Carbon acquisition and water relations of lichens in polar regions – potentials and limitations. Lichenologist, 27, 531–545.Google Scholar
Kardish, N., Silberstein, L., Fleminger, G. and Galun, M. (1991). Lectin from the lichen Nephroma laevigatum Ach. Localization and function. Symbiosis, 11, 47–62.Google Scholar
Kärnefelt, I. (1989). Morphology and phylogeny in the Teloschistales. Cryptogamic Botany, 1, 147–203.Google Scholar
Kärnefelt, I. (1990). Evidence of a slow evolutionary change in the speciation of lichens. Bibliotheca Lichenologica, 38, 291–306.Google Scholar
Kärnefelt, E. I. and Thell, A. (1995). Genotypical variation and reproduction in natural populations of Thamnolia. Bibliotheca Lichenologica, 58, 213–234.Google Scholar
Karnieli, A., Kokaly, R., West, N. E. and Clark, R. N. (2001). Remote sensing of biological soil crusts. In Biological Soil Crusts: Structure, Function and Management, ed. Belnap, J. and Lange, O. L., pp. 431–455. Berlin: Springer.Google Scholar
Kauff, F. and Büdel, B. (2005). Ascoma ontogeny and apothecial anatomy in the Gyalectaceae (Ostropales, Ascomycota) support the re-establishment of the Coenogoniaceae. Bryologist, 108, 272–281.CrossRefGoogle Scholar
Kauff, F. and Lutzoni, F. (2002). Phylogeny of the Gyalectales and Ostropales (Ascomycota, Fungi): among and within order relationships based on nuclear ribosomal RNA small and large subunits. Molecular Phylogenetics and Evolution, 25, 138–156.CrossRefGoogle ScholarPubMed
Keller, N. P. and Hohn, T. M. (1997). Metabolic pathway gene clusters in filamentous fungi. Fungal Genetics and Biology, 21, 17–21.CrossRefGoogle ScholarPubMed
Kelly, B. B. and Becker, V. E. (1975). Effects of light intensity and temperature on nitrogen fixation by Lobaria pulmonaria, Sticta weigelii, Leptogium cyanescens and Collema subfurvum. Bryologist, 78, 350–355.CrossRefGoogle Scholar
Kelly, B. C. and Gobas, F. A. P. C. (2001). Bioaccumulation of persistent organic pollutants in lichen-caribou-wolf food chains of Canada's central and western Arctic. Environmental Science and Technology, 35, 325–334.CrossRefGoogle ScholarPubMed
Kerfin, W. and Boger, P. (1982). Light-induced hydrogen evolution by blue-green algae (Cyanobacteria). Physiologia Plantarum, 54, 93–98.CrossRefGoogle Scholar
Kershaw, K. A. (1974). Dependence of the level of nitrogenase activity on the water content of the thallus in Peltigera canina, P. evansiana, P. polydactyla, and P. praetextata. Canadian Journal of Botany, 52, 1423–1427.CrossRefGoogle Scholar
Kershaw, K. A. (1977). Physiological-environmental interactions in lichens. II. The pattern of net photosynthetic acclimation of Peltigera canina (L.) Willd. var. praetextata (Floreke in Somm.) Hue, and P. polydactyla (Neck.) Hoff. New Phytologist, 79, 377–390.CrossRefGoogle Scholar
Kershaw, K. A. (1985). Physiological Ecology of Lichens. Cambridge: Cambridge University Press.Google Scholar
Kershaw, K. A. and Larson, D. W. (1974). Studies on lichen-dominated systems. IX. Topographic influences on microclimate and species distribution. Canadian Journal of Botany, 52, 1935–1945.CrossRefGoogle Scholar
Kershaw, K. A. and Looney, J. H. H. (1985). Quantitative and Dynamic Plant Ecology, 3rd edn. London: Edward Arnold.Google Scholar
Kershaw, K. A. and Rouse, W. R. (1971). Studies on lichen dominated ecosystems. I. The water relation of Cladonia alpestris in spruce-lichen woodland in northern Ontario. Canadian Journal of Botany, 49, 1389–1399.CrossRefGoogle Scholar
Kershaw, M. J. and Talbot, N. J. (1998). Hydrophobins and repellents: proteins with fundamental roles in fungal morphogenesis. Fungal Genetics and Biology, 23, 18–33.CrossRefGoogle ScholarPubMed
Kershaw, M. J., Thornton, C., Wakley, G. and Talbot, N. J. (2005). Four conserved intramolecular disulphide linkages are required for secretion and cell wall localization of a hydrophobin during fungal morphogenesis. Molecular Microbiology, 56, 117–125.CrossRefGoogle ScholarPubMed
Kets, E., Galinski, E., Wit, M., Bont, J. and Heipieper, H. (1996). Mannitol, a novel bacterial compatible solute in Pseudomonas putida S12. Journal of Bacteriology, 178, 6665–6670.CrossRefGoogle ScholarPubMed
Kidd, D. M. and Ritchie, M. G. (2006). Phylogeographic information systems: putting the geography into phylogeography. Journal of Biogeography, 33, 1851–1865.CrossRefGoogle Scholar
Kieft, T. L. and Ruscetti, T. (1990). Characterization of biological ice nuclei from a lichen. Journal of Bacteriology, 172, 3519–3523.CrossRefGoogle ScholarPubMed
Kinoshita, Y. (1993). The Production of Lichen Substances for Pharmaceutical Use by Lichen Tissue Culture. Osaka: Nippon Paint Publication.Google Scholar
Kirk, P. M., Cannon, P. F., David, J. C. and Stalpers, J. A. (2001). Ainsworth and Bisby's Dictionary of the Fungi. 9th edn. Wallingford, UK: CAB International.Google Scholar
Kirkpatrick, R. C., Long, Y. C., Zhong, T. and Xia, L. (1998). Social organization and range use in the Yunnan snub-nosed monkey Rhinopithecus bieti. International Journal of Primatology, 19, 13–51.CrossRefGoogle Scholar
Knops, J. M.H, Nash, T. H. III, Boucher, V. L. and Schlesinger, W. L. (1991). Mineral cycling and epiphytic lichens: implications at the ecosystem level. Lichenologist, 23, 309–321.CrossRefGoogle Scholar
Knops, J. M. H., Nash, T. H. III and Schlesinger, W. H. (1996). The influence of epiphytic lichens on the nutrient cycling of an oak woodland. Ecological Monographs, 66, 159–179.CrossRefGoogle Scholar
Knowles, R. D., Pastor, J. and Biesboer, D. D. (2006). Increased soil nitrogen associated with dinitrogen-fixing, terricolous lichens of the genus Peltigera in northern Minnesota. Oikos, 114, 37–48.CrossRefGoogle Scholar
Koch, J. and Kilian, R. (2005). ‘Little Ice Age’ glacier fluctuations, Gran Campo Nevado, southernmost Chile. Holocene, 15, 20–28.CrossRefGoogle Scholar
Köck, M., Schlee, D. and Metzger, U. (1985). Sulfite-induced changes of oxygen metabolism in the action of superoxide dismutase in Euglena gracilis and Trebouxia sp. Biochemie und Physiologie der Pflanzen, 180, 213–224.CrossRefGoogle Scholar
Koga, S., Echigo, A. and Nunomura, K. (1966). Physical properties of cell water in partially dried Saccharomyces cerevisiae. Biophysics Journal, 6, 665–674.CrossRefGoogle ScholarPubMed
Komárek, J. and Anagnostidis, K. (1998). Cyanoprokaryota. 1. Teil: Chroococcales. Jena: Gustav Fischer.Google Scholar
Komárek, J. and Anagnostidis, K. (2005). Cyanoprokaryota. 2. Teil: Oscillatoriales. München: Elsevier.Google Scholar
Kondratyuk, S. and Kärnefelt, I. (1997). Josefpoeltia and Xanthomendoxa, two new genera in the Teloschistaceae (lichenized Ascomycotina). Bibliotheca Lichenologica, 68, 19–44.Google Scholar
Kong, F. X., Hu, W., Chao, S. Y., Sang, W. L. and Wang, L. S. (1999). Physiological responses of the lichen Xanthoparmelia mexicana to oxidative stress of SO2. Environmental and Experimental Botany, 42, 201–209.CrossRefGoogle Scholar
Kopecky, J., Azarkovich, M., Pfündel, E. E., Shuvalov, V. A. and Heber, U. (2005). Thermal dissipation of light energy is regulated differently and by different mechanisms in lichens and higher plants. Plant Biology, 7, 156–167.CrossRefGoogle ScholarPubMed
Koptsik, S. V., Koptsik, G. N. and Meryashkina, L. V. (2004). Ordination of plant communities in forest biogeocenoses under conditions of air pollution in the northern Kola Peninsula. Russian Journal of Ecology, 35, 190–199.CrossRefGoogle Scholar
Korf, R. P. (1973). Discomycetes and Tuberales. In The Fungi. Vol. IVA: A Taxonomic Review with Keys: Ascomycetes and Fungi Imperfecti, ed. Ainsworth, G. C., Sparrow, F. K. and Sussman, A. S., pp. 249–319. New York: Academic Press.Google Scholar
Kostner, B. and Lange, O. L. (1986). Epiphytische Flechten in bayerischen Waldschadensgebieten des nordlichen Alpenraumes: Floristisch-soziologische Untersuchungen und Vitalitätstests durch Photosynthesemessungen. Berichte der Akademie fur Naturschutz und Landschaftspflege, 10, 185–210.Google Scholar
Kranner, I. and Birtić, S. (2005). A modulating role for antioxidants in desiccation tolerance. Integrative and Comparative Biology, 45, 734–740.CrossRefGoogle ScholarPubMed
Kranner, I. and Grill, D. (1994). Rapid changes of the glutathione status and the enzymes involved in the reduction of glutathione-disulfide during the initial stage of wetting of lichens. Cryptogamic Botany, 4, 203–206.Google Scholar
Kranner, I. and Lutzoni, F. (1999). Evolutionary consequences of transition to a lichen symbiotic state and physiological adaptation to oxidative damage associated with poikilohydry. In Plant Response to Environmental Stress: From Phytohormones to Genome Reorganisation, ed. Lerner, H. R., pp. 591–628. New York: Marcel Dekker.Google Scholar
Kranner, I., Beckett, R., Hochman, A. & Nash, T. H. III (2008). Desiccation tolerance in lichens: a review. Bryologist, 111 (in press).CrossRefGoogle Scholar
Kranner, I., Cram, W. J., Zorn, M., et al. (2005). Antioxidants and photoprotection in a lichen as compared with its isolated symbiotic partners. Proceedings of the National Academy of Sciences, USA, 102, 3141–3146.CrossRefGoogle Scholar
Kranner, I., Zorn, M., Turk, B., et al. (2003). Biochemical traits of lichens differing in relative desiccation tolerance. New Phytologist, 160, 167–176.CrossRefGoogle Scholar
Kroken, S. and Taylor, J. W. (2000). Phylogenetic species, reproductive mode, and specificity of the green alga Trebouxia forming lichens with the fungal genus Letharia. Bryologist, 103, 645–660.CrossRefGoogle Scholar
Kroken, S. and Taylor, J. W. (2001). A gene geneology approach to recognize phylogenetic species boundaries in the lichenized fungus Letharia. Mycologia, 93, 38–53.CrossRefGoogle Scholar
Kroken, S., Glass, N. L., Taylor, J. W., Yoder, O. C. and Turgeon, B. G. (2003). Phylogenomic analysis of type I polyketide synthase genes in pathogenic and saprobic ascomycetes. Proceedings of the National Academy of Sciences, USA, 100, 15 670–15 675.CrossRefGoogle ScholarPubMed
Kurina, L. M. and Vitousek, P. M. (1999). Controls over the accumulation and decline of a nitrogen-fixing lichen, Stereocaulon vulcani on young Hawaiian lava flows. Journal of Ecology, 87, 784–799.CrossRefGoogle Scholar
Kurina, L. M. and Vitousek, P. M. (2001). Nitrogen fixation rates of Stereocaulon vulcani on young Hawaiian lava flows. Biogeochemistry, 55, 179–194.CrossRefGoogle Scholar
Kytöviita, M. M. and Crittenden, P. D. (1994). Effects of simulated acid rain on nitrogenase activity (acetylene reduction) in the lichen Stereocaulon paschale (L.) Hoffm., with special reference to nutritional aspects. New Phytologist, 128, 263–271.CrossRefGoogle Scholar
Kytöviita, M. M. and Crittenden, P. D. (2002). Seasonal variation in growth rate in Stereocaulon paschale. Lichenologist, 34, 533–537.CrossRefGoogle Scholar
Laaksovirta, K. and Olkkonen, H. (1979). Effect of air pollution on epiphytic lichen vegetation and element contents of a lichen and pine needles at Valkeakoski, S. Finland. Annales Botanici Fennici, 16, 285–296.Google Scholar
Laaksovirta, K., Olkkonen, H. and Alakijala, P. (1976). Observations on the lead content of lichen and bark adjacent to a highway in southern Finland. Environmental Pollution, 11, 247–255.CrossRefGoogle Scholar
Lakatos, M. (2002). Ökologische Untersuchungen wuchsformbedingter Verbreitungsmuster von Flechten im tropischen Regenwald. Ph.D. thesis. Kaiserslautern: University of Kaiserslautern.
Lakatos, M., Rascher, U. and Büdel, B. (2006). Functional characteristics of corticolous lichens in the understory of a tropical lowland rain forest. New Phytologist, 172, 679–695.CrossRefGoogle ScholarPubMed
Lambers, H. (1985). Respiration in intact plants and tissues: its regulation and dependence on environmental factors, metabolism and invaded organisms. In Higher Plant Respiration, ed. Douce, R. and Day, D. A., pp. 418–465. Berlin: Springer.CrossRefGoogle Scholar
Lambers, H., Chapin, F. S., III and Pons, T. L. (1998). Photosynthesis, respiration, and long distance transport. In Plant Physiological Ecology, ed. Lambers, H., Chapin, F. S. III and Pons, T. L., pp. 10–95. Berlin: Springer.CrossRefGoogle Scholar
Lang, G. E., Reiners, W. A. and Heier, R. K. (1976). Potential alteration of precipitation chemistry by epiphytic lichens. Oecologia, 25, 229–241.CrossRefGoogle ScholarPubMed
Lang, G. E., Reiners, W. A. and Pike, L. H. (1980). Structure and biomass of epiphytic lichen communities of balsam fir forests in New Hampshire. Ecology, 61, 541–550.CrossRefGoogle Scholar
Lange, O. L. (1953). Hitze- und Trockenresistenz der Flechten in Beziehung zu ihrer Verbreitung. Flora, 140, 39–97.Google Scholar
Lange, O. L. (1965). Der CO2-Gaswechsel von Flechten bei tiefen Temperaturen. Planta, 64, 1–19.CrossRefGoogle Scholar
Lange, O. L. (1969). Experimentell-ökologische Untersuchungen an Flechten der Negev-Wüste. I. CO2-Gaswechsel von Ramalina maciformis (Del.) Bory unter kontrollierten Bedingungen im Laboratorium. Flora, 158, 324–359.Google Scholar
Lange, O. L. (1980). Moisture content and CO2 exchange of lichens. I. Influence of temperature on moisture-dependent net photosynthesis and dark respiration in Ramalina maciformis. Oecologia, 45, 82–87.CrossRefGoogle ScholarPubMed
Lange, O. L. (2002). Photosynthetic productivity of the epilithic lichen Lecanora muralis: long-term field monitoring of CO2 exchange and its physiological interpretation. I. Dependence of photosynthesis on water content, light, temperature, and CO2 concentration from laboratory measurements. Flora, 197, 233–249.CrossRefGoogle Scholar
Lange, O. L. (2003 a). Photosynthetic productivity of the epilithic lichen Lecanora muralis: long-term field monitoring of CO2 exchange and its physiological interpretation. II. Diel and seasonal patterns of net photosynthesis and respiration. Flora, 198, 55–70.Google Scholar
Lange, O. L. (2003 b). Photosynthetic productivity of the epilithic lichen Lecanora muralis: long-term field monitoring of CO2 exchange and its physiological interpretation. III. Diel, seasonal, and annual carbon budgets. Flora, 198, 277–292.Google Scholar
Lange, O. L. and Bertsch, A. (1965). Photosynthese der Wüstenflechte Ramalina maciformis nach Wasserdampfaufnahme aus dem Luftraum. Naturwissenschaften, 52, 215–216.CrossRefGoogle Scholar
Lange, O. L. and Green, T. G. A. (2003). Photosynthetic performance of a foliose lichen of biological soil crust communities: long-term monitoring of the CO2 exchange of Cladonia convoluta under temperate habitat conditions. Bibliotheca Lichenologica, 86, 257–280.Google Scholar
Lange, O. L. and Green, T. G. A. (2005). Lichens show that fungi can acclimate their respiration to seasonal changes in temperature. Oecologia, 142, 11–19.CrossRefGoogle ScholarPubMed
Lange, O. L. and Green, T. G. A. (2006). Nocturnal respiration in lichens in their natural habitat is not affected by preceding diurnal net photosynthesis. Oecologia, 148, 396–404.CrossRefGoogle Scholar
Lange, O. L. and Metzner, H. (1965). Lichtabhängiger Kohlenstoff-Einbau in Flechten bei tiefen Temperaturen. Naturwissenschaften, 52, 191.CrossRefGoogle Scholar
Lange, O. L. and Tenhunen, J. D. (1981). Moisture content and CO2 exchange of lichens. II. Depression of net photosynthesis in Ramalina maciformis at high water content is caused by increased thallus carbon dioxide diffusion resistance. Oecologia, 51, 426–429.CrossRefGoogle Scholar
Lange, O. L. and Wagenitz, G. (2004). Vernon Ahmadjian introduced the term “chlorolichen”. Lichenologist, 36, 171.CrossRefGoogle Scholar
Lange, O. L. and Ziegler, H. (1963). Der Schwermetallgehalt von Flechten aus dem Acarosporetum sinopicae auf Erzschlackenhalden des Harzes. q. Eisen und Kupfer. Mitteilungen der floristischsoziologischen Arbeitsgemeinschaft, neue Folge, 10, 156–183.Google Scholar
Lange, O. L., Bilger, W., Rimke, S. and Schreiber, U. (1989). Chlorophyll fluorescence of lichens containing green and blue-green algae during hydration by water vapor uptake and by addition of liquid water. Botanica Acta, 102, 306–313.CrossRefGoogle Scholar
Lange, O. L., Büdel, B., Heber, U., et al. (1993 a). Temperate rainforest lichens in New Zealand: high thallus water content can severely limit photosynthetic CO2 exchange. Oecologia, 95, 303–313.CrossRefGoogle ScholarPubMed
Lange, O. L., Büdel, B., Meyer, A. and Kilian, E. (1993 b). Further evidence that activation of net photosynthesis by dry cyanobacterial lichens requires liquid water. Lichenologist, 25, 175–189.CrossRefGoogle Scholar
Lange, O. L., Büdel, B., Meyer, A., Zellner, H. and Zotz, G. (2000). Lichen carbon gain under tropical conditions: water relations and CO2 exchange of three Leptogium species of a lower montane rainforest in Panama. Flora, 195, 172–190.CrossRefGoogle Scholar
Lange, O. L., Büdel, B., Zellner, H., Zotz, G. and Meyer, A. (1994). Field measurements of water relations and CO2 exchange of the tropical, cyanobacterial basidiolichen Dictyonema glabratum in a Panamanian rainforest. Botanica Acta, 107, 279–290.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A. and Heber, U. (2001). Hydration-dependent photosynthetic production of lichens: what do laboratory studies tell us about field performance?Journal of Experimental Botany, 52, 2033–2042.CrossRefGoogle ScholarPubMed
Lange, O. L., Green, T. G. A., Melzer, B., Meyer, A. and Zellner, H. (2006). Water relations and CO2 exchange of the terrestrial lichen Teloschistes capensis in the Namib fog desert: measurements during two seasons in the field and under controlled conditions. Flora, 201, 268–280.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A. and Reichenberger, H. (1999 b). The response of lichen photosynthesis to external CO2 concentration and its interaction with thallus water-status. Journal of Plant Physiology, 154, 157–166.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A. and Ziegler, H. (1988). Water status related photosynthesis and carbon isotope discrimination in species of the lichen genus Pseudocyphellaria with green or blue-green photobionts and in photosymbiodemes. Oecologia, 75, 494–501.CrossRefGoogle ScholarPubMed
Lange, O. L., Hahn, S. C., Meyer, A. and Tenhunen, J. D. (1998). Upland tundra in the foothills of the Brooks Range, Alaska, U.S.A.: lichen long-term photosynthetic CO2 uptake and net carbon gain. Arctic and Alpine Research, 30, 252–261.CrossRefGoogle Scholar
Lange, O. L., Kilian, E. and Ziegler, H. (1986). Water vapour uptake and photosynthesis of lichens: performance differences in species with green and blue-green algae as phycobionts. Oecologia, 71, 104–110.CrossRefGoogle Scholar
Lange, O. L., Kilian, E. and Ziegler, H. (1990 b). Photosynthese von Blattflechten mit hygroskopischen Thallusbewegungen bei Befeuchtung durch Wasserdampf oder mit flüssigem Wasser. Bibliotheca Lichenologica, 38, 311–323.Google Scholar
Lange, O. L., Leisner, J. M. R. and Bilger, W. (1999 a). Chlorophyll fluorescence characteristics of the cyanobacterial lichen Peltigera rufescens under field conditions. II. Diel and annual distribution of metabolic activity and possible mechanisms to avoid photoinhibition. Flora, 194, 413–430.CrossRefGoogle Scholar
Lange, O. L., Meyer, A., Zellner, H., Ullmann, I. and Wessels, D. C. J. (1990 a). Eight days in the life of a desert lichen: water relations and photosynthesis of Teloschistes capensis in the coastal fog zone of the Namib Desert. Madoqua, 17, 17–30.Google Scholar
Lange, O. L., Reichenberger, H. and Meyer, A. (1995). High thallus water content and photosynthetic CO2 exchange of lichens. Laboratory experiments with soil crust species from local xerothermic steppe formations in Franconia, Germany. In Flechten Follmann. Contributions to Lichenology in Honour of Gerhard Follmann, ed. Daniëls, F. J. A., Schulz, M., and Peine, J., pp. 139–153. Cologne: Geobotanical and Phytotaxonomical Study Group, University of Cologne.Google Scholar
Lange, O. L., Reichenberger, H. and Walz, H. (1997). Continuous monitoring of CO2 exchange of lichens in the field: short-term enclosure with an automatically operating cuvette. Lichenologist, 29, 259–274.CrossRefGoogle Scholar
Lange, O. L., Tenhunen, J. D., Harley, P. C. and Walz, H. (1985). Method for field measurements of CO2-exchange. The diurnal changes in net photosynthesis and photosynthetic capacity of lichens under mediterranean climatic conditions. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 23–39. New York: Plenum Press.CrossRefGoogle Scholar
Larcher, W. (2003). Physiological Plant Ecology. Berlin: Springer.CrossRefGoogle Scholar
Larocque, S. J. and Smith, D. J. (2004). Calibrated Rhizocarpon spp. growth curve for the Mount Waddington area, British Columbia coast mountains, Canada. Arctic, Antarctic, and Alpine Research, 36, 407–418.CrossRefGoogle Scholar
Larson, D. W. (1983). The pattern of production within individual Umbilicaria lichen thalli. New Phytologist, 94, 409–419.CrossRefGoogle Scholar
Larson, D. W. (1984). Thallus size as a complicating factor in the physiological ecology of lichens. New Phytologist, 97, 87–97.CrossRefGoogle Scholar
Larson, D. W. (1987). The absorption and release of water by lichens. Bibliotheca Lichenologica, 25, 351–360.Google Scholar
Larson, D. W. and Carey, C. K. (1986). Phenotypic variation with “individual” lichen thalli. American Journal of Botany, 73, 214–223.CrossRefGoogle Scholar
Larson, D. W. and Kershaw, K. A. (1974). Acclimation in arctic lichens. Nature, 254, 421–423.CrossRefGoogle Scholar
Larson, D. W. and Kershaw, K. A. (1975). Studies on lichen-dominated systems. XIII. Seasonal and geographical variation of net CO2 exchange of Alectoria ochroleuca. Canadian Journal of Botany, 53, 2598–2607.CrossRefGoogle Scholar
Larson, D. W., Matthes-Sears, U. and Nash, T. H. III (1985). The ecology of Ramalina menziesii. I. Geographical variation in form. Canadian Journal of Botany, 63, 2062–2068.CrossRefGoogle Scholar
Laufer, Z., Beckett, R. P. and Minibayeva, F. V. (2006 a). Co-occurrence of the multicopper oxidases tyrosinase and laccase in lichens in sub-order Peltigerineae. Annals of Botany, 98, 1035–1042.CrossRefGoogle ScholarPubMed
Laufer, Z., Beckett, R. P., Minibayeva, F. V., Lüthje, S. and Böttger, M. (2006 b). Occurrence of laccases in lichenized ascomycetes of the Peltigerineae. Mycological Research, 110, 846–853.CrossRefGoogle ScholarPubMed
Laundon, J. R. (1978). Haematomma chemotypes form fused thalli. Lichenologist, 10, 221–225.CrossRefGoogle Scholar
Lawrey, J. D. (1984). Biology of Lichenized Fungi. New York: Praeger.Google Scholar
Lawrey, J. D. (1986). Biological role of lichen substances. Bryologist, 89, 111–122.CrossRefGoogle Scholar
Lawrey, J. and Diederich, P. (2003). Lichenicolous fungi: interactions, evolution, and biodiversity. Bryologist, 106, 80–120.CrossRefGoogle Scholar
Laxen, D. P. H. and Thompson, M. N. A. (1987). Sulphur dioxide in Greater London, 1931–1985. Environmental Pollution, 43, 103–114.CrossRefGoogle Scholar
LeBlanc, F. and Sloover, J. (1970). Relation between industrialization and the distribution and growth of epiphytic lichens and mosses in Montreal. Canadian Journal of Botany, 48, 1485–1496.CrossRefGoogle Scholar
LeBlanc, F., Rao, D. N. and Comeau, G. (1972). Indices of atmospheric purity and fluoride pollution pattern in Arvida, Quebec. Canadian Journal of Botany, 50, 991–998.CrossRefGoogle Scholar
LeBlanc, F., Robitaille, G. and Rao, D. N. (1974). Biological response of lichens and bryophytes to environmental pollution in the Murdochville Copper Mine area, Quebec. Journal of the Hattori Botanical Laboratory, 38, 405–433.Google Scholar
Lechowicz, M. J. (1982). Ecological trends in lichen photosynthesis. Oecologia, 53, 330–336.CrossRefGoogle ScholarPubMed
Lee, D. E., Lee, W. G. and Mortimer, N. (2001). Where and why have all the flowers gone? Depletion and turnover in the New Zealand Cenozoic angiosperm flora in relation to palaeography and climate. Australian Journal of Botany, 49, 341–356.CrossRefGoogle Scholar
Legaz, M. E., Fontaniella, B., Millanes, A. M. and Vicente, C. (2004). Secreted arginases from phylogenetically far-related lichen species act as cross-recognition factors for two different algal cells. European Journal of Cell Biology, 83, 435–446.CrossRefGoogle Scholar
Leisner, J. M. R., Green, T. G. A. and Lange, O. L. (1997). Photobiont activity of a temperate crustose lichen: long-term chlorophyll fluorescence and CO2 exchange measurements in the field. Symbiosis, 23, 165–182.Google Scholar
Leprince, O., McKersie, B. D. and Hendry, G. A. (1993). The mechanisms of desiccation tolerance in developing seeds. Seed Science Research, 3, 231–246.CrossRefGoogle Scholar
Leuckert, C., Ahmadjian, V., Culberson, C. F. and Johnson, A. (1990). Xanthones and depsidones of the lichen Lecanaora dispersa in nature and of its mycobiont in culture. Mycologia, 82, 370–378.CrossRefGoogle Scholar
Leverenz, J. and Jarvis, P. G. (1979). Photosynthesis in Sitka spruce. VIII. The effects of light flux density and direction on the rate of net photosynthesis and the stomatal conductance of needles. Journal of Applied Ecology, 16, 919–932.CrossRefGoogle Scholar
Leverenz, J. W., Falk, S., Pilström, C.-M. and Samuelsson, G. (1990). The effects of photoinhibition on the photosynthetic light-response curve of green plant cells (Chlamydomonas reinhardtii). Planta, 182, 161–168.CrossRefGoogle Scholar
Lewis, D. H. (1973). Concepts in fungal nutrition and the origin of parasitism and mutualism. Biological Reviews, 48, 261–278.CrossRefGoogle Scholar
Lewis, D. H. and Smith, D. C. (1967). Sugar alcohols (polyols) in fungi and green plants. I. Distribution, physiology and metabolism. New Phytologist, 66, 143–184.CrossRefGoogle Scholar
Lewis, L. A. and McCourt, R. M. (2004). Green algae and the origin of land plants. American Journal of Botany, 91, 1535–1556.CrossRefGoogle ScholarPubMed
Smith, Lewis R. I. (1995). Colonization by lichens and the development of lichen-dominated communities in the maritime Antarctic. Lichenologist, 27, 473–483.CrossRefGoogle Scholar
Lex, M., Silvester, W. B. and Stewart, W. D. P. (1972). Photorespiration and nitrogenase activity in the blue-green alga, Anabaena cylindrica. Proceedings of the Royal Society of London B, 180, 87–102.CrossRefGoogle ScholarPubMed
Li, Y. [M.] (2006). Seasonal variation of diet and food availability in a group of Sichuan snub-nosed monkeys in Shennongjia Nature Reserve, China. American Journal of Primatology, 68, 217–233.Google Scholar
Liberatore, S., Garibotti, G. and Calvelo, S. (2002). Phytogeography of Argentinean lichens. Bibliotheca Lichenologica, 82, 221–234.Google Scholar
Lidén, K. and Gustafsson, M. (1967). Relationships and seasonal variation of 137Cs in lichen, reindeer and man in northern Sweden 1961–1965. In Radioecological Concentration Processes, ed. Aberg, B. and Hungate, F. P., pp. 193–208. Oxford: Pergamon Press.Google Scholar
Lilly, V. G. and Barnett, H. L. (1951). Physiology of the Fungi. New York: McGraw-Hill Book Co.Google Scholar
Lindblom, L. and Ekman, S. (2006). Genetic variation and population differentiation in the lichen-forming ascomycete Xanthoria parietina on the island Storfosna, central Norway. Molecular Ecology, 15, 1545–1559.CrossRefGoogle ScholarPubMed
Linder, M. B., Szilvay, G. R., Nakari-Setälä, T. and Penttilä, M. E. (2005). Hydrophobins: the protein-amphiphiles of filamentous fungi. FEMS Microbiology Reviews, 29, 877–896.CrossRefGoogle ScholarPubMed
Lines, C. E. M., Ratcliffe, R. G., Rees, T. A. V. and Southon, T., E. (1989). A 13C NMR study of photosynthate transport and metabolism in the lichen Xanthoria calcicola Oxner. New Phytologist, 111, 447–456.CrossRefGoogle Scholar
Link, S. O. and Nash, T. H. III (1984 a). Ecophysiological studies of the lichen, Parmelia praesignis Nyl. Population variation and the effect of storage conditions. New Phytologist, 96, 249–256.CrossRefGoogle Scholar
Link, S. O. and Nash, T. H. III (1984 b). A mathematical description of the effect of resaturation on net photosynthesis in the lichen, Parmelia praesignis Nyl. New Phytologist, 96, 257–262.CrossRefGoogle Scholar
Link, S. O., Nash, T. H., III and Driscoll, M. (1985). CO2 exchange in lichens: towards a mechanistic model. In Lichen Physiology and Cell Biology, ed. Brown, D. H., pp. 77–91. New York: Plenum Press.CrossRefGoogle Scholar
Linnaeus, C. (1753). Species plantarum, Vol. 2. Stockholm: L. Salvi.Google Scholar
Lipscomb, D. L., Farris, J. S., Källersjö, M. and Tehler, A. (1998). Support, ribosomal sequences, and the phylogeny of the eukaryotes. Cladistics, 14, 303–38.CrossRefGoogle Scholar
Litterski, B. and Ahti, T. (2004). World distribution of selected European Cladonia species. Symbolae Botanicae Upsalienses, 34, 205–236.Google Scholar
Litterski, B. and Otte, V. (2002). Biogeographical research on European species of selected lichen genera. Bibliotheca Lichenologica, 82, 83–90.Google Scholar
Liu, Y. J. and Hall, B. D. (2004). Body plan evolution of ascomycetes, as inferred from an RNA polymerase II phylogeny. Proceedings of the National Academy of Sciences, USA, 101, 4507–4512.CrossRefGoogle ScholarPubMed
Llimona, X. and Hladun, N. L. (2001). Checklist of the Lichens and lichenicolous Fungi of the Iberian Peninsula and the Balearic Islands. Bocconea, 14, 1–581.Google Scholar
Lodenius, M. and Laaksovirta, K. (1979). Mercury content of Hypogymnia physodes and pine needles affected by a chlor-alkali works in Kuusankoski, SE Finland. Annales Botanici Fennici, 16, 7–10.Google Scholar
Loewus, F. A. (1999). Biosynthesis and metabolism of ascorbic acid in plants and of analogs of ascorbic acid in fungi. Phytochemistry, 52, 193–210.CrossRefGoogle Scholar
Lohtander, K., Oksanen, I. and Rikkinen, J. (2003). Genetic diversity of green algal and cyanobacterial photobionts in Nephroma (Peltigerales). Lichenologist, 4, 325–339.CrossRefGoogle Scholar
Lomolino, M. V., Riddle, B. R. and Brown, J. H. (2006). Biogeography. 3rd edn. Sunderland: Sinauer Associates.Google Scholar
Longton, R. E. (1988). Biology of Polar Bryophytes and Lichens. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lorenzini, G., Landi, U., Loppi, S. and Nali, C. (2003). Lichen distribution and bioindicator tobacco plants give discordant response: a case study from Italy. Environmental Monitoring and Assessment, 82, 243–264.CrossRefGoogle ScholarPubMed
Loso, M. G. and Doak, D. F. (2006). The biology behind lichenometric dating curves. Oecologia, 147, 223–229.CrossRefGoogle ScholarPubMed
Louwhoff, S. H. J. J. (2001). Biogeography of Hypotrachyna, Parmotrema and allied genera (Parmeliaceae) in the Pacific islands. Bibliotheca Lichenologica, 78, 223–246.Google Scholar
Lücking, R. (1997). The use of foliicolous lichens as bioindicators in the tropics, with special reference to the microclimate. Abstracta Botanica, 21, 99–116.Google Scholar
Lücking, R. (2003). Takhtajan's floristic regions and foliicolous lichen biogeography: a compatibility analysis. Lichenologist 35, 33–54.CrossRefGoogle Scholar
Lücking, R. and Bernecker-Lücking, A. (2002). Distance, dynamics, and diversity in tropical rainforests: an experimental approach using foliicolous lichens on artificial leaves. I. Growth performance and succession. Ecotropica, 8, 1–13.Google Scholar
Lücking, R. and Kalb, K. (2001). New Caledonia, foliicolous lichens and island biogeography. Bibliotheca Lichenologica, 78, 247–273.Google Scholar
Lücking, R., Sérusiaux, E. and Vezda, A. (2005). Phylogeny and systematics of the lichen family Gomphillaceae (Ostropales) inferred from cladistic analysis of phenotype data. Lichenologist, 37, 123–170.CrossRefGoogle Scholar
Lücking, R., Wirth, V., Ferraro, L. and Caceres, M. E. S. (2003). Foliicolous lichens from Valdivian temperate rain forest of Chile and Argentina: evidence of an austral element, with the description of seven new taxa. Global Ecology and Biogeography, 12, 21–36.CrossRefGoogle Scholar
Lumbsch, H. T. (1998). The use of metabolic data in lichenology at the species and subspecific levels. Lichenologist, 30, 357–367.CrossRefGoogle Scholar
Lumbsch, H. T. and Elix, J. A. (1985). A new species of the lichen genus Diploschistes from Australia. Plant Systematics and Evolution, 150, 275–279.CrossRefGoogle Scholar
Lumbsch, H. T. and Kothe, H. W. (1988). Anatomical features of Chondropsis semiviridis (Nyl.) Nyl. in relation to its vagrant habit. Lichenologist, 20, 25–29.CrossRefGoogle Scholar
Lumbsch, H. T., Prado, R. and Kantvilas, G. (2005). Gregorella, a new genus to accommodate Moelleropsis humida and a molecular phylogeny of Arctomiaceae. Lichenologist, 37, 291–302.CrossRefGoogle Scholar
Lumbsch, H. T., Schmitt, I., Döring, H. and Wedin, M. (2001 a). ITS sequence data suggest variability of ascus types and support ontogenetic characters as phylogenetic discriminators in the Agyriales (Ascomycota). Mycological Research, 105, 265–274.CrossRefGoogle Scholar
Lumbsch, H. T., Schmitt, I., Döring, H. and Wedin, M. (2001 b). Molecular systematics supports the recognition of an additional order of Ascomycota: the Agyriales. Mycological Research, 105, 16–23.CrossRefGoogle Scholar
Lumbsch, H. T., Schmitt, I., Lücking, R., Wiklund, E. and Wedin, M. (2007). The phylogenetic placement of Ostropales within Lecanoromycetes (Ascomycota) revisited. Mycological Research, 111, 257–267.CrossRefGoogle ScholarPubMed
Lumbsch, H. T., Schmitt, I., Palice, Z., Wiklund, E. and Wedin, M. (2004). Supraordinal phylogenetic relationships of Lecanoromycetes based on a Bayesian analysis of combined nuclear and mitochondrial sequences. Molecular Phylogenetics and Evolution, 31, 822–832.CrossRefGoogle ScholarPubMed
Lumbsch, H. T., Wirtz, N., Lindemuth, R. and Schmitt, I. (2002). Higher level phylogenetic relationships of Euascomycetes (Pezizomycotina) inferred from a combined analysis of nuclear and mitochondrial sequence data. Mycological Progress, 1, 57–70.CrossRefGoogle Scholar
Lutzoni, F., Kauff, F., Cox, C., et al. (2004). Assembling the fungal tree of life: progress, classification, and evolution of subcellular traits. American Journal of Botany, 91, 1446–1480.CrossRefGoogle ScholarPubMed
Lutzoni, F., Pagel, M. and Reeb, V. (2001). Major fungal lineages are derived from lichen symbiotic ancestors. Nature, 411, 937–940.CrossRefGoogle ScholarPubMed
MacCracken, J. G., Alexander, L. E. and Uresk, D. W. (1983). An important lichen of southeastern Montana rangelands. Journal of Range Management, 36, 35–37.CrossRefGoogle Scholar
MacDonald, G. M. (2003). Biogeography: Space, Time, and Life. New York: John Wiley.Google Scholar
MacFarlane, J. D. and Kershaw, K. A. (1977). Physiological-environmental interactions in lichens. IV. Seasonal changes in the nitrogenase activity in Peltigera canina (L.) Willd. var. praetextata (Floerke in Somm.) Hue, and P. canina (L.) Willd. var. rufescens (Weiss) Mudd. New Phytologist, 69, 403–408.CrossRefGoogle Scholar
MacFarlane, J. D. and Kershaw, K. A. (1980). Physiological-environmental interactions in lichens. IX. Thermal stress and lichen ecology. New Phytologist, 84, 669–685.CrossRefGoogle Scholar
MacFarlane, J. D. and Kershaw, K. A. (1982). Physiological-environmental interactions in lichens. XIV. The environmental control of glucose movement from alga to fungus in Peltigera polydactyla, P. rufescens, and Collema furfuraceum. New Phytologist, 91, 93–101.CrossRefGoogle Scholar
MacGinitie, H. (1937). The flora of the Weaverville beds of Trinity County, California, with descriptions of the plant-bearing beds. In Eocene Flora of Western America, pp. 83–151. Publication 465. Washington: Carnegie Institution of Washington.Google Scholar
MacKenzie, T. D. B., Król, M., Huner, N. P. A. and Campbell, D. A. (2002). Seasonal changes in chlorophyll fluorescence quenching and the induction and capacity of the photoprotective xanthophyll cycle in Lobaria pulmonaria. Canadian Journal of Botany, 80, 255–261.CrossRefGoogle Scholar
MacKenzie, T. D. B., MacDonald, T. M., Dubois, L. A. and Campbell, D. A. (2001). Seasonal changes in temperature and light driven acclimation of photosynthetic physiology and macromolecular content in Lobaria pulmonaria. Planta, 214, 57–66.CrossRefGoogle Scholar
Madelin, M. F. (1968). Parasitism on other fungi and lichens. In The Fungi, Vol. III: The Fungal Population, ed. Ainsworth, G. C. and Sussman, A. S., pp. 253–269. New York: Academic Press.Google Scholar
Magan, N. (1997). Fungi in extreme environments. In Environmental and Microbial Relationships, Vol. IV: The Mycota, ed. Wicklow, D. and Soderstrom, B., pp. 99–114. Berlin: Springer.Google Scholar
Mägdefrau, K. (1957). Flechten und Moose im baltischen Bernstein. Berichte der deutschen botanischen Gesellschaft, 9, 433–435.Google Scholar
Máguas, C. and Griffiths, H. (2003). Applications of stable isotopes in plant ecology. Progress in Botany, 64, 472–505.Google Scholar
Máguas, C., Valladares, F. and Brugnoli, E. (1997). Effects of thallus size on morphology and physiology of foliose lichens: new findings with a new approach. Symbiosis, 23, 149–164.Google Scholar
Majerus, M. E. N. (1998). Melanism: Evolution in Action. Oxford: Oxford University Press.Google Scholar
Makkonen, S., Hurri, R. and Hyvärinen, M. (2007). Differential responses of lichen symbionts to enhanced nitrogen and phosphorous availability: an experiment with Cladina stellaris. Annals of Botany, 99, 877–884.CrossRefGoogle ScholarPubMed
Malcolm, W. M. and Galloway, D. J. (1997). New Zealand Lichens. Checklist, Key, and Glossary. Wellington: Museum of New Zealand Te Papa Tongarewa.Google Scholar
Malhotra, S. S. and Khan, A. A. (1983). Sensitivity to SO2 of various metabolic processes in an epiphytic lichen, Evernia mesomorpha. Biochemie und Physiologie der Pflanzen, 178, 121–130.CrossRefGoogle Scholar
Manodori, A. M. and Melis, A. (1984). Photochemical apparatus organization in Anacystis nidulans (Cyanophyceae). Effect of CO2 concentration during cell growth. Plant Physiology, 74, 67–71.CrossRefGoogle Scholar
Margot, J. (1973). Experimental study of the effects of sulphur dioxide on the soredia of Hypogymnia physodes. In Air Pollution and Lichens, ed. Ferry, B. W., Baddeley, M. S. and Hawksworth, D. L., pp. 314–329. Toronto: University of Toronto Press.Google Scholar
Margules, C. R. and Pressey, R. L. (2000). Systematics conservation planning. Nature, 405, 243–253.CrossRefGoogle Scholar
Margulis, L. and Fester, R. (eds.) (1991). Symbiosis as a Source of Evolutionary Innovation: Speciation and Morphogenesis. Cambridge: Massachusetts Institute of Technology Press.Google ScholarPubMed
Marsh, J. E. and Nash, T. H. III (1979). Lichens in relation to the Four Corners Power Plant in New Mexico. Bryologist, 82, 20–28.CrossRefGoogle Scholar
Marshall, W. A. (1996). Aerial dispersal of lichen soredia in the maritime antarctic. New Phytologist, 134, 523–530.CrossRefGoogle Scholar
Marti, J. (1983). Sensitivity of lichen phycobionts to dissolved air pollutants. Canadian Journal of Botany, 61, 1647–1653.CrossRefGoogle Scholar
Martin, D., Ciulla, R. and Roberts, M. (1999). Osmoadaptation in Archaea. Applied and Environmental Microbiology, 65, 1815–1825.Google ScholarPubMed
Masterson, C. L. and Murphy, P. M. (1984). The acetylene reduction technique. In Current Developments in Biological Nitrogen-fixation, ed. Rao, N. S. Subba, pp. 8–33. Baltimore: Edward Arnold.Google Scholar
Matthes-Sears, U. and Nash, T. H. III (1986). The ecology of Ramalina menziesii. V. Estimation of gross carbon and thallus hydration source from diurnal measurements and climatic data. Canadian Journal of Botany, 64, 1698–1702.CrossRefGoogle Scholar
Matthews, J. A. (2005). ‘Little Ice Age’ glacier variations in Jotunheimen, southern Norway: a study in regionally controlled lichenometric dating of recessional moraines with implications for climate and lichen growth rates. Holocene, 15, 1–19.CrossRefGoogle Scholar
Mattox, K. R. and Stewart, K. D. (1984). Classification of the green algae: a concept based on comparative cytology. In Systematics of the Green Algae, ed. Irvine, D. E. G. and John, D. M., pp. 29–72. London: Academic Press.Google Scholar
Mattson, J. E. (1991). Protein banding patterns in some American and European species of Cetraria. Bryologist, 94, 261–269.CrossRefGoogle Scholar
Mayaba, N. and Beckett, R. P. (2001). The effect of desiccation on the activities of antioxidant enzymes in lichens from habitats of contrasting water status. Symbiosis, 31, 113–121.Google Scholar
McCall, K. K. and Martin, C. E. (1991). Chlorophyll concentrations and photosynthesis in three forest understorey mosses in northeastern Kansas. Bryologist, 94, 25–29.CrossRefGoogle Scholar
McCarroll, D. (1993). Modelling late-Holocene snow-avalanche activity, incorporating a new approach to lichenometry. Earth Surface Processes and Landforms, 18, 527–539.CrossRefGoogle Scholar
McCarroll, D., Shakesby, R. and Matthews, J. A. (1998). Spatial and temporal patterns of late Holocene rockfall activity on a Norwegian talus slope: a lichenometric and simulation-modeling approach. Arctic and Alpine Research, 30, 51–60.CrossRefGoogle Scholar
McCarthy, D. P. (1999). A biological basis for lichenometry?Journal of Biogeography, 26, 379–386.CrossRefGoogle Scholar
McCarthy, P. M. (2001). Polyblastia. Flora of Australia, 58A, 171–172.
McCarthy, P. M. (2004). Maronina. Flora of Australia, 56A, 62–63.
McCarthy, P. M. (2006). Checklist of the Lichens of Australia and its Island Territories. Australian Biological Resources Study, Canberra. Version 6 April 2006. Online: www.anbg.gov.au/abrs/lichenlist/introduction.html.
McCarthy, P. M. and Healey, J. A. (1978). Dispersal of lichen propagules by slugs. Lichenologist, 10, 131–132.CrossRefGoogle Scholar
McCune, B. (1988). Lichen communities along O3 and SO2 gradients in Indianapolis. Bryologist, 91, 223–228.CrossRefGoogle Scholar
McCune, B. (1994). Using epiphytic litter to estimate epiphyte biomass. Bryologist, 97, 396–401.CrossRefGoogle Scholar
McCune, B. and Daly, W. J. (1994). Consumption and decomposition of lichen litter in a temperate coniferous rainforest. Lichenologist, 26, 67–71.Google Scholar
McCune, B. and Grace, J. B. (2002). Analysis of Ecological Communities. Gleneden Beach, OR: MjM Software.Google Scholar
McCune, B., Berryman, S. D., Cissel, J. H. and Gitelman, A. I. (2003). Use of a smoother to forecast occurrence of epiphytic lichens under alternative forest management plans. Ecological Applications, 13, 1110–1123.CrossRefGoogle Scholar
McDowall, R. M. (2004). What biogeography is: a place for process. Journal of Biogeography, 31, 345–351.CrossRefGoogle Scholar
McEvoy, M. (2006). Acclimation of the photobiont and mycobiont partners in lichens to high solar radiation. Ph.D. thesis. Ås, Norway: Department of Ecology and Natural Resource Management. Norwegian University of Life Sciences.
McGlone, M. S. (2005). Goodbye Gondwana. Journal of Biogeography, 32, 739–740.CrossRefGoogle Scholar
McGlone, M. S., Duncan, R. P. and Heenan, P. B. (2001). Endemism, species selection and the origin and distribution of the vascular plant flora of New Zealand. Journal of Biogeography, 28, 199–216.CrossRefGoogle Scholar
McKersie, B. D. and Lesham, Y. Y. (1994). Stress and Stress Coping in Cultivated Plants. Dordrecht: Kluwer.CrossRefGoogle Scholar
McNabb, D. H. and Geist, J. M. (1979). Acetylene reduction assay of symbiotic N2 fixation under field conditions. Ecology, 60, 1070–1072.CrossRefGoogle Scholar
Meier, F. A., Scherrer, S. and Honegger, R. (2002). Faecal pellets of lichenivorous mites contain viable cells of the lichen-forming ascomycete Xanthoria parietina and its green algal photobiont, Trebouxia arboricola. Biological Journal of the Linnean Society, 76, 259–268.CrossRefGoogle Scholar
Melkonian, M. (1990). Chlorophyte orders of uncertain affinities: Order Microthamniales. In Handbook of Protoclista, ed. Margulis, L., Corliss, J. O., Melkonian, M. and Chapman, D. J., pp. 652–654. Boston: Jones and Bartlett.Google Scholar
Melkonian, M. and Peveling, E. (1988). Zoospore ultrastructure in species of Trebouxia and Pseudotrebouxia (Chlorophyta). Plant Systematics and Evolution, 158, 183–210.CrossRefGoogle Scholar
Miadlikowska, J. and Lutzoni, F. (2004). Phylogenetic classification of peltigeralean fungi (Peltigerales, Ascomycota) based on ribosomal RNA small and large subunits. American Journal of Botany, 91, 449–464.CrossRefGoogle ScholarPubMed
Miadlikowska, J., Arnold, A. E., Hofstetter, V. and Lutzoni, F. (2004 a). High diversity of cryptic fungi inhabiting healthy lichen thalli in a temperate and tropical forest. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 43. Tartu: Tartu University Press.Google Scholar
Miadlikowska, J., Arnold, A. and Lutzoni, F. (2004 b). Diversity of cryptic fungi inhabiting healthy lichen thalli in a temperate and tropical forest. Ecological Society of America Annual Meeting, 89, 349–350.Google Scholar
Miao, V. P. W., Rabenau, A. and Lee, A. (1997). Cultural and molecular characterization of photobionts of Peltigera membranacea. Lichenologist, 29, 571–587.CrossRefGoogle Scholar
Micallef, A. and Colls, J. J. (1999). Analysis of long-term measurements of airborne concentrations of sulphur dioxide and SO42 − in the rural United Kingdom. Environmental Monitoring and Assessment, 57, 277–290.CrossRefGoogle Scholar
Mies, B. and Printzen, C. (1997). Notes on the lichens of Socotra (Yemen, Indian Ocean). Bibliothecia Lichenologica, 68, 223–239.Google Scholar
Mietzsch, E., Lumbsch, H. T. and Elix, J. A. (1993). Notice: a new computer program for the identification of lichen substances. Mycotaxon, 47, 475–479.Google Scholar
Mikhailova, I. N. and Scheidegger, C. (2001). Early development of Hypogymnia physodes (L.) Nyl. in response to emissions from a copper smelter. Lichenologist, 33, 527–538.CrossRefGoogle Scholar
Millbank, J. W. (1972). Nitrogen metabolism in lichens. IV. The nitrogenase activity of the Nostoc phycobiont in Peltigera canina. New Phytologist, 71, 1–10.CrossRefGoogle Scholar
Millbank, J. W. (1974). Nitrogen metabolism in lichens. V. The forms of nitrogen released by the blue-green phycobiont in Peltigera spp. New Phytologist, 73, 1171–1181.CrossRefGoogle Scholar
Millbank, J. W. (1976). Aspects of nitrogen metabolism in lichens. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 441–455. London: Academic Press.Google Scholar
Millbank, J. W. (1981). The assessment of nitrogen fixation and throughput by lichens. I. The use of a controlled environment chamber to relate acetylene reduction estimates to fixation. New Phytologist, 89, 647–655.CrossRefGoogle Scholar
Millbank, J. W. (1982 a). Nitrogenase and hydrogenase in cyanophilic lichens. New Phytologist, 92, 221–228.CrossRefGoogle Scholar
Millbank, J. W. (1982 b). The assessment of nitrogen fixation and throughput by lichens. III. Losses of nitrogenous compounds by Peltigera membranacea, P. polydactyla and Lobaria pulmonaria in simulated rainfall episodes. New Phytologist, 92, 229–234.CrossRefGoogle Scholar
Millbank, J. W. and Olsen, J. D. (1986). The assessment of nitrogen fixation and throughput by lichens. IV. Nitrogen losses from Peltigera membranacea (Ach.) Nyl. in autumn, winter and spring. New Phytologist, 104, 643–651.CrossRefGoogle Scholar
Miller, G. H. (1973). Variations in lichen growth from direct measurements: preliminary curves for Alectoria minuscula from eastern Baffin Island, N. W. T., Canada. Arctic and Alpine Research, 5, 33–42.CrossRefGoogle Scholar
Miller, J. E. and Brown, D. H. (1999). Studies of ammonia uptake and loss by lichens. Lichenologist, 31, 85–93.Google Scholar
Miller, P. R. and McBride, J. R. (eds.) (1998). Oxidant Air Pollution Impacts in the Montane Forests of Southern California. Ecological Studies 134. New York: Springer.Google Scholar
Miller, P. R., Longbotham, G. J. and Longbotham, C. R. (1983). Sensitivity of selected western conifers to ozone. Plant Disease, 67, 1113–1115.CrossRefGoogle Scholar
Miszalski, Z. and Niewiadomska, E. (1993). Comparison of sulphite oxidation mechanisms in three lichen species. New Phytologist, 123, 345–349.CrossRefGoogle Scholar
Mitchell, R. J., Truscot, A. M., Leith, I. D., et al. (2005). A study of the epiphytic communities of Atlantic oak woods along an atmospheric nitrogen deposition gradient. Journal of Ecology, 93, 482–492.CrossRefGoogle Scholar
Moberg, R. (1990). Waynea, a new lichen genus in the Bacidiaceae from California. Lichenologist, 22, 249–252.CrossRefGoogle Scholar
Modenesi, P. (1993). An SEM study of injury symptoms in Parmotrema reticulatum treated with paraquat or growing in sulphur dioxide-polluted air. Lichenologist, 25, 423–433.CrossRefGoogle Scholar
Mohr, F., Ekman, S. and Heegaard, E. (2004). Evolution and taxonomy of the marine Collemopsidium species (lichenized Ascomycota) in north-west Europe. Mycological Research, 108, 515–532.CrossRefGoogle Scholar
Molina, M. C., Crespo, A., Vicente, C. and Elix, J. A. (2003). Differences in the composition of phenolics and fatty acids of cultured mycobiont and thallus of Physconia distorta. Plant Physiology and Biochemistry, 41, 175–180.CrossRefGoogle Scholar
Mollenhauer, D. (1992). Geosiphon pyriforme. In Algae and Symbioses: Plants, Animals, Fungi, Viruses, Interactions Explored, ed. Reisser, W., pp. 339–351. Bristol: Biopress.Google Scholar
Mollenhauer, D., Mollenhauer, R. and Kluge, M. (1996). Studies on initiation and development of the partner association in Geosiphon pyriforme (Kütz.) v. Wettstein, a unique endocytobiotic system of a fungus (Glomales) and the cyanobacterium Nostoc punctiforme (Kütz.) Hariot. Protoplasma, 193, 3–9.CrossRefGoogle Scholar
Möller, C. and Dreyfuss, M. M. (1996). Microfungi from Antarctic lichens, mosses and vascular plants. Mycologia, 88, 922–933.CrossRefGoogle Scholar
Møller, I. M. (2001). Plant mitochondria and oxidative stress. Electron transport, NADPH turnover and metabolism of reactive oxygen species. Annual Review of Plant Physiology and Plant Molecular Biology, 52, 561–591.CrossRefGoogle ScholarPubMed
Monnet, F., Bordas, F., Deluchat, V., et al. (2005). Use of the aquatic lichen Dermatocarpon luridum as bioindicator of copper pollution: accumulation and cellular distribution tests. Environmental Pollution, 138, 455–461.CrossRefGoogle ScholarPubMed
Montalvo, A. M. and Ellstrand, N. C. (2001). Non-local transplantation and outbreeding depression in the subshrub Lotus scoparius (Fabaceae). American Journal of Botany, 88, 258–269.CrossRefGoogle Scholar
Montieth, J. L. (1977). Climate and the efficiency of crop production in Britain. Philosophical Transactions of the Royal Society of London B, 281, 277–294.CrossRefGoogle Scholar
Morrone, J. J. (2005). Cladistic biogeography: identity and place. Journal of Biogeography, 32, 1281–1284.CrossRefGoogle Scholar
Mosbach, K. (1969). Biosynthesis of lichen substances, products of a symbiotic association. Angewandte Chemie, International Edition, 8, 240–250.CrossRefGoogle Scholar
Moser, T. J., Nash, T. H. III and Clark, W. D. (1980). Effects of a long-term sulfur dioxide fumigation on arctic caribou forage lichens. Canadian Journal of Botany, 58, 2235–2240.CrossRefGoogle Scholar
Moser, T. J., Nash, T. H. III and Link, S. O. (1983 a). Diurnal gross photosynthetic patterns and potential seasonal CO2 assimilation in Cladonia stellaris and Cladonia rangiferina. Canadian Journal of Botany, 61, 642–655.CrossRefGoogle Scholar
Moser, T. J., Nash, T. H. III and Olafsen, A. G. (1983 b). Photosynthetic recovery in arctic caribou forage lichens following a long-term field sulfur dioxide fumigation. Canadian Journal of Botany, 61, 367–370.CrossRefGoogle Scholar
Moxham, T. H. (1980). Lichens and perfume manufacture. Bulletin of the British Lichen Society, 47, 1–2.Google Scholar
Muir, D. C. G., Segstro, M. D., Welbourn, P. M., et al. (1993). Patterns of accumulation of airborne organochlorine contaminants in lichens from the Upper Great Lakes Region of Ontario. Environmental Science and Technology, 27, 1201–1210.CrossRefGoogle Scholar
Muir, P. S., Shirazi, A. M. and Patrie, J. (1997). Seasonal growth dynamics in the lichen Lobaria pulmonaria. Bryologist, 100, 458–464.CrossRefGoogle Scholar
Mukhtar, A., Garty, J. and Galun, M. (1994). Does the lichen alga Trebouxia occur free-living in nature? – Further immunological evidence. Symbiosis, 17, 247–253.Google Scholar
Munne-Bosch, S. and Alegre, L. (2002). The function of tocopherols and tocotrienols in plants. Critical Reviews in Plant Science, 21, 31–57.CrossRefGoogle Scholar
Muñoz, J., Felicísmo, A., Cabezas, F., Burgaz, A. and Martínez, I. (2004). Wind as long-distance dispersal vehicle in the Southern Hemisphere. Science, 304, 1144–1147.CrossRefGoogle ScholarPubMed
Murashige, T. and Skoog, F. (1962). A revised medium for rapid growth and bioassays with tobacco tissue cultures. Physiologia Plantarum, 15, 473–494.CrossRefGoogle Scholar
Murtagh, G. J., Dyer, P. S. and Crittenden, P. D. (2000). Reproductive systems – sex and the single lichen. Nature, 404, 564.CrossRefGoogle Scholar
Murtagh, G. J., Dyer, P. S., McClure, P. C. and Crittenden, P. D. (1999). Use of randomly amplified polymorphic DNA markers as a tool to study variation in lichen- forming fungi. Lichenologist, 31, 257–267.CrossRefGoogle Scholar
Myachi, S., Nakayama, O., Yokohama, , et al. (1989). World Catalogue of Algae, 2nd. edn. Tokyo: Japan Scientific Societies Press.Google Scholar
Myllys, L., Högnabba, F., Lohtander, K., et al. (2005). Phylogenetic relationships of Stereocaulaceae based on simultaneous analysis of beta-tubulin, GAPDH and SSU rDNA sequences. Taxon, 54, 605–618.CrossRefGoogle Scholar
Myllys, L, Stenroos, S., Thell, A. and Ahti, T. (2003). Phylogeny of bipolar Cladonia arbuscula and Cladonia mitis (Lecanorales, Euascomycetes). Molecular Phylogenetics and Evolution, 27, 58–69.CrossRefGoogle Scholar
Naef, A., Roy, B. A., Kaiser, R. and Honegger, R. (2002). Insect-mediated reproduction of systemic infections by Puccinia arrhenatheri on Berberis vulgaris. New Phytologist, 154, 717–730.CrossRefGoogle Scholar
Nakano, T., Handa, S. and Takeshita, S. (1991). Some corticolous algae from the Taishaku-kyô Gorge, western Japan. Nova Hedwigia, 52, 427–451.Google Scholar
Nash, T. H. III (1971). Lichen sensitivity to hydrogen fluoride. Bulletin of the Torrey Botanical Club, 98, 103–106.Google Scholar
Nash, T. H. III (1972). Simplification of the Blue Mountain lichen communities near a zinc factory. Bryologist, 75, 315–324.Google Scholar
Nash, T. H. III (1973). Sensitivity of lichens to sulfur dioxide. Bryologist, 76, 333–339.Google Scholar
Nash, T. H. III (1975). Influence of effluents from a zinc factory on lichens. Ecological Monographs, 45, 183–196.CrossRefGoogle Scholar
Nash, T. H., III (1988). Correlating fumigation studies with field effects. In Lichens, Bryophytes and Air Quality ed. Nash, T. H. III and Wirth, V., pp. 201–216. Bibliotheca Lichenologica 30. Berlin: J. Cramer.Google Scholar
Nash, T. H., III (1989). Metal tolerance in lichens. In Heavy Metal Tolerance in Plants: Evolutionary Aspects, ed. Shaw, A. J., pp. 119–131. Boca Raton: CRC Press.Google Scholar
Nash, T. H., III (1996). Photosynthesis, respiration, productivity and growth. In Lichen Biology, ed. Nash, T. H. III, pp. 88–120. Cambridge: Cambridge University Press.Google Scholar
Nash, T. H. III and Gries, C. (2002). Lichens as bioindicators of sulfur dioxide. Symbiosis, 33, 1–21.Google Scholar
Nash, T. H. III and Lange, O. L. (1988). Responses of lichens to salinity: concentration and time-course relationships and variability among Californian species. New Phytologist, 109, 361–367.CrossRefGoogle Scholar
Nash, T. H III, and Moser, T. J. (1982). Vegetational and physiological patterns of lichens in North American deserts. Journal of the Hattori Botanical Laboratory, 53, 331–336.Google Scholar
Nash, T. H. III, and Olafsen, A. G. (1995). Climate change and the ecophysiological response of Arctic lichens. Lichenologist, 27, 559–565.CrossRefGoogle Scholar
Nash, T. H., III and Riddell, J. (2006). Historical perspectives and new opportunities in the use of lichens as air pollutant monitors. Botany 2006 Symposium Abstracts, p. 51. St. Louis: Botanical Society of America.
Nash, T. H. III and Sigal, L. L. (1979). Gross photosynthetic response of lichens to short-term ozone fumigations. Bryologist, 82, 280–285.Google Scholar
Nash, T. H., III and Sigal, L. L. (1980). Sensitivity of lichens to air pollution with an emphasis on oxidant air pollutants. In Proceedings of the Symposium on Effects of Air Pollution on Mediterranean and Temperate Forest Ecosystems, June 22–27, 1980, Riverside, California, U.S.A. Gen. Tech. Rep. PSW-43, ed. Miller, P. R. (prin. coord.), pp. 117–124. Berkley: Pacific Southwest Forest and Range Experiment Station, Forest Service, U.S. Department of Agriculture.Google Scholar
Nash, T. H. III, Moser, T. J. and Link, S. O. (1980). Nonrandom variation of gas exchange within arctic lichens. Canadian Journal of Botany, 58, 1181–1186.CrossRefGoogle Scholar
Nash, T. H. III, Moser, T. J., Link, S. O., et al. (1983). Lichen photosynthesis in relation to CO2 concentration. Oecologia, 58, 52–56.CrossRefGoogle ScholarPubMed
Nash, T. H. III, Reiner, A., Demmig-Adams, B., et al. (1990). The effect of atmospheric desiccation and osmotic water stress on photosynthesis and dark respiration of lichens. New Phytologist, 116, 269–276.CrossRefGoogle Scholar
Nash, T. H. III, Ryan, B. D., Diederich, P., Gries, C. and Bungartz, F., eds., (2004). Lichen Flora of the Greater Sonoran Desert Region. Vol. 2. Tempe: Lichens Unlimited.Google Scholar
Nash, T. H. III, Ryan, B. D., Gries, C. and Bungartz, F. (2002). Lichen Flora of the Greater Sonoran Desert Region. Vol. 1. Tempe: Lichens Unlimited.Google Scholar
Nash, T. H., III, Thomas, M. A., Hoober, J. K., Gries, C. and Zheng, S. X. (2001). Free amino acids in lichens and their symbionts. In Lichenological Contributions in Honour of Jack Elix, ed. McCarthy, P. M., Kantvilas, G. and Louwhoff, S. H. J. J., pp. 313–319. Bibliotheca Lichenologica 78. Berlin: J. Cramer.Google Scholar
Nash, T. H. III, White, S. L. and Marsh, J. E. (1977). Lichen and moss distribution and biomass in hot desert ecosystems. Bryologist, 80, 470–479.Google Scholar
Nathan, R. (2005). Long-distance dispersal research: building a network of yellow brick roads. Diversity and Distributions 11, 125–130.CrossRefGoogle Scholar
Nevo, E., Apelbaum-Elkahar, I., Garty, J. and Beiles, A. (1997). Natural selection caused microscale allozyme diversity in wild barley and a lichen at “Evolution Canyon”, Mt. Carmel, Israel. Heredity, 78, 373–382.CrossRefGoogle Scholar
Newmaster, S. T., Bell, F. W. and Vitt, D. H. (1999). The effects of glyphosate and triclopyr on common bryophytes and lichens in northwestern Ontario. Canadian Journal of Forest Research, 29, 1101–1111.CrossRefGoogle Scholar
Newsham, K. K., Low, M. N. R., McLeod, A. R., Greenslade, P. D. and Emmett, B. C. (1997). Ultraviolet-B radiation influences the abundance and distribution of phylloplane fungi on pedunculate oak (Quercus robur). New Phytologist, 136, 287–297.CrossRefGoogle Scholar
Nicholson, T. P., Rudd, B. A. M., Dawson, M., et al. (2001). Design and utility of oligonucleotide gene probes for fungal polyketide synthases. Chemistry and Biology, 8, 57–178.CrossRefGoogle ScholarPubMed
Nieboer, E. and Richardson, D. H. S. (1980). The replacement of the nondescript term ‘heavy metals’ by a biologically and chemically significant classification of metal ions. Environmental Pollution, 1, 3–26.Google Scholar
Nieboer, E. and Richardson, D. H. S. (1981). Lichens as monitors of atmospheric deposition. In Atmospheric Pollutants in Natural Waters, ed. Eisenreich, S. J., pp. 339–388. Ann Arbor: Ann Arbor Science.Google Scholar
Nieboer, E., MacFarlane, J. D. and Richardson, D. H. S. (1984). Modification of plant cell buffering capacities by gaseous air pollutants. In Gaseous Air Pollutants and Plant Metabolism, ed. Kozsol, M. J. and Whatley, F. R., pp. 313–330. London: Butterworths.Google Scholar
Nieboer, E., Richardson, D. H. S., Lavoie, P. and Padovan, D. (1979). The role of metal-ion binding in modifying the toxic effects of sulphur dioxide on the lichen Umbilicaria muhlenbergii. I. Potassium efflux studies. New Phytologist, 82, 621–632.CrossRefGoogle Scholar
Nieboer, E., Richardson, D. H. S., Puckett, K. J. and Tomassini, F. D. (1976). The phytotoxicity of sulphur dioxide in relation to measurable responses in lichens. In Effects of Air Pollutants on Plants, ed. Mansfield, T. A., pp. 61–85. Cambridge: Cambridge University Press.Google Scholar
Nieboer, E., Richardson, D. H. S. and Tomassini, F. D. (1978). Mineral uptake and release by lichens: an overview. Bryologist, 81, 226–246.CrossRefGoogle Scholar
Nieboer, E., Tomassini, F. D., Puckett, K. J. and Richardson, D. H. S. (1977). A model for the relationship between gaseous and aqueous concentrations of sulphur dioxide in lichen exposure studies. New Phytologist, 79, 157–162.CrossRefGoogle Scholar
Nikonov, A. A. and Shebalina, T. Y. (1979). Lichenometry and earthquake age determination in central Asia. Nature, 280, 675–677.CrossRefGoogle Scholar
Nimis, P. L. (1996). Towards a checklist of Mediterranean lichens. Bocconea, 6, 5–17.Google Scholar
Nimis, P. L. and Martellos, S. (2003). A second checklist of the lichens of Italy, with a thesaurus of synonyms. Museo Regionale di Scienze Naturali Monografie, 4, 1–192.Google Scholar
Nimis, P. L. and Poelt, J. (1987). The lichens and lichenicolous fungi of Sardinia (Italy). An annotated list. Studia Geobotanica Trieste, 7 (suppl.1), 1–269.Google Scholar
Nimis, P. L., Castello, M. and Perotti, M. (1990). Lichens as biomonitors of sulphur dioxide pollution in La Spezia (Northern Italy). Lichenologist, 22, 333–344.CrossRefGoogle Scholar
Nimis, P. L., Pinna, D. and Salvadori, O. (1992). Licheni e Conservazione dei Monumenti. Bologna: Cooperativa Libraria Universitaria Editrice Bologna.Google Scholar
Nimis, P. L., Scheidegger, C. and Wolseley, P. A., eds. (2002). Monitoring with Lichens – Monitoring Lichens. Nato Science Series IV: Earth and Environmental Sciences 7. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Noctor, G. and Foyer, C. (1998). Ascorbate and glutathione: keeping active oxygen under control. Annual Review of Plant Physiology and Plant Molecular Biology, 49, 249–279.CrossRefGoogle ScholarPubMed
Noeske, O., Läuchli, A., Lange, O. L., Vieweg, G. H. and Ziegler, H. (1970). Konzentration und Kokalisierung von Schwermetallen in Flechten der Erzschlackenhalden des Harzes. Berichte der deutschen botanischen Gesellschaft, 4, 67–79.Google Scholar
Nordberg, M. L. and Allard, A. (2002). A remote sensing methodology for monitoring cover. Canadian Journal of Remote Sensing, 28, 262–274.CrossRefGoogle Scholar
Notcutt, G. and Davies, F. (1993). Dispersion of gaseous volcanogenic fluoride, island of Hawaii. Journal of Volcanology and Geothermal Research, 56, 125–131.CrossRefGoogle Scholar
Notcutt, G. and Davies, F. (1999). Biomonitoring of volcanogenic fluoride, Furnas Caldera, Sao Miguel, Azores. Journal of Volcanology and Geothermal Research, 92, 209–214.CrossRefGoogle Scholar
Nriagu, J. O. and Pacyna, J. (1988). Quantitative assessment of worldwide contamination of air, water and soils by trace metals. Nature, 333, 134–139.CrossRefGoogle ScholarPubMed
Nyati, S. (2006). Photobiont diversity in Teloschistaceae. Ph.D. thesis, Mathematisch-Naturwissenschaftliche Fakultät, Universität Zürich.
Nybakken, L., Solhaug, K. A., Bilger, W. and Gauslaa, Y. (2004). The lichens Xanthoria elegans and Cetraria islandica maintain a high protection against UV-B radiation in Arctic habitats. Oecologia, 140, 211–216.CrossRefGoogle ScholarPubMed
Nylander, W. (1866). Circa novum in studio Lichenum criterium chemicum. Flora, 49, 198–201.Google Scholar
Obermayer, W. and Poelt, J. (1992). Contributions to the knowledge of the lichen flora of the Himalayas. III. On Lecanora somervellii Paulson (lichenised Ascomycotina, Lecanoraceae). Lichenologist, 24, 111–117.Google Scholar
Oberwinkler, F. (1984). Fungus-alga interactions in basidiolichens. Nova Hedwigia, 79, 739–774.Google Scholar
Oberwinkler, F. (2001). Basidiolichens. In Fungal Associations, The Mycota, Vol. IX, ed. Hock, B., pp. 211–225. Berlin: Springer.CrossRefGoogle Scholar
O'Brien, H., Miadlikowska, J. and Lutzoni, F. (2005). Assessing host specialization in symbiotic cyanobacteria associated with four closely related species of the lichen fungus Peltigera. European Journal of Phycology, 40, 363–378.CrossRefGoogle Scholar
Ochiai, E. (1977). Bioinorganic Chemistry: An Introduction. Boston: Allyn and Bacon.Google Scholar
Ockinger, E., Niklasson, M. and Nilsson, S. G. (2005). Is local distribution of the epiphytic lichen Lobaria pulmonaria limited by dispersal capacity or habitat quality?Biodiversity and Conservation, 14, 759–773.CrossRefGoogle Scholar
Ogren, E. (1993). Convexity of the photosynthetic light-response curve in relation to intensity and direction of light during growth. Plant Physiology, 101, 1013–1019.CrossRefGoogle ScholarPubMed
O'Hare, G. P. and Williams, P. (1975). Some effects of sulphur dioxide flow on lichens. Lichenologist, 7, 116–120.CrossRefGoogle Scholar
Ohmura, Y., Kawachi, M., Kasai, F. and Watanabe, M. (2006). Genetic combinations of symbionts in a vegetatively reproducing lichen, Parmotrema tinctorum, based on ITS rDNA sequences. Bryologist, 109, 43–59.CrossRefGoogle Scholar
Olafsen, A. G. (1989). Nitrogen and carbon fixation in two Arctic lichens, Stereocaulon tomentosum and Peltigera canina. M.S. thesis. Tempe: Arizona State University.
Oliver, A. E., Hinchab, D. K. and Crowe, J. H. (2002). Looking beyond sugars: the role of amphiphilic solutes in preventing adventitious reactions in anhydrobiotes at low water contents. Comparative Biochemistry and Physiology Part A, 131, 515–525.CrossRefGoogle ScholarPubMed
Oliver, A. E., Leprince, O., Wolkers, W. W., et al. (2001). Non-disaccharide-based mechanisms of protection during drying. Cryobiology, 43, 151–167.CrossRefGoogle ScholarPubMed
Olmez, I., Gulovali, M. C. and Gordon, G. E. (1985). Trace element concentrations in lichens near a coal-fired power plant. Atmospheric Environment, 19, 1663–1669.CrossRefGoogle Scholar
O'Neill, A. L. (1994). Reflectance spectra of microphytic soil crusts in semi-arid Australia. International Journal of Remote Sensing, 15, 675–681.CrossRefGoogle Scholar
Öquist, G., Brunes, L. and Hällgren, J.-E. (1982). Photosynthetic efficiency of Betula pendula acclimated to different quantum flux densities. Plant Cell and Environment, 5, 9–15.Google Scholar
Orange, A., James, P. W. and White, F. J. (2001). Microchemical Methods for the Identification of Lichens. London: British Lichen Society.Google Scholar
Ott, S. (1987 a). The juvenile development of lichen thalli from vegetative diaspores. Symbiosis, 3, 57–74.Google Scholar
Ott, S. (1987 b). Reproductive strategies in lichens. Bibliotheca Lichenologica, 25, 81–93.Google Scholar
Ott, S. and Zvoch, I. (1992). Ethylene production by lichens. Lichenologist, 24, 73–80.Google Scholar
Ott, S., Krieg, T., Spanier, U. and Schieleit, P. (2000 a). Phytohormones in lichens with emphasis on ethylene biosynthesis and functional aspects on lichen symbiosis. Phyton, 40, 83–94.Google Scholar
Ott, S., Schröder, T. and Jahns, H. M. (2000 b). Colonization strategies and interactions of lichens on twigs. Bibliotheca Lichenologica, 75, 445–455.Google Scholar
Ott, S., Treiber, K. and Jahns, H. M. (1993). The development of regenerative thallus structures. Botanical Journal of the Linnean Society, 113, 61–76.CrossRefGoogle Scholar
Otte, V., Esslinger, T. L. and Litterski, B. (2002). Biogeographical research on European species of the lichen genus Physconia. Journal of Biogeography, 29, 1125–1141.CrossRefGoogle Scholar
Otte, V., Esslinger, T. L. and Litterski, B. (2005). Global distribution of the European species of the lichen genus Melanelia Essl. Journal of Biogeography, 32, 1221–1241.CrossRefGoogle Scholar
Øvstedal, D. O. and Smith, Lewis R. I. (2001). The Lichens of Antarctica and South Georgia: A Guide to their Identification and Ecology. Cambridge: Cambridge University Press.Google Scholar
Pakarinen, P. and Häsänen, E. (1983). Mercury concentrations of bog mosses and lichens. Suo, 34, 17–20.Google Scholar
Palacios, D., Parrilla, G. and Zamorano, J. J. (1999). Paraglacial and postglacial debris flows on a Little Ice Age terminal moraine: Jamapa Glacier, Pico de Orizaba (Mexico). Geomorphology, 28, 95–118.CrossRefGoogle Scholar
Palice, Z. and Printzen, C. (2004). Genetic variability in tropical and temperate populations of Trapeliopsis glaucolepidea: evidence against long-range dispersal in a lichen with a disjunct distribution. Mycotaxon, 90, 43–54.Google Scholar
Palmer, H. E., Hanson, W. C., Griffin, B. I. and Roesch, W. C. (1963). Cesium-137 in Alaskan eskimos. Science, 142, 64–66.CrossRefGoogle ScholarPubMed
Palmer, R. J. Jr. and Friedmann, E. I. (1990). Water relations, thallus structure and photosynthesis in Negev Desert lichens. New Phytologist, 116, 597–603.CrossRefGoogle ScholarPubMed
Palmqvist, K. (1993). Photosynthetic CO2 use efficiency in lichens and their isolated photobionts: the possible role of a CO2 concentrating mechanism in cyanobacterial lichens. Planta, 191, 48–56.CrossRefGoogle Scholar
Palmqvist, K. (2000). Carbon economy in lichens. New Phytologist, 148, 11–36.CrossRefGoogle Scholar
Palmqvist, K. (2002). Carbon metabolism in cyanobacterial lichens. In Cyanobacteria in Symbiosis, ed. Rai, A. N., and, B. BergmanRasmussen, U., pp. 73–96. Amsterdam: Kluwer Academic.CrossRefGoogle Scholar
Palmqvist, K. and Dahlman, L. (2006). Responses of the green algal foliose lichen Platismatia glauca to increased nitrogen supply. New Phytologist, 171, 343–356.CrossRefGoogle ScholarPubMed
Palmqvist, K. and Sundberg, B. (2000). Light use efficiency of dry matter gain in five macro-lichens: relative impact of microclimate and species-specific traits. Plant Cell and Environment, 23, 1–14.CrossRefGoogle Scholar
Palmqvist, K., los Ríos, A., Ascaso, C. and Samuelsson, G. (1997). Photosynthetic carbon acquisition in the lichen photobionts Coccomyxa and Trebouxia (Chlorophyta). Physiologia Plantarum, 101, 67–76.CrossRefGoogle Scholar
Palmqvist, K., Campbell, D., Ekblad, A. and Johansson, H. (1998). Photosynthetic capacity in relation to nitrogen content and its partitioning in lichens with different photobionts. Plant, Cell and Environment, 21, 361–372.CrossRefGoogle Scholar
Palmqvist, K., Dahlman, L., Valladares, F., et al. (2002). CO2 exchange and thallus nitrogen across 75 contrasting lichen associations from different climate zones. Oecologia, 133, 295–306.CrossRefGoogle ScholarPubMed
Palmqvist, K., Samuelsson, G. and Badger, M. R. (1994). Photobiont-related differences in carbon acquisition among green-algal lichens. Planta, 195, 70–79.CrossRefGoogle Scholar
Pannewitz, S., Green, T. G. A., Schlensog, M., et al. (2006). Photosynthetic performance of Xanthoria mawsonii C. W. Dodge in coastal habitats, Ross Sea region, continental Antarctica. Lichenologist, 38, 67–81.CrossRefGoogle Scholar
Pannewitz, S., Schlensog, M., Green, T. G. A., Sancho, L. G. and Schroeter, B. (2003). Are lichens active under snow in continental Antarctica?Oecologia, 135, 30–38.CrossRefGoogle ScholarPubMed
Parguey-Leduc, A. and Janex-Favre, M. (1981). The ascocarps of ascohymenial pyrenomycetes. In Ascomycete Systematics: The Luttrellian Concept, ed. Reynolds, D. R., pp. 102–123. New York: Springer.CrossRefGoogle Scholar
Parmasto, E. (2004). Integrating molecular and morphological data in the systematics of fungi. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 5. Tartu: Tartu University Press.Google Scholar
Peake, J. F. and James, P. W. (1967). Lichens and mollusca. Lichenologist, 3, 425–428.CrossRefGoogle Scholar
Pearce, C. (1997). Biologically active fungal metabolites. Advances in Applied Microbiology, 44, 1–80.CrossRefGoogle ScholarPubMed
Pearson, L. C. and Henriksson, E. (1981). Air pollution damage to cell membranes in lichens. II. Laboratory experiments. Bryologist, 84, 515–520.CrossRefGoogle Scholar
Pearson, L. C. and Skye, E. (1965). Air pollution affects patterns of photosynthesis in Parmelia sulcata, a corticolous lichen. Science, 148, 1600–1602.CrossRefGoogle Scholar
Peiser, G. and Yang, S. F. (1985). Biochemical and physiological effects of SO2 on nonphotosynthetic processes in plants. In Sulfur Dioxide and Vegetation: Physiology, Ecology, and Policy Issues, ed. Winner, W. E., Mooney, H. A. and Goldstein, R. A., pp. 148–161. Stanford: Stanford University Press.Google Scholar
Pennycook, S. R. and Galloway, D. J. (2004). Checklist of New Zealand “Fungi”. In Fungi of New Zealand, Vol. 1: Introduction to Fungi of New Zealand, ed. McKenzie, E. H. C., pp. 401–488. Hong Kong: Fungal Diversity Press.Google Scholar
Pérez, F. L. (1994). Vagrant cryptogams in a paramo of the high Venezuelan Andes. Flora, 189, 263–276.CrossRefGoogle Scholar
Perkins, D. F. (1992). Relationship between fluoride contents and loss of lichens near an aluminium works. Water, Air, and Soil Pollution, 64, 503–510.CrossRefGoogle Scholar
Perkins, D. F. and Millar, R. O. (1987). Effects of airborne fluoride emissions near an aluminium works in Wales: part 2 – saxicolous lichens growing on rocks and walls. Environmental Pollution, 48, 185–196.CrossRefGoogle ScholarPubMed
Peršoh, D. (2004). Diversity of lichen inhabiting fungi in the Letharietum vulpinae. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 34. Tartu: Tartu University Press.Google Scholar
Peršoh, D., Beck, A. and Rambold, G. (2004). The distribution of ascus types and photobiontal selection in Lecanoromycetes (Ascomycota) against the background of a revised SSU nrDNA phylogeny. Mycological Progress, 3, 103–121.CrossRefGoogle Scholar
Peterson, E. B. (2000). An overlooked fossil lichen (Lobariaceae). Lichenologist, 32, 298–300.CrossRefGoogle Scholar
Peterson, E. B. and McCune, B. (2003). The importance of hotspots for lichen diversity in forests of western Oregon. Bryologist, 106, 246–256.CrossRefGoogle Scholar
Peterson, R. B. and Burris, R. H. (1976). Conversion of acetylene reduction rates in natural populations of blue-green algae. Analytical Biochemistry, 73, 404–410.CrossRefGoogle ScholarPubMed
Petrini, O., Hake, U. and Dreyfuss., M. M. (1990). An analysis of fungal communities isolated from fruticose lichens. Mycologia, 82, 444–451.CrossRefGoogle Scholar
Petzold, D. E. and Goward, S. N. (1988). Reflectance spectra of subarctic lichens. Remote Sensing of Environment, 24, 481–492.CrossRefGoogle Scholar
Pfanz, H., Martinoia, E, Lange, O. L. and Heber, U. (1987). Flux of SO2 into leaf cells and cellular acidification by SO2. Plant Physiology, 85, 928–933.CrossRefGoogle Scholar
Piercey-Normore, M. D. (2004). Selection of algal genotypes by three species of lichen fungi in the genus Cladonia. Canadian Journal of Botany, 82, 947–961.CrossRefGoogle Scholar
Piercey-Normore, M. D. (2006). The lichen-forming ascomycete Evernia mesomorpha associates with multiple genotypes of Trebouxia jamesii. New Phytologist, 169, 331–344.CrossRefGoogle ScholarPubMed
Piercey-Normore, M. D. and DePriest, P. T. (2001). Algal switching among lichen symbioses. American Journal of Botany, 88, 1490–1498.CrossRefGoogle ScholarPubMed
Pike, L. H. (1978). The importance of epiphytic lichens in mineral cycling. Bryologist, 81, 247–257.CrossRefGoogle Scholar
Pike, L. H. (1981). Estimation of lichen biomass and production with special reference to the use of ratios. In The Fungal Community: Its Organization and Role in the Ecosystem, ed. Wicklow, D. and Carroll, G., pp. 533–552. New York: Marcel Dekker.Google Scholar
Pinna, D., Salvadori, O. and Tretiach, M. (1998). An anatomical investigation of calcicolous endolithic lichens from the Trieste karst (NE Italy). Plant Biosystems, 132, 183–195.CrossRefGoogle Scholar
Pintado, A. and Sancho, L. G. (2002). Ecological significance of net photosynthesis activation by water vapour uptake in Ramalina capitata from rain-protected habitats in central Spain. Lichenologist, 34, 403–413.CrossRefGoogle Scholar
Pišút, I. and Lisicka, E. (1985). A study of cryptogamic epiphytes on an oak trunk in the vicinity of Bratislava in the years 1973–1983. Ekologia (CSSR), 4, 225–234.Google Scholar
Plakunova, O. V. and Plakunova, V. G. (1987). Ultrastructure of Cladina stellaris lichen components in normal conditions and at SO2 environment pollution. Izvestiya Akademii Nauk SSR, Seriya Biologicheskaya, 1987, 361–369.Google Scholar
Platt, J. L. and Spatafora, J. W. (2000). Evolutionary relationships of nonsexual lichenized fungi: molecular phylogenetic hypotheses for the genera Siphula and Thamnolia from SSU and LSU rDNA. Mycologia, 92, 475–487.CrossRefGoogle Scholar
Plessl, A. (1963). Über die Beziehungen von Pilz und Alge im Flechtenthallus. Österreichische Botanische Zeitschrift, 110, 194–269.CrossRefGoogle Scholar
Poelt, J. (1970). Das Konzept der Artenpaare bei den Flechten. Vorträge aus dem Gesamtgebeit der Botanik herausgegeben von der deutschen botanischen Gesellschaft, Berlin, 4, 187–198.Google Scholar
Poelt, J. (1972). Die Taxonomische Behandlung von Artenpaaren bei den Flechten. Botaniska Notiser, 125, 77–81.Google Scholar
Poelt, J. (1980). Physcia opuntiella und die Lebensform der sprossenden Flechten. Flora, 169, 22–31.CrossRefGoogle Scholar
Poelt, J. (1985). Über auf Moosen parasitierende Flechten. Sydowia, Annales Mycologici Series II, 38, 241–254.Google Scholar
Poelt, J. (1986). Morphologie der Flechten – Fortschritte und Probleme. Berichte der deutschen botanischen Gesellschaft, 99, 3–29.Google Scholar
Poelt, J. (1989). Die Entstehung einer Strauchflechte aus einem Formenkreis krustiger Verwandter. Flora, 183, 65–72.CrossRefGoogle Scholar
Poelt, J. (1993). La riproduzione asessuale nei licheni. Notiziario della Società Lichenologica Italiana, 6, 9–28.Google Scholar
Poelt, J. (1994). On lichenized asexual diaspores in foliose lichens – a contribution towards a more differentiated nomenclature (Lichens, Lecanorales). Cryptogamic Botany, 5, 150–162.Google Scholar
Poelt, J. and Doppelbauer, H. (1956). Über parasitische Flechten. Planta, 46, 467–480.CrossRefGoogle Scholar
Poelt, J. and Mayrhofer, H. (1988). Über Cyanotrophie bei Flechten. Plant Systematics and Evolution, 158, 265–281.CrossRefGoogle Scholar
Poelt, J. and Obermayer, W. (1990 a). Über Thallosporen bei einigen Krustenflechten. Herzogia, 8, 273–288.Google Scholar
Poelt, J. and Obermayer, W. (1990 b). Lichenisierte Bulbillen als Diasporen bei der Basidiolichene Multiclavula vernalis spec. coll. Herzogia, 6, 289–294.Google Scholar
Poelt, J. and Vězda, A. (1990). Über kurzlebige Flechten. Bibliotheca Lichenologica, 38, 377–394.Google Scholar
Pöggeler, S. (1999). Phylogenetic relationships between mating-type sequences from homothallic and heterothallic ascomycetes. Current Genetics, 36, 222–231.Google ScholarPubMed
Poinar, G., Peterson, E. and Platt, J. (2000). Fossil Parmelia in New World amber. Lichenologist, 32, 263–269.CrossRefGoogle Scholar
Pole, M. (1993). Keeping in touch: vegetation prehistory on both sides of the Tasman. Australian Systematic Botany, 6, 387–397.CrossRefGoogle Scholar
Pole, M. (1994). The New Zealand flora – entirely long-distance dispersal?Journal of Biogeography, 21, 625–635.CrossRefGoogle Scholar
Ponzetti, J. M. and McCune, B. P. (2001). Biotic soil crusts of Oregon's shrub steppe: community composition in relation to soil chemistry, climate, and livestock activity. Bryologist, 104, 212–225.CrossRefGoogle Scholar
Popp, M. and Smirnoff, N. (1995). Polyol accumulation and metabolism during water deficit. In Environment and Plant Metabolism, ed. Smirnoff, N., pp. 199–215. Oxford: BIOS Scientific Publishers.Google Scholar
Potts, M. (1994). Desiccation tolerance in prokaryotes. Microbiology Reviews, 58, 755–805.Google ScholarPubMed
Potts, M. and Bowman, M. (1985). Sensitivity of Nostoc commune UTEX 584 (Cyanobacteria) to water stress. Archiv für Microbiologie, 141, 51–56.CrossRefGoogle Scholar
Pressel, S., Ligrone, R. and Duckett, J. G. (2006). The effects of de- and rehydration on food-conducting cells in the moss Polytrichum formosum Hedw.: a cytological study. Annals of Botany, 98, 67–76.CrossRefGoogle Scholar
Price, S. and Long, S. P. (1989). An in vivo analysis of the effect of SO2 fumigation on photosynthesis in Zea mays. Physiologica Plantarum, 76, 193–200.CrossRefGoogle Scholar
Prillinger, H. (1982). Zur genetischen Kontrolle und Evolution der sexuellen Fortplanzung und Heterothallie der Chitinpilzen. Zeitschrift für Mykologie, 48, 297–324.Google Scholar
Printzen, C. and Ekman, S. (2002). Genetic variability and its geographical distribution in the widely disjunct Cavernularia hultenii. Lichenologist, 34, 101–111.CrossRefGoogle Scholar
Printzen, C. and Ekman, S. (2003). Local population subdivision in the lichen Cladonia subcervicornis as revealed by mitochondrial cytochrome oxidase subunit 1 intron sequences. Mycologia, 95, 399–406.CrossRefGoogle ScholarPubMed
Printzen, C. and Kantvilas, G. (2004). Hertelidea, genus novum Stereocaulaearum (Ascomycetes lichenisati). Bibliotheca Lichenologica, 88, 539–553.Google Scholar
Printzen, C. and Lumbsch, H. T. (2000). Molecular evidence for the diversification of extant lichens in the late Cretaceous and Tertiary. Molecular Phylogenetics and Evolution, 17, 379–387.CrossRefGoogle ScholarPubMed
Printzen, C., Ekman, S. and Tønsberg, T. (2003). Phylogeography of Cavernularia hultenii: evidence of slow genetic drift in a widely disjunct lichen. Molecular Ecology, 12, 1473–1486.CrossRefGoogle Scholar
Proctor, M. C. F., Ligrone, R. and Duckett, J. G. (2006). Desiccation tolerance in the moss Polytrichum formosum Hedw.: physiological and fine-structural changes during desiccation and recovery. Annals of Botany, 99, 75–93.CrossRefGoogle Scholar
Puckett, K. J. (1976). The effect of heavy metals in some aspects of lichen physiology. Canadian Journal of Botany, 54, 2695–2703.CrossRefGoogle Scholar
Puckett, K. J. (1978). Element Levels in Lichens from the Northwest Territories. Report ARQA–56–76. Downsview: Atmospheric Environment Service, Environment Canada.Google Scholar
Puckett, K. J. (1988). Bryophytes and lichens as monitors of metal deposition. Bibliotheca Lichenologica, 30, 231–267.Google Scholar
Puckett, K. J. and Burton, M. A. S. (1981). The effect of trace elements on lower plants. In Effect of Heavy Metal Pollution on Plants, Vol. 2: Metals in the Environment, ed. Lepp, N. W., pp. 213–238. London: Applied Science Publishers.Google Scholar
Puckett, K. J. and Finegan, E. J. (1980). An analysis of the element content of lichens from the Northwest Territories, Canada. Canadian Journal of Botany, 58, 2073–2089.CrossRefGoogle Scholar
Puckett, K. J., Nieboer, E., Flora, W. P. and Richardson, D. H. S. (1973). Sulphur dioxide: its effect on photosynthetic 14C fixation in lichens and suggested mechanisms of phytotoxicity. New Phytologist, 72, 141–154.CrossRefGoogle Scholar
Puckett, K. J., Richardson, D. H. S., Flora, W. P. and Nieboer, E. (1974). Photosynthetic 14C fixation by the lichen Umbilicaria muhlenbergii (Ach.) Tuck. following short exposures to aqueous sulphur dioxide. New Phytologist, 73, 1183–1192.CrossRefGoogle Scholar
Puckett, K. J., Tomassini, F. D., Nieboer, E. and Richardson, D. H. S. (1977). Potassium efflux by lichen thalli following exposure to aqueous sulphur dioxide. New Phytologist, 79, 135–145.CrossRefGoogle Scholar
Punz, W. (1979). Der Einfluss isolierter und kombinierter Schadstoffe auf die Flechtenphotosynthese. Photosynthetica, 13, 428–433.Google Scholar
Purvis, A., Gittleman, J. L. and Brooks, T., eds. (2005). Phylogeny and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Purvis, O. W. (1984). The occurrence of copper oxalate in lichens growing on copper sulphide-bearing rocks in Scandinavia. Lichenologist, 16, 197–204.CrossRefGoogle Scholar
Purvis, O. W. (2000). Lichens. London: Natural History Museum and Washington: Smithsonian Institution.Google Scholar
Purvis, O. W., Coppins, B. J., Hawksworth, D. J., James, P. W. and Moore, M. D. (1992). The Lichen Flora of Great Britain and Ireland. London: Natural History Publications.Google Scholar
Purvis, O. W., Elix, J. A.Broomhead, J. A. and Jones, G. C. (1987). The occurrence of copper-norstictic acid in lichens from cupriferous substrata. Lichenologist, 19, 193–203.CrossRefGoogle Scholar
Qui, B. S. and Gao, K. S. (2001). Photosynthetic characteristics of the terrestrial blue-green alga Nostoc flagelliforme. European Journal of Phycology, 36, 147–156.Google Scholar
Quilhot, W., Fernández, E., Rubio, C., Goddard, M. and Hidalgo, M. E. (1998). Lichen secondary products and their importance in environmental studies. In Lichenology in Latin America: History, Current Knowledge and Applications, ed. Marcelli, M. P. and Seaward, M. R. D., pp. 171–179. São Paulo: CETESB.Google Scholar
Quispel, A. (1960). Respiration of lichens. In Pflanzenatmung einschliesslich Gärung und Säurestoffwechsel (Handbuch der Pflanzenphysiologie vol XII/2), ed. Wolf, J., pp. 455–460. Berlin: Springer.Google Scholar
Rai, A. N. (1988). Nitrogen metabolism. In Handbook of Lichenology, Vol. I, ed. Galun, M., pp. 201–237. Boca Raton: CRC Press.Google Scholar
Rai, A. N. (2002). Cyanolichens: nitrogen metabolism. In Cyanobacteria in Symbiosis, ed. Rai, A. N., Bergman, B. and Rasmussen, U., pp. 97–115. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Rai, A. N., Rowell, P. and Stewart, W. D. P. (1980). NH4+ assimilation and nitrogenase regulation in the lichen Peltigera aphthosa Willd. New Phytologist, 85, 545–555.CrossRefGoogle Scholar
Rai, A. N., Rowell, P. and Stewart, W. D. P. (1981). Nitrogenase activity and dark CO2 fixation in the lichen Peltigera aphthosa Willd. Planta, 151, 256–264.CrossRefGoogle ScholarPubMed
Rai, A. N., Rowell, P. and Stewart, D. P. (1983). Mycobiont-cyanobiont interactions during dark nitrogen fixation by the lichen Peltigera aphthosa. Physiologia Plantarum, 57, 285–290.CrossRefGoogle Scholar
Rambold, G. and Triebel, D. (1992). The inter-lecanoralean associations. Bibliotheca Lichenologica, 48, 3–201.Google Scholar
Rambold, G., Friedl, T. and Beck, A. (1998). Photobionts in lichens: possible indicators of phylogenetic relationships?Bryologist, 101, 392–397.CrossRefGoogle Scholar
Ramstad, S. and Hestmark, G. (2001). Population structure and size-dependent reproductive effort in Umbilicaria sporochroa. Mycologia, 93, 453–458.CrossRefGoogle Scholar
Rancan, F., Rosan, S., Boehm, K., et al. (2002). Protection against UVB irradiation by natural filters extracted from lichens. Journal of Photochemistry and Photobiology B: Biology, 68, 133–139.CrossRefGoogle ScholarPubMed
Randlane, T. and Saag, A. (1998). Synopsis of the genus Nephromopsis (Fam. Parmeliaceae, lichenized Ascomycota). Cryptogamie, Bryologie et Lichénologie, 19, 175–191.Google Scholar
Randlane, T., Saag, A. and Thell, A. (1997). A second updated world list of cetrarioid lichens. Bryologist, 100, 109–122.CrossRefGoogle Scholar
Rao, D. N. and LeBlanc, F. (1966). Effects of sulfur dioxide on the lichen alga with special reference to chlorophyll. Bryologist, 69, 69–75.CrossRefGoogle Scholar
Raven, J. A. (1992). Energy and nutrient acquisition by autotrophic symbioses and their asymbiotic ancestors. (Review). Symbiosis, 14, 33–60.Google Scholar
Raven, J. A., Johnston, A. M., Handley, L. L. and McInroy, S. G. (1990). Transport and assimilation of inorganic carbon by Lichina pygmea under emersed and submersed conditions. New Phytologist, 114, 407–417.CrossRefGoogle Scholar
Reeb, V., Lutzoni, F. and Roux, C. (2004). Contribution of RPB2 to multilocus phylogenetic studies of the euascomycetes (Pezizomycotina, Fungi) with special emphasis on the lichen-forming Acarosporaceae and evolution of polyspory. Molecular Phylogenetics and Evolution, 32, 1036–1060.CrossRefGoogle ScholarPubMed
Reich, P. B., Ellsworth, D. S., Walters, M. B., et al. (1999). Generality of leaf trait relationships: a test across six biomes. Ecology, 80, 1955–1969.CrossRefGoogle Scholar
Reich, P. B., Walters, M. B., Ellsworth, D. S., et al. (1998). Relationships of leaf dark respiration to leaf nitrogen, specific leaf area and leaf life-span: a test across biomes and functional groups. Oecologia, 114, 471–482.CrossRefGoogle ScholarPubMed
Reiners, W. A. and Olson, R. K. (1984). Effects of canopy components on throughfall chemistry: an experimental analysis. Oecologia, 63, 320–330.CrossRefGoogle ScholarPubMed
Reinke, J. (1894–1896). Abhandlungen über Flechten. Jahrbücher wissenschaftliche Botanik, 26, 28, 29.Google Scholar
Reisser, W. (1992). Endosymbiotic associations of algae with freshwater protozoa and invertebrates. In Algae and Symbioses: Plants, Animals, Fungi, Viruses, Interactions Explored, ed. Reisser, W., pp. 1–19. Bristol: Biopress.Google Scholar
Reiter, R. and Türk, R. (2000 a). Investigations on the CO2 exchange of lichens in the alpine belt. I. Comparative patterns of net CO2 exchange in Cladonia mitis, Thamnolia vermicularis and Umbilicaria cylindrica. Bibliotheca Lichenologica, 75, 333–351.Google Scholar
Reiter, R. and Türk, R. (2000 b). Investigations on the CO2 exchange of lichens in the alpine belt. II. Comparative patterns of net CO2 exchange in Cetraria islandica and Flavocetraria nivalis. Phyton (Austria), 40, 161–177.Google Scholar
Renhorn, K. E., Esseen, P.-A., Palmqvist, K. and Sundberg, B. (1997). Growth and vitality of epiphytic lichens. I. Responses to microclimate along a forest edge-interior gradient. Oecologia, 109, 1–9.CrossRefGoogle Scholar
Rennenberg, H. (1984). The fate of excess sulfur in higher plants. Annual Review of Plant Physiology, 35, 121–153.CrossRefGoogle Scholar
Rennenberg, H. and Polle, A. (1994). Metabolic consequences of atmospheric sulphur influx into plants. In Plant Responses to the Gaseous Environment, ed. Alscher, R. G. and Wellburn, A. R., pp. 165–180. London: Chapman and Hall.CrossRefGoogle Scholar
Reutimann, P. and Scheidegger, C. (1987). Importance of lichen secondary products in food choice of two oribatic mites (Acari) in an alpine meadow ecosystem. Journal of Chemical Ecology, 13, 363–369.CrossRefGoogle Scholar
Rhoades, F. M. (1977). Growth rates of Lobaria oregana as determined from sequential photographs. Canadian Journal of Botany, 55, 2226–2233.CrossRefGoogle Scholar
Rhoades, F. M. (1981). Biomass of epiphytic lichens and bryophytes on Abies lasiocarpa on a Mt. Baker lava flow. Bryologist, 84, 39–47.CrossRefGoogle Scholar
Rhoades, F. M. (1983). Distribution of thalli in a population of the epiphytic lichen Lobaria oregana and a model of population dynamics and production. Bryologist, 86, 309–331.CrossRefGoogle Scholar
Richardson, D. H. S. (1974). The Vanishing Lichens. New York: Hafner Press.Google Scholar
Richardson, D. H. S. (1988). Medicinal and other economic aspects of lichens. In Handbook of Lichenology, Vol. 3, ed. Galun, M., pp. 93–108. Boca Raton: CRC Press.Google Scholar
Richardson, D. H. S. (1991). Lichens and man. In Frontiers in Mycology, ed. Hawksworth, D. L., pp. 187–210. Kew: CAB International.Google Scholar
Richardson, D. H. S. (1999). War in the world of lichens: parasitism and symbiosis as exemplified by lichens and lichenicolous fungi. Mycological Research, 6, 641–650.CrossRefGoogle Scholar
Richardson, D. H. S. and Cameron, R. P. (2004). Cyanolichens: their response to pollution and possible management strategies for their conservation in northeastern North America. Northeastern Naturalist, 11, 1–22.CrossRefGoogle Scholar
Richardson, D. H. S. and Nieboer, E. (1983). Ecophysiological responses of lichens to sulphur dioxide. Journal of the Hattori Botanical Laboratory, 54, 331–351.Google Scholar
Richardson, D. H. S. and Smith, D. C. (1966). The physiology of the symbiosis in Xanthoria aureola (Ach.) Erichs. Lichenologist, 3, 202–206.CrossRefGoogle Scholar
Richardson, D. H. S. and Smith, D. C. (1968). Lichen physiology. IX. Carbohydrate movement from the Trebouxia symbiont of Xanthoria aureola. New Phytologist, 67, 61–68.CrossRefGoogle Scholar
Richardson, D. H. S. and Young, C. M. (1977). Lichens and vertebrates. In Lichen Ecology, ed. Seaward, M. R. D., pp. 121–144. London: Academic Press.Google Scholar
Richardson, D. H. S., Beckett, P. J. and Nieboer, E. (1980). Nickel in lichens, bryophytes, fungi and algae. In Nickel in the Environment, ed. Nriagu, J. O., pp. 367–406. New York: John Wiley.Google Scholar
Richardson, D. H. S., Hill, D. J. and Smith, D. C. (1968). Lichen physiology. XI. The role of the alga in determining the pattern of carbohydrate movement between lichen symbionts. New Phytologist, 67, 469–486.CrossRefGoogle Scholar
Richardson, D. H. S., Nieboer, E., Lavoie, P. and Padovan, D. (1979). The role of metal-ion binding in modifying the toxic affects of sulphur dioxide on the lichen Umbilicaria muhlenbergii. II. 14C-fixation studies. New Phytologist, 82, 633–643.CrossRefGoogle Scholar
Richardson, D. H. S., Nieboer, E., Lavoie, P. and Padovan, D. (1984). Anion accumulation by lichens. I. The characteristics and kinetics of arsenate uptake by Umbilicaria muhlenbergii. New Phytologist, 96, 71–82.CrossRefGoogle Scholar
Riddle, B. R. (2005). Is biogeography emerging from its identity crisis?Journal of Biogeography, 32, 185–186.CrossRefGoogle Scholar
Riddle, B. R. and Hafner, D. J. (2004). The past and future roles of phylogeography. In Frontiers of Biogeography: New Directions in the Geography of Nature, ed. Lomolino, M. V. and Heaney, L. R., pp. 93–110. Sunderland: Sinauer Associates.Google Scholar
Rikkinen, J. (2003). Calicioid lichens from European tertiary amber. Mycologia, 95, 1032–1036.CrossRefGoogle ScholarPubMed
Rikkinen, J., Oksanen, I. and Lohtander, K. (2002). Lichen guilds share related cyanobacterial symbionts. Science, 297, 357.CrossRefGoogle ScholarPubMed
Rindi, F. and Guiry, M. D. (2003). Composition and distribution of subaerial algal assemblages in Galway City, western Ireland. Cryptogamie Algologie, 24, 245–267.Google Scholar
Roberts, B. A. and Thompson, L. K. (1980). Lichens as indicators of fluoride emission from a phosphorous plant, Long Harbour, Newfoundland, Canada. Canadian Journal of Botany, 58, 2218–2228.CrossRefGoogle Scholar
Robinson, C. H. (2001). Cold adaptation in Arctic and Antarctic fungi. New Phytologist, 151, 341–353.CrossRefGoogle Scholar
Rogers, R. W. (1990). Ecological strategies of lichens. Lichenologist, 22, 149–162.CrossRefGoogle Scholar
Rogers, R. W. (1992). Lichen ecology and biogeography. Flora of Australia 54, 30–42.Google Scholar
Rollin, E. M., Milton, E. J. and Roche, P. (1994). The influence of weathering and lichen cover on the reflectance spectra of granitic rocks. Remote Sensing of Environment, 50, 194–199.CrossRefGoogle Scholar
Rolstad, J., Gjerde, I., Storaunet, K. O. and Rolstad, E. (2001). Epiphytic lichens in Norwegian coastal spruce forest: historical logging and present forest structure. Ecological Applications, 11, 421–436.CrossRefGoogle Scholar
Romagni, J. G., Thomas, M. A., Gries, C. and Nash, T. H. III (1997). Sulfite reductase activity in six lichen species as a response to fumigations with sulfur dioxide. Supplement to Plant Physiology, 114, 58.Google Scholar
Romagni, J. G., Thomas, M. A. and Nash, T. H. III (1998). Detoxification of SO2 in lichens: total glutathione. Supplement to Plant Physiology, 115, 89.Google Scholar
Romeike, J., Friedl, T., Helms, G. and Ott, S. (2002). Genetic diversity of algal and fungal partners in four species of Umbilicaria (lichenized ascomycetes) along a transect of the Antarctic peninsula. Molecular Biology and Evolution, 19, 1209–1217.CrossRefGoogle ScholarPubMed
Rorat, T. (2006). Plant dehydrins: tissue location, structure and function. Cell Molecular Biology Letters, 11, 536–556.CrossRefGoogle ScholarPubMed
Rose, C. I. and Hawksworth, D. L. (1981). Lichen recolonization in London's cleaner air. Nature, 289, 289–292.CrossRefGoogle Scholar
Rose, F. (1976). Lichenological indicators of age and environmental quality in woodlands. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 279–307. London: Academic Press.Google Scholar
Rosentreter, R. (1993). Vagrant lichens in North America. Bryologist, 96, 333–338.CrossRefGoogle Scholar
Rosentreter, R. and Eldridge, D. J. (2002). Monitoring biodiversity and ecosystem function: grasslands, deserts, and steppe. In Monitoring with Lichens – Monitoring Lichens. Nato Science Series IV: Earth and Environmental Sciences, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 223–237. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Roser, D. J., Mellick, D. R., Ling, H. U. and Seppelt, R. D. (1992). Polyol and sugar content of terrestrial plants from continental Antarctica. Antarctic Science, 4, 413–420.CrossRefGoogle Scholar
Ross, L. J. (1982). Lichens on coastal live oak in relation to ozone. M. S. Thesis. Arizona State University, Tempe, Arizona.
Ross, L. J. and Nash, T. H. III (1983). Effect of ozone on gross photosynthesis of lichens. Environmental and Experimental Botany, 23, 71–77.CrossRefGoogle Scholar
Rosso, A. L., McCune, B. and Rambo, T. R. (2000). Ecology and conservation of a rare, old-growth-associated canopy lichen in a silvicultural landscape. Bryologist, 103, 117–127.CrossRefGoogle Scholar
Roux, C. (1981). Étude écologique et phytosociologique des peuplements lichéniques saxicoles–calcicoles du sud-est de la France. Bibliotheca Lichenologica, 15, 1–557.Google Scholar
Rubio, C., Fernández, E., Hidalgo, M. E. and Quilhot, W. (2002). Effects of solar UV-B radiation in the accumulation of rhizocarpic acid in a lichen species from alpine zones of Chile. Boletín, Sociedad Chilena de Química, 47, 67–72.CrossRefGoogle Scholar
Ruchty, A., Rosso, A. L. and McCune, B. (2001). Changes in epiphyte communities as the shrub, Acer circinatum, develops and ages. Bryologist, 104, 272–281.CrossRefGoogle Scholar
Rundel, P. W. (1969). Clinal variation in the production of usnic acid in Cladonia subtenuis along light gradients. Bryologist, 72, 40–44.CrossRefGoogle Scholar
Rundel, P. W. (1978). The ecological role of secondary lichen substances. Biochemical Systematics and Ecology, 6, 157–170.CrossRefGoogle Scholar
Rundel, P. W. (1982). Water uptake by organs other than roots. In Physiological Plant Ecology. Vol. II: Water Relations and Carbon Assimilation, ed. Lange, O. L., Nobel, P. S., Osmond, C. B. and Ziegler, H., pp. 111–134. Encyclopedia of Plant Physiology 12B. Berlin: Springer.CrossRefGoogle Scholar
Rundel, P. W. (1988). Water relations. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 17–36. Boca Raton: CRC Press.Google Scholar
Ruoss, E. (1987). Species differentiation in a group of reindeer lichens (Cladonia subg. Cladina). Bibliotheca Lichenologica, 25, 197–206.Google Scholar
Rychert, R. C. and Skujins, J. (1974). Nitrogen fixation by blue-green algae–lichen crusts in the Great Basin Desert. Proceedings of the Soil Science Society of America, 38, 768–771.CrossRefGoogle Scholar
Salisbury, F. B. and Ross, C. W. (1992). Plant Physiology. 4th edn. Belmont: Wadsworth Publishing.Google Scholar
Sancho, L. G. and Kappen, L. (1989). Photosynthesis and water relations and the role of anatomy in Umbilicaricaeae (Lichenes) from central Spain. Oecologia, 81, 473–480.CrossRefGoogle ScholarPubMed
Sancho, L. G. and Pintado, A. (2004). Evidence of high annual growth rate for lichens in the maritime Antarctic. Polar Biology, 27, 312–319.CrossRefGoogle Scholar
Sancho, L. G., Torre, R., Horneck, G., et al. (2007). Lichens survive in space: Results from the 2005 LICHENS experiment. Astrobiology, 7, 443–454.CrossRefGoogle ScholarPubMed
Sancho, L. G., Pintado, A., Blanquer, J. M., Raggio, J. and Vilches, R. (2004). Lichen morphology, thallus water content and photosynthetic performance. Looking for a single trait. In Book of Abstracts of the 5th IAL Symposium: Lichens in Focus, ed. Randlane, T. and Saag, A., p. 47. Tartu: Tartu University Press.Google Scholar
Sancho, L. G., Pintado, A., Green, T. G. A., Pannewitz, S. and Schroeter, B. (2003). Photosynthetic and morphological variation within and among populations of the Antarctic lichen Umbilicaria aprina: implications of the thallus size. Bibliotheca Lichenologica, 86, 299–311.Google Scholar
Sancho, L. G., Pintado, A., Valladares, F., Schroeter, B. and Schlensog, M. (1997). Photosynthetic performance of cosmopolitan lichens in the maritime Antarctic. Bibliotheca Lichenologica, 67, 197–210.Google Scholar
Sancho, L. G., Schroeter, B. and Del-Prado, R. (2000 a). Ecophysiology and morphology of the globular erratic lichen Aspicilia fruticulosa (Eversm.) Flag. from central Spain. Bibliotheca Lichenologica, 75, 137–147.Google Scholar
Sancho, L. G., Valladares, F., Schroeter, B. and Kappen, L. (2000 b). Ecophysiology of Antarctic versus temperate populations of a bipolar lichen: the key role of the photosynthetic partner. In Antarctic Ecosystems: Models for Wider Ecological Understanding, ed. Davison, W., Williams, C. H. and Broady, P., pp. 190–194. Christchurch: New Zealand Natural Sciences Publications.Google Scholar
Sanders, W. B. (1989). Growth and development of the reticulate thallus in the lichen Ramalina menziesii. American Journal of Botany, 76, 666–678.CrossRefGoogle Scholar
Sanders, W. B. (2001 a). Preliminary light microscope observations of fungal and algal colonization and lichen thallus initiation on glass slides placed near foliicolous lichen communities within a lowland tropical forest. Symbiosis, 31, 85–94.Google Scholar
Sanders, W. B. (2001 b). Lichens: the interface between mycology and plant morphology. BioScience, 51, 1025–1035.CrossRefGoogle Scholar
Sanders, W. B. (2005). Observing microscopic phases of lichen life cycles on transparent substrata placed in situ. Lichenologist, 37, 373–382.CrossRefGoogle Scholar
Sanders, W. B. and Lücking, R. (2002). Reproductive strategies, relichenization and thallus development observed in situ in leaf-dwelling lichen communities. New Phytologist, 155, 425–435.CrossRefGoogle Scholar
Sanders, W. B., Moe, R. L. and Ascaso, C. (2004). The intertidal marine lichen formed by the pyrenomycete fungus Verrucaria tavaresiae (Ascomycotina) and the brown alga Petroderma maculiforme (Phaeophyceae): thallus organization and symbiont interaction. American Journal of Botany, 91, 511–522.CrossRefGoogle ScholarPubMed
Sanmartín, I. and Ronquist, F. (2004). Southern Hemisphere biogeography inferred by event-based models: plant versus animal patterns. Systematic Biology, 53, 216–243.CrossRefGoogle ScholarPubMed
Santesson, J. (1969). Chemical studies on lichens. 10. Mass spectrometry on lichens. Arkiv för Chemie, 30, 363–377.Google Scholar
Santesson, R. (1939). Amphibious pyrenolichens I. Arkiv för Botanik, 29A, 1–67.Google Scholar
Santesson, R. (1952). Foliicolous lichens. I. A revision of the taxonomy of the obligately foliicolous, lichenized fungi. Symbolae Botanicae Upsalienses, 12, 1–590.Google Scholar
Santesson, R., Moberg, R., Nordin, A., Tønsberg, T. and Vitikainen, O. (2004). Lichen-forming and Lichenicolous Fungi of Fennoscandia. Uppsala: Museum of Evolution, Uppsala University.Google Scholar
Sanz, M.-J., Gries, C. and Nash, T. H. III (1992). Dose-response relationships for SO2 fumigations in the lichens Evernia prunastri (L.) Ach. and Ramalina fraxinea (L.) Ach. New Phytologist, 122, 313–319.CrossRefGoogle Scholar
Schaper, T. and Ott, S. (2003). Photobiont selectivity and interspecific interactions in lichen communities. I. Culture experiments with the mycobiont Fulgensia bracteata. Plant Biology, 5, 441–450.CrossRefGoogle Scholar
Scheidegger, C. (1985). Systematische Studien zur Krustenflechte Anzina carneonivea (Trapeliacae, Lecanorales). Nova Hedwigia, 41, 191–218.Google Scholar
Scheidegger, C. (1993). A revision of saxicolous species of the genus Buellia De Not. and formerly included genera in Europe. Lichenologist, 25, 315–364.CrossRefGoogle Scholar
Scheidegger, C. (1994 a). Low-temperature scanning electron microscopy: the localization of free and perturbed water and its role in the morphology of the lichen symbionts. Cryptogamic Botany, 4, 290–299.Google Scholar
Scheidegger, C. (1994 b). Reproductive strategies in Vezdaea (Lecanorales, lichenized Ascomycetes): a low-temperature scanning electron microscopy study of a ruderal species. Cryptogamic Botany, 5, 163–171.Google Scholar
Scheidegger, C. (1998). Erioderma pedicellatum: a critically endangered lichen species. Species, 30, 68–69.Google Scholar
Scheidegger, C. and Schroeter, B. (1995). Effects of ozone fumigation on epiphytic macrolichens: ultrastructure, CO2 gas exchange and chlorophyll fluorescence. Environmental Pollution, 88, 345–354.CrossRefGoogle ScholarPubMed
Scheidegger, C., Schroeter, B. and Frey, B. (1995 a). Structural and functional processes during water vapour uptake and desiccation in selected lichens with green algal photobionts. Planta, 197, 399–409.CrossRefGoogle Scholar
Scheidegger, C., Wolseley, P. A. and Thor, G., eds., (1995 b). Conservation biology of lichenised fungi. Mitteilungen der Eidgenössischen Forschungsanstalt für Wald, Schnee und Landschaft, 70, 1–173.Google Scholar
Scherrer, S. and Honegger, R. (2003). Inter- and intraspecific variation of homologous hydrophobin (H1) gene sequences among Xanthoria spp. (lichen-forming ascomycetes). New Phytologist, 158, 375–389.CrossRefGoogle Scholar
Scherrer, S., Vries, O. M. H., Dudler, R., Wessels, J. G. H. and Honegger, R. (2000). Interfacial self-assembly of fungal hydrophobins of the lichen-forming ascomycetes Xanthoria parietina and X. ectaneoides. Fungal Genetics and Biology, 30, 81–93.CrossRefGoogle ScholarPubMed
Scherrer, S., Haisch, A. and Honegger, R. (2002). Characterization and expression of XPH1, the hydrophobin gene of the lichen-forming ascomycete Xanthoria parietina. New Phytologist, 154, 175–184.CrossRefGoogle Scholar
Scherrer, S., Zippler, U. and Honegger, R. (2005). Characterisation of the mating-type locus in the genus Xanthoria (lichen-forming ascomycetes, Lecanoromycetes). Fungal Genetics and Biology, 42, 976–988.CrossRefGoogle Scholar
Schlee, D., Kandzia, R., Tintemann, H. and Türk, R. (1995). Activity of superoxide dismutase and malondialdehyde content in lichens along an altitude profile. Phyton, 35, 233–242.Google Scholar
Schlensog, M., Pannewitz, S., Green, T. G. A. and Schroeter, B. (2004). Metabolic recovery of continental antarctic cryptogams after winter. Polar Biology, 27, 399–408.CrossRefGoogle Scholar
Schlensog, M., Schroeter, B. and Green, T. G. A. (2000). Water dependent photosynthetic activity of lichens from New Zealand: differences in the green algal and the cyanobacterial thallus parts of photosymbiodemes. Bibliotheca Lichenologica, 75, 149–160.Google Scholar
Schlensog, M., Schroeter, B., Sancho, L. G., Pintado, A. and Kappen, L. (1997). Effect of strong irradiance on photosynthetic performance of the melt-water dependent cyanobacterial lichen Leptogium puberulum (Collemataceae) Hue from the maritime Antarctic. Bibliotheca Lichenologica, 67, 235–246.Google Scholar
Schmitt, I. and Lumbsch, H. T. (2004). Molecular phylogeny of the Pertusariaceae supports secondary chemistry as an important systematic character set in lichen-forming ascomycetes. Molecular Phylogenetics and Evolution, 33, 43–55.CrossRefGoogle ScholarPubMed
Schmitt, I., Lumbsch, H. T. and Søchting, U. (2003). Phylogeny of the lichen genus Placopsis and its allies based on Bayesian analyses of nuclear and mitochondrial sequences. Mycologia, 95, 827–835.CrossRefGoogle ScholarPubMed
Schmitt, I., Martin, M. P., Kautz, S. and Lumbsch, T. H. (2005 a). Diversity of non-reducing polyketide synthase genes in the Pertusariales (lichenized Ascomycota). A phylogenetic perspective. Phytochemistry, 66, 1241–1253.CrossRefGoogle ScholarPubMed
Schmitt, I., Messuti, M. I., Feige, G. B. and Lumbsch, H. T. (2001). Molecular data support rejection of the generic concept in the Coccotremataceae (Ascomycota). Lichenologist, 33, 315–321.CrossRefGoogle Scholar
Schmitt, I., Mueller, G. M. and Lumbsch, H. T. (2005 b). Ascoma morphology is homoplaseous and phylogenetically misleading in some pyrenocarpous lichens. Mycologia, 97, 362–374.CrossRefGoogle ScholarPubMed
Schmitt, I., Yamamoto, Y. and Lumbsch, H T. (2006). Phylogeny of Pertusariales (Ascomycotina): resurrection of Ochrolechiaceae and a new circumscription of Megasporaceae. Journal of the Hattori Botanical Laboratory, 100, 753–764.Google Scholar
Schmull, M. and Hauck, M. (2003). Element microdistribution in the bark of Abies balsamea and Picea rubens and its impact on epiphytic lichen abundance on Whiteface Mountain, New York. Flora, 198, 293–303.CrossRefGoogle Scholar
Schreiber, U., Bilger, W. and Neubauer C. (1994). Chlorophyll fluorescence as a nonintrusive indicator for rapid assessment of in vivo photosynthesis. In Ecophysiology of Photosynthesis ed. Schulze, E.-D. and Caldwell, M. M., pp. 49–70. Ecological Studies 100. Berlin: Springer.CrossRefGoogle Scholar
Schroeter, B. and Scheidegger, C. (1995). Water relations in lichens at subzero temperatures: structural changes and carbon dioxide exchange in the lichen Umbilicaria aprina from continental Antarctica. New Phytologist, 131, 273–285.CrossRefGoogle Scholar
Schroeter, B., Green, T. G. A., Kappen, L. and Seppelt, D. (1994). Carbon dioxide exchange at subzero temperatures. Field measurements on Umbilicaria aprina in Antarctica. Cryptogamic Botany, 4, 233–241.Google Scholar
Schroeter, B., Kappen, L., Schulz, F. and Sancho, L. G. (2000). Seasonal variation in the carbon balance of lichens in the maritime Antarctic: long-term measurements of photosynthetic activity in Usnea aurantiaco-atra. In Antarctic Ecosystems: Model for Wider Ecological Understanding, ed. Davison, W., Howard-Williams, C. and Broady, P., pp. 220–224. Christchurch: Caxton Press.Google Scholar
Schroeter, B., Schulz, F. and Kappen, L. (1997). Hydration-related spatial and temporal variation of photosynthetic activity in Antarctic lichens. In Antarctic Communities: Species, Structure and Survival, ed. Battaglia, B., Valencia, J. and Walton, D. W. H., pp. 221–225. Cambridge: Cambridge University Press.Google Scholar
Schulz, M. (1995). Protein and ubiquitin conjugate patterns of Peltigera horizontalis (Huds.) Baumg. during desiccation and rehydration. In Flechten Follmann. Contributions to Lichenology in Honour of Gerhard Follmann, ed. Daniels, F. J. A., Schultz, M. and Peine, J., pp. 87–96. Cologne: Botanical Institute, University of Cologne.Google Scholar
Schulze, E.-D., Beck, E. and Müller-Hohenstein, K. (2002). Pflanzenökologie. Heidelberg: Spektrum Akademischer.Google Scholar
Schulze, E.-D. and Chapin, F. S., III (1987). Plant specialization to environments of different resource availability. In Potentials and Limitations of Ecosystem Analysis, ed. Schulze, E.-D. and Zwolfer, H., pp. 120–148. Berlin: Springer.CrossRefGoogle Scholar
Schüssler, A. (2002). Molecular phylogeny, taxonomy, and evolution of Geosiphon pyriformis and arbuscular mycorrhizal fungi. Plant and Soil, 244, 75–83.CrossRefGoogle Scholar
Schüssler, A., Schnepf, E., Mollenhauer, D. and Kluge, M. (1995). The fungal bladders of the endocyanosis Geosiphon pyriforme, a Glomus-related fungus: cell wall permeability indicates a limiting pore radius of only 0.5 nm. Protoplasma, 185, 131–139.CrossRefGoogle Scholar
Schüssler, A., Schwarzott, D. and Walker, C. (2001). A new fungal phylum, the Glomeromycota: phylogeny and evolution. Mycological Research, 105, 1413–1421.CrossRefGoogle Scholar
Schuster, G., Ott, S. and Jahns, H. M. (1985). Artificial cultures of lichens in the natural environment. Lichenologist, 17, 247–253.CrossRefGoogle Scholar
Schwartzman, D. W. and Volk, T. (1989). Biotic enhancement of weathering and the habitability of Earth. Nature, 340, 457–460.CrossRefGoogle Scholar
Schwendener, S. (1867). Über die wahre Natur der Flechtengonidien. Verhandlungen der schweizerischen naturforschenden Gesellschaft, 57, 9–11.Google Scholar
Schwendener, S. (1869). Die Algentypen der Flechtengonidien. Basel: Schultze.Google Scholar
Scotter, G. W. (1965). Chemical composition of forage lichens from northern Saskatchewan as related to use by barren-ground caribou in the taiga of northern Canada. Canadian Journal of Plant Sciences, 45, 246–250.CrossRefGoogle Scholar
Seaward, M. R. D. (1973). Lichen ecology of the Scunthorpe heathlands. I. Mineral accumulation. Lichenologist, 5, 423–433.CrossRefGoogle Scholar
Seaward, M. R. D. (1976). Performance of Lecanora muralis in an urban environment. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 323–357. London: Academic Press.Google Scholar
Seaward, M. R. D. (1982 a). Lichen ecology of changing urban environments. In Urban Ecology, ed. Bornkamm, R., Lee, J. A. and Seaward, M. R. D., pp. 181–189. Oxford: Blackwell Scientific.Google Scholar
Seaward, M. R. D. (1982 b). Principles and priorities of lichen conservation. Journal of the Hattori Botanical Laboratory, 52, 401–406.Google Scholar
Seaward, M. R. D. (1988). Contribution of lichens to ecosystems. In CRC Handbook of Lichenology, Vol. 2, ed. Galun, M., pp. 107–129. Boca Raton: CRC Press.Google Scholar
Seaward, M. R. D. (1993). Lichens and sulphur dioxide air pollution: field studies. Environmental Reviews, 1, 73–91.CrossRefGoogle Scholar
Seaward, M. R. D. (1996 a). Checklist of Tunisian lichens. Bocconea, 6, 115–148.Google Scholar
Seaward, M. R. D. (1996 b). Lichens and the environment. In A Century of Mycology, ed. Sutton, B. C., pp. 293–320. Cambridge: Cambridge University Press.Google Scholar
Seaward, M. R. D. (1997). Urban deserts bloom: a lichen renaissance. Bibliotheca Lichenologica, 67, 297–309.Google Scholar
Seaward, M. R. D. (ed.) (1998). Lichen Atlas of the British Isles. London: British Lichen Society.Google Scholar
Seaward, M. R. D. (2004). The use of lichens for environmental impact assessment. Symbiosis, 37, 293–305.Google Scholar
Seaward, M. R. D. and Aptroot, A. (2003). Lichens of Silhouette Island (Seychelles). Bibliotheca Lichenologica, 86, 423–439.Google Scholar
Seaward, M. R. D. and Coppins, B. J. (2004). Lichens and hypertrophication. Bibliotheca Lichenologica, 88, 561–572.Google Scholar
Seaward, M. R. D. and Edwards, H. G. M. (1997). Biological origin of major chemical disturbances on ecclesiastical architecture studied by Fourier Transform Raman spectroscopy. Journal of Raman Spectroscopy, 28, 691–696.3.0.CO;2-4>CrossRefGoogle Scholar
Seaward, M. R. D. and Letrouit-Galinou, M. (1991). Lichens return to the Jardin du Luxembourg after an absence of almost a century. Lichenologist, 23, 181–186.CrossRefGoogle Scholar
Sedelnikova, N. V. and Cheremisin, D. V. (2001). The use of lichens for dating of petroglyphs. Siberian Journal of Ecology, 8, 479–481.Google Scholar
Sensen, M. and Richardson, D. H. S. (2002). Mercury levels in lichens from different host trees around a chlor-alkali plant in New Brunswick, Canada. Science of the Total Environment, 293, 31–45.CrossRefGoogle ScholarPubMed
Sérusiaux, E. (1985). Goniocysts, goniocystangia and Opegrapha lambinonii and related species. Lichenologist, 17, 1–25.CrossRefGoogle Scholar
Sérusiaux, E. (1986). The nature and origin of campylidia in lichenized fungi. Lichenologist, 18, 1–35.CrossRefGoogle Scholar
Sérusiaux, E. (1989). Liste Rouge des Macrolichens dans la Communauté Européenne. Liège: Centre des Recherches sur les Lichens.Google Scholar
Seymour, F. A., Crittenden, P. D., Dickinson, M. J., et al. (2005 a). Breeding systems in the lichen-forming fungal genus Cladonia. Fungal Genetics and Biology, 42, 554–563.CrossRefGoogle ScholarPubMed
Seymour, F. A., Crittenden, P. D. and Dyer, P. S. (2005 b). Sex in the extremes: lichen-forming fungi. Mycologist, 19, 51–58.CrossRefGoogle Scholar
Sharma, P., Bergman, B., Hallbom, L. and Hofsten, A. (1982). Ultrastructural changes of Nostoc of Peltigera canina in presence of SO2. New Phytologist, 92, 573–579.CrossRefGoogle Scholar
Sheridan, R. P. (1979). Impact of emissions from coal-fired electricity generating facilities on N2-fixing lichens. Bryologist, 82, 54–58.CrossRefGoogle Scholar
Shibata, S. (1973). Some aspects of lichen chemotaxonomy. In Chemistry in Botanical Classification, Vol. 25, ed. Bendz, G. and Santesson, J., pp. 241–249. Stockholm: Nobel Symposia: Medicine and Natural Sciences.Google Scholar
Shibata, S. (1992). Studies on some lichen metabolites and their development. Journal of Japanese Botany, 67, 63–71.Google Scholar
Shibuya, M., Ebizuka, Y., Noguchi, H., Iitaka, Y. and Sankawa, U. (1983). Inhibition of prostaglandin biosynthesis by 4-O-methylcryptochlorophaeic acid: synthesis of monomeric arylcarboxylic acids for inhibitory activity testing and X-ray analysis of 4-O-methylcryptochlorophaeic acid. Chemical Pharmaceutical Bulletin of Tokyo, 31, 407–413.CrossRefGoogle ScholarPubMed
Shuvalov, V. A. and Heber, U. (2003). Photochemical reactions in dehydrated photosynthetic organisms, leaves, chloroplasts, and photosystem II particles: reversible reduction of pheophytin and chlorophyll and oxidation of β-carotene. Chemical Physics, 294, 227–237.CrossRefGoogle Scholar
Sigal, L. L. and Johnston, J. W. Jr. (1986). Effects of acidic rain and ozone on nitrogen fixation and photosynthesis in the lichen Lobaria pulmonaria (L.) Hoffm. Environmental and Experimental Botany, 26, 59–64.CrossRefGoogle Scholar
Sigal, L. L. and Nash, T. H. III (1983). Lichen communities on conifers in southern California: an ecological survey relative to oxidant air pollution. Ecology, 64, 1343–1354.CrossRefGoogle Scholar
Sigal, L. L. and Taylor, O. C. (1979). Preliminary studies of the gross photosynthetic response of lichens to peroxyacetylnitrite fumigations. Bryologist, 82, 564–575.CrossRefGoogle Scholar
Silberstein, L., Siegel, B. Z., Siegel, S. M., Mukhtar, A. and Galun, M. (1996 a). Comparative studies on Xanthoria parietina, a pollution-resistant lichen, and Ramalina duriaei, a sensitive species. I. Effects of air pollution on physiological processes. Lichenologist, 28, 355–365.CrossRefGoogle Scholar
Silberstein, L., Siegel, B. Z., Siegel, S. M., Mukhtar, A. and Galun, M. (1996 b). Comparative studies on Xanthoria parietina, a pollution-resistant lichen, and Ramalina duriaei, a sensitive species. II. Evaluation of possible air pollution-protection mechanisms. Lichenologist, 28, 367–383.CrossRefGoogle Scholar
Sillett, S. C. and Goslin, M. N. (1999). Distribution of epiphytic macrolichens in relation to remnant trees in a multiple-age Douglas-fir forest. Canadian Journal of Forest Research, 29, 1204–1215.CrossRefGoogle Scholar
Sillett, S. C., McCune, B., Peck, J. E., Rambo, T. R. and Ruchty, A. (2000). Dispersal limitations of epiphytic lichens result in species dependent on old-growth forests. Ecological Applications, 10, 789–799.CrossRefGoogle Scholar
Simpson, T. J. (1995). Polyketide biosynthesis. Chemistry and Industries, 1995, 407–411.Google Scholar
Sinnemann, S. J., Andrésson, Ó. S., Brown, D. W. and Miao, V. P. (2000). Cloning and heterologous expression of Solorina crocea pyrG. Current Genetics, 37, 333–338.CrossRefGoogle ScholarPubMed
Sipman, H. J. M. (1994). Foliicolous lichens on plastic tape. Lichenologist, 26, 311–312.CrossRefGoogle Scholar
Sipman, H. J. M. (1997). Observations on the foliicolous lichen and bryophyte flora in the canopy of a semi-deciduous tropical forest. Abstracta Botanica, 21, 153–161.Google Scholar
Sipman, H. J. M. (2002). The significance of the northern Andes for lichens. Botanical Review, 68, 88–99.CrossRefGoogle Scholar
Sipman, H. J. M. (2006 a). Diversity and biogeography of lichens in neotropical montane oak forests. In Ecology and Conservation of Neotropical Oak Forests, ed. Kappelle, M., pp. 69–81. Ecological Studies 185. Berlin: Springer.CrossRefGoogle Scholar
Sipman, H. J. M. (2006b). Identification key and literature guide to the genera of lichenized fungi (Lichens) in the Neotropics (provisional version). Online: www.bgbm.org/sipman/keys/neokeyA.htm.
Sipman, H. J. M. and Harris, R. C. (1989). Lichens. In Tropical Rain Forest Ecosystems, ed. Lieth, H. and Werger, M. J. A., pp. 303–309. Amsterdam: Elsevier.Google Scholar
Skujins, J. and Klubek, B. (1978). Nitrogen fixation and cycling by blue-green algae-lichen-crusts in arid rangeland soils. Ecological Bulletin (Stockholm), 26, 164–171.Google Scholar
Skulachev, V. P. (1998). Uncoupling: new approaches to an old problem of bioenergetics. Biochimica Biophysica Acta, 1363, 100–124.CrossRefGoogle Scholar
Skult, H. (1984). The Parmelia omphaloides (Ascomycetes) complex in Eastern Fennoscandia. Annales Botanici Fennici, 21, 117–142.Google Scholar
Skye, E. (1968). Lichens and air pollution: a study of cryptogamic epiphytes and environment in the Stockholm region. Acta Phytogeographica Suecica, 52, 8–123.Google Scholar
Slocum, R. D., Ahmadijan, V. and Hildreth, K. C. (1980). Zoosporogenesis in Trebouxia gelatinosa: ultrastrucutral potential for zoospore release and implications for the lichen association. Lichenologist, 12, 173–187.CrossRefGoogle Scholar
Sluiman, H. J. (1989). The green algal class Ulvophyceae – an ultrastructural survey and classification. Cryptogamic Botany, 1, 83–94.Google Scholar
Sluiman, H. J., Kouwets, F. A. C. and Blommers, P. C. J. (1989). Classification and definition of cytokinetic patterns in Green Algae: sporulation versus (vegetative) cell division. Archiv für Protistenkunde, 137, 277–90.CrossRefGoogle Scholar
Smith, D. C. and Douglas, A. (1987). The Biology of Symbiosis. London: Edward Arnold.Google Scholar
Smith, D. C. and Molesworth, S. (1973). Lichen physiology. XIII. Effects of rewetting of dry lichens. New Phytologist, 72, 525–533.CrossRefGoogle Scholar
Smith, E. C. and Griffiths, H. (1996). The occurrence of the chloroplast pyrenoid is correlated with the activity of a CO2 concentrating mechanism and carbon isotope discrimination in lichens and bryophytes. Planta, 198, 6–16.CrossRefGoogle Scholar
Snelgar, W. P. (1981). The ecophysiology of New Zealand forest lichens with special reference to carbon dioxide exchange. Ph.D. thesis, Waikato University, Hamilton, New Zealand.
Snelgar, W. P. and Green, T. G. A. (1980). Carbon dioxide exchange in lichens. II. Low carbon dioxide compensation levels and lack of apparent photorespiratory activity in some lichens. Bryologist, 83, 505–507.CrossRefGoogle Scholar
Snelgar, W. P. and Green, T. G. A. (1981). Carbon dioxide exchange in lichens: apparent photorespiration and the possible role of CO2 refixation in some members of the Stictaceae (Lichenes). Journal of Experimental Botany, 32, 661–668.CrossRefGoogle Scholar
Snelgar, W. P. and Green, T. G. A. (1982). Growth rates of Stictaceae lichens in New Zealand beech forests. Bryologist, 85, 301–306.CrossRefGoogle Scholar
Snelgar, W. P., Green, T. G. A. and Beltz, C. K. (1981 a). Carbon dioxide exchange in lichens: estimation of internal thallus CO2 transport resistances. Physiologia Plantarum, 52, 417–422.CrossRefGoogle Scholar
Snelgar, W. P., Green, T. G. A. and Wilkins, A. L. (1981 b). Carbon dioxide exchange in lichens. I. Resistances to CO2 uptake at different thallus water contents. New Phytologist, 88, 353–361.CrossRefGoogle Scholar
Søchting, U. (1995). Lichens as monitors of nitrogen deposition. Cryptogamic Botany, 5, 264–269.Google Scholar
Søchting, U. and Johnsen, I. (1978). Lichen transplants as biological indicators of SO2 air pollution in Copenhagen. Bulletin of Environmental Contamination and Toxicology, 19, 1–7.CrossRefGoogle ScholarPubMed
Solhaug, K. A. and Gauslaa, Y. (1996). Parietin, a photoprotective secondary product of the lichen Xanthoria parietina. Oecologia, 108, 412–418.CrossRefGoogle ScholarPubMed
Solhaug, K. A. and Gauslaa, Y. (2004 a). Testing of ecological roles of secondary lichen compounds. In Lichens in Focus, ed. Randlane, T. and Saag, A., p. 40. Tartu: Tartu University Press.Google Scholar
Solhaug, K. A. and Gauslaa, Y. (2004 b). Photosynthates stimulate the UV-B induced fungal anthraquinone synthesis in the foliose lichen Xanthoria parietina. Plant Cell and Environment, 27, 167–176.CrossRefGoogle Scholar
Solhaug, K. A., Gauslaa, Y., Nybakken, L. and Bilger, W. (2003). UV-induction of sun-screening pigments in lichens. New Phytologist, 158, 91–100.CrossRefGoogle Scholar
Solomina, O. and Calkin, P. E. (2003). Lichenometry as applied to moraines in Alaska, U.S.A., and Kamchatka, Russia. Arctic, Antarctic, and Alpine Research, 35, 129–143.CrossRefGoogle Scholar
Sommerkorn, M. (2000). The ability of lichens to benefit from natural CO2 enrichment under a spring snow-cover: a study with two arctic-alpine species from contrasting habitats. Bibliotheca Lichenologica, 75, 365–380.Google Scholar
Sonesson, M., Osborne, C. and Sandberg, G. (1994). Epiphytic lichens as indicators of snow depth. Arctic and Alpine Research, 26, 159–165.CrossRefGoogle Scholar
Sonesson, M., Schipperges, B. and Carlsson, B. Å. (1992). Seasonal patterns of photosynthesis in alpine and subalpine populations of the lichen Nephroma arcticum. Oikos, 65, 3–12.CrossRefGoogle Scholar
Spiro, B., Morrisson, J. and Purvis, O. W. (2002). Sulphur isotopes in lichens as indicators of sources. In Monitoring with Lichens – Monitoring Lichens. Nato Science Series IV: Earth and Environmental Sciences, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 311–315. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Clair, St. L. L. and Seaward, M. R. D., eds. (2004). Biodeterioration of Stone Surfaces. Dordrecht: Kluwer.CrossRefGoogle Scholar
Clair, St., Johansen, L. L., , J. R. and Rushforth, S. R. (1993). Lichens of soil crust communities in the intermountain area of the western United States. Great Basin Naturalist, 53, 5–12.Google Scholar
Staiger, B. (2002). Die Flechtenfamilie Graphidaceae. Studien in Richtung einer natürlicheren Gliederung. Bibliotheca Lichenologica, 85, 1–526.Google Scholar
Staiger, B., Kalb, K. and Grube, M. (2006). Phylogeny and phenotypic variation in the lichen family Graphidaceae (Ostropomycetidae, Ascomycota). Mycological Research, 110, 765–772.CrossRefGoogle Scholar
Stålfelt, M. G. (1939). Der Gasaustausch der Flechten. Planta, 29, 11–31.CrossRefGoogle Scholar
Stamatakis, A. (2006). RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics, 22, 2688–2690.CrossRefGoogle ScholarPubMed
Stamp, N. (2004). Can the growth-differentiation balance hypothesis be tested rigorously?Oikos, 107, 439–448.CrossRefGoogle Scholar
Steinkötter, J., Bhattacharya, D., Semmelroth, I., Bibeau, C. and Melkonian, M. (1994). Prasinophytes form independent lineages within the Chlorophyta: evidence from ribosomal RNA sequence comparisons. Journal of Phycology, 30, 340–345.CrossRefGoogle Scholar
Steinnes, E. and Krog, H. (1977). Mercury, arsenic and selenium fall-out from an industrial complex studied by means of lichen transplants. Oikos, 28, 160–164.CrossRefGoogle Scholar
Stenroos, S. (1993). Taxonomy and distribution of the lichen family Cladoniaceae in the Antarctic and peri-Antarctic regions. Cryptogamic Botany, 3, 310–344.Google Scholar
Stenroos, S. and DePriest, P. T. (1998). SSU rDNA phylogeny of cladoniiform lichens. American Journal of Botany, 85, 1548–1559.CrossRefGoogle ScholarPubMed
Stenroos, S., Feuerer, T. and Ahti, T. (2002 a). Chilean Cladoniaceae online. Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg, 30–32, 241–251.Google Scholar
Stenroos, S., Hyvönen, J., Myllys, L, Thell, A. and Ahti, T. (2002 b). Phylogeny of the genus Cladonia s. lat. (Cladoniaceae, Ascomycetes) inferred from molecular, morphological, and chemical data. Cladistics, 18, 237–278.CrossRefGoogle Scholar
Stenroos, S., Stocker-Wörgötter, E., Yoshimura, I., et al. (2003). Culture experiments and DNA sequence data confirm the identity of Lobaria photomorphs. Canadian Journal of Botany, 81, 232–247.CrossRefGoogle Scholar
Stewart, W. D. P. (1980). Some aspects of structure and function in N2-fixing cyanobacteria. Annual Reviews in Microbiology, 34, 497–536.CrossRefGoogle ScholarPubMed
Stewart, W. D. P. and Rodgers, G. A. (1978). Studies on the symbiotic blue-green algae of Anthoceros, Blasia, and Peltigera. In Environmental Role of Nitrogen-fixing Blue-green Algae and Asymbiotic Bacteria, ed. Granhall, U., pp. 247–259. Stockholm: Swedish Natural Science Research Council.Google Scholar
Stewart, W. D. P. and Rowell, P. (1975). Effects of L-methionine DL-sulphoximine on the assimilation of newly fixed NH3, C2H2 reduction and heterocyst production in Anabaena cylindrica. Biochemical Biophysical Research Communications, 65, 846–856.CrossRefGoogle Scholar
Stewart, W. D. P. and Rowell, P. (1977). Modifications of nitrogen-fixing algae in lichen symbioses. Nature (London), 265, 371–372.CrossRefGoogle Scholar
Stewart, W. D. P., Fitzgerald, G. P. and Burris, R. H. (1967). In situ studies on N2 fixation using the acetylene reduction technique. Proceedings of the National Academy of Sciences, USA, 58, 2071–2088.CrossRefGoogle ScholarPubMed
Stocker, O. (1927). Physiologische und ökologische Untersuchungen an Laub- und Strauchflechten. Flora, 21, 334–415.Google Scholar
Stocker-Wörgötter, E. (1995). Experimental cultivation of lichens and lichen symbionts. Canadian Journal of Botany, 73, S579–S589.CrossRefGoogle Scholar
Stocker-Wörgötter, E. (2001). Experimental lichenology and microbiology of lichens: culture experiments, secondary chemistry of cultured mycobionts, resynthesis and thallus morphogenesis. Bryologist, 104, 576–581.CrossRefGoogle Scholar
Stocker-Wörgötter, E. (2002 a). Resynthesis of Photosymbiodemes. In Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 47–60. Berlin: Springer.Google Scholar
Stocker-Wörgötter, E. (2002 b). Analysis of secondary compounds in cultured mycobionts. In Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 296–306. Berlin: Springer.Google Scholar
Stocker-Wörgötter, E. (2005). Approaches to a biotechnology of lichen-forming fungi: induction of polyketide pathways and chemosyndromes in axenically cultured mycobionts. Recent Research Developments in Phytochemistry, 9, 115–131.Google Scholar
Stocker-Wörgötter, E. (2008). Metabolic diversity of lichen-forming ascomycetous fungi: culturing, polyketide and shikimate metabolite production, and PKS genes. Natural Product Reports, 25, 188–200.CrossRefGoogle ScholarPubMed
Stocker-Wörgötter, E. and Elix, J. A. (2004). Experimental studies of lichenized fungi: formation of rare depsides and dibenzofurans by the cultured mycobiont of Bunodophoron patagonicum (Sphaerophoraceae, lichenized Ascomycota). In Contributions to Lichenology, Festschrift in Honour of Hannes Hertel, ed. Döbbeler, P. and Rambold, G., pp. 659–669. Bibliotheca Lichenologica 88. Berlin: J. Cramer.Google Scholar
Stocker-Wörgötter, E. and Elix, J. A. (2006). Morphogenetic strategies and induction of secondary metabolite biosynthesis in cultured lichen-forming Ascomycota, as exemplified by Cladia retipora (Labill.) Nyl. and Dactylina arctica (Richards.) Nyl. Symbiosis, 41, 9–20.Google Scholar
Stocker-Wörgötter, E., Elix, J. A. and Grube, M. (2004). Secondary chemistry of lichen-forming fungi: chemosyndromic variation and DNA-analyses of cultures and chemotypes in the Ramalina farinacea complex. Bryologist, 107, 152–162.CrossRefGoogle Scholar
Stone, D. F. (1989). Epiphytic succession on Quercus garryana branches in the Willamette Valley of western Oregon. Bryologist, 92, 81–94.CrossRefGoogle Scholar
Stulen, I. and De Kok, L. J. (1993). Whole plant regulation of sulfur-metabolism – a theoretical approach and comparison with current ideas on regulation of nitrogen metabolism. In Sulfur Nutrition and Assimilation in Higher Plants, ed. Kok, L. J., Stulen, I., Rennenberg, H., Brunold, C. and Rauser, W. E., pp. 77–91. The Hague: SPB Academic Publishing.Google Scholar
Sugiyama, K., Kurokawa, S. and Okada, G. (1976). Studies on lichens as a bioindicator of air pollution. I. Correlation of distribution of Parmelia tinctorum with SO2 air pollution. Japanese Journal of Ecology, 26, 209–212.Google Scholar
Sun, H. J. and Friedmann, E. I. (2005). Communities adjust their temperature optima by shifting producer-to-consumer ratio, shown in lichens as models. I. Experimental verification. Microbial Ecology, 49, 528–535.CrossRefGoogle ScholarPubMed
Sun, H. J., DePriest, P. T., Gargas, A., Rossman, A. Y. and Friedmann, E. I. (2002). Pestalotiopsis maculans: a dominant parasymbiont in North American lichens. Symbiosis, 33, 215–226.Google Scholar
Sundberg, B., Ekblad, A, Näsholm, T. and Palmqvist, K. (1999 a). Lichen respiration in relation to active time, nitrogen and ergosterol concentrations. Functional Ecology, 13, 119–125.CrossRefGoogle Scholar
Sundberg, B., Lundberg, P., Ekblad, A. and Palmqvist, K. (1999 b). In vivo13C NMR spectroscopy of carbon fluxes in four diverse lichens. In Physiological Ecology of Lichen Growth, by Sundberg, B.. Ph.D. thesis paper VI. Umeå: Umeå University.Google Scholar
Sundberg, B., Näsholm, T. and Palmqvist, K. (2001). The effect of nitrogen on growth and key thallus components in the two tripartite lichens, Nephroma arcticum and Peltigera aphthosa. Plant, Cell and Environment, 24, 517–527.CrossRefGoogle Scholar
Sundberg, B., Palmqvist, K., Esseen, P.-A. and Renhorn, K.-E. (1997). Growth and vitality of epiphytic lichens. II. Modelling of carbon gain using field and laboratory data. Oecologia, 109, 10–18.CrossRefGoogle Scholar
Suryanarayanan, T. S., Thirunavukkarasu, N., Hariharan, G. N. and Balaji, P. (2005). Occurrence of non-obligate microfungi inside lichen thalli. Sydowia, 57, 120–130.Google Scholar
Svenning, M. M., Eriksson, T. and Rasmussen, U. (2005). Phylogeny of symbiotic cyanobacteria within the genus Nostoc based on 16S rDNA sequence analyses. Archiv für Mikrobiologie, 183, 19–26.CrossRefGoogle ScholarPubMed
Swanson, A. and Fahselt, D. (1997). Effects of ultraviolet light on the polyphenolics of Umbilicaria americana Poelt & Nash. Canadian Journal of Botany, 75, 284–289.CrossRefGoogle Scholar
Syers, J. K. and Iskander, I. K. (1973). Pedogenetic significance of lichens. In The Lichens, ed. Ahmadjian, V. and Hale, M. E., pp. 225–248. New York: Academic Press.Google Scholar
Szabo, I., Bergantino, E. and Giacometti, G. M. (2005). Light and oxygenic photosynthesis: energy dissipation as a protection mechanism against photo-oxidation. EMBO Reports, 6, 629–634.CrossRefGoogle ScholarPubMed
Takala, K., Kaurenen, P. and Olkkonen, H. (1978). Fluorine content of two lichen species in the vicinity of a fertilizer factory. Annales Botanici Fennici, 15, 158–167.Google Scholar
Takala, K., Olkkonen, H., Ikonen, J., Jääskeläinen, J. and Puumalainen, P. (1985). Total sulphur contents of epiphytic and terricolous lichens in Finland. Annales Botanici Fennici, 22, 91–100.Google Scholar
Tanabe, Y., Watanabe, M. M. and Sugiyama, J. (2002). Are Microsporidia really related to Fungi?: a reappraisal based on additional gene sequences from basal fungi. Mycological Research, 106, 1380–1391.CrossRefGoogle Scholar
Tapper, R. (1976). Dispersal and changes in the local distribution of Evernia prunastri and Ramalina farinacea. New Phytologist, 77, 725–734.CrossRefGoogle Scholar
Tarhanen, S., Holopainen, T. and Oksanen, J. (1997). Ultrastructural changes and electrolyte leakage from ozone fumigated epiphytic lichens. Annals of Botany, 80, 611–621.CrossRefGoogle Scholar
Taylor, O. C. (1969). Importance of peroxyacetyl nitrate (PAN) as a phytotoxic air pollutant. Journal of the Air Pollution Control Association, 19, 347–351.CrossRefGoogle Scholar
Taylor, R. J. and Bell, M. (1983). Effects of SO2 on the lichen flora in an industrial area, northwest Whatcom County, Washington. Northwest Science, 57, 157–166.Google Scholar
Taylor, T. N., Hass, H. and Kerp, H. (1997). A cyanolichen from the Lower Devonian Rhynie Chert. American Journal of Botany, 84, 992–1004.CrossRefGoogle ScholarPubMed
Taylor, T. N., Hass, H., Remy, W. and Kerp, H. (1995 b). The oldest fossil lichen. Nature, 378, 244.CrossRefGoogle Scholar
Taylor, T. N., Remy, W., Hass, H. and Kerp, H. (1995 a). Fossil arbuscular mycorrhizae from the early Devonian. Mycologia, 87, 560–573.CrossRefGoogle Scholar
Taylor, W. A., Free, C., Boyce, C., Helgemo, R. and Ochoada, J. (2004). SEM analysis of Spongiophyton interpreted as a fossil lichen. International Journal of Plant Sciences, 165, 875–881.CrossRefGoogle Scholar
Tehler, A. (1982). The species pair concept in lichenology. Taxon, 31, 708–714.CrossRefGoogle Scholar
Tehler, A. (1988). A cladistic outline of the Eumycota. Cladistics, 4, 227–277.CrossRefGoogle Scholar
Tehler, A. (1990). A new approach to the phylogeny of Euascomycetes with a cladistic outline of Arthoniales focussing on Roccellaceae. Canadian Journal of Botany, 68, 2458–2592.CrossRefGoogle Scholar
Tehler, A. (1995). Morphological data, molecular data and total evidence in phylogenetic analysis. Canadian Journal of Botany, 73, S667–S676.CrossRefGoogle Scholar
Tehler, A. (1996). Systematics, phylogeny and classification. In Lichen Biology, ed. Nash, T. H. III, pp. 217–239. Cambridge: Cambridge University Press.Google Scholar
Tehler, A. and Irestedt, M. (2007). Parallel evolution of lichen growth forms in the family Roccellaceae (Arthoniales, Ascomycota). Cladistics, 23, 432–454.CrossRefGoogle Scholar
Tehler, A., Dahlkild, Å., Eldenäs, P. and Feige, G. B. (2004). The phylogeny and taxonomy of Macaronesian, European and Mediterranean Roccella (Roccellaceae, Arthoniales). Symbolae Botanicae Upsalienses, 34, 405–428.Google Scholar
Tehler, A., Farris, J. S., Lipscomb, D. L. and Källersjö, M. (2000). Phylogenetic analyses of the fungi based on large rDNA data sets. Mycologia, 92, 459–474.CrossRefGoogle Scholar
Tehler, A., Little, D. P. and Farris, J. S. (2003). The full-length phylogenetic tree from 1551 ribosomal sequences of chitinous fungi, Fungi. Mycological Research, 107, 901–916.CrossRefGoogle ScholarPubMed
Tel-Or, E. and Stewart, W. D. P. (1976). Photosynthetic electron transport, ATP synthesis and nitrogenase activity in isolated heterocysts of Anabaena cylindrica. Biochimica et Biophysica Acta, 423, 189–195.CrossRefGoogle ScholarPubMed
Tel-Or, E. and Stewart, W. D. P. (1977). Photosynthetic components of activities of nitrogen-fixing isolated heterocysts of Anabaena cylindrica. Proceedings of the Royal Society of London B, 198, 61–96.CrossRefGoogle Scholar
Théau, J. and Duguay, C. R. (2003). Mapping lichen changes in the summer range of the George River caribou herd (Québec-Labrador, Canada) using Landsat imagery (1976–1998). Rangifer, 24, 31–49.CrossRefGoogle Scholar
Théau, J. and Duguay, C. R. (2004). Lichen mapping in the summer range of the George River caribou herd using Landsat TM imagery. Canadian Journal of Remote Sensing, 30, 867–881.CrossRefGoogle Scholar
Théau, J., Peddle, D. R. and Duguay, C. R. (2005). Mapping lichen in a caribou habitat of northern Quebec, Canada, using an enhancement-classification method and spectral mixture analysis. Remote Sensing of Environment, 94, 232–243.CrossRefGoogle Scholar
Thell, A. and Goward, T. (1996). The new cetrarioid genus Kaernefeltia and related groups in the Parmeliaceae (lichenized Ascomycotina). Bryologist, 99, 125–136.CrossRefGoogle Scholar
Thell, A., Feuerer, T.Kärnefelt, I., Myllys, L. and Stenroos, S. (2004 a). Monophyletic groups within the Parmeliaceae identified by IST rDNA, β-tubulin and GAPDH sequences. Mycological Progress, 3, 297–314.CrossRefGoogle Scholar
Thell, A., Goward, T., Randlane, T., Kärnefewlt, E. I. and Saag, A. (1995). A revision of the North American lichen genus Ahtiana (Parmeliaceae). Bryologist, 98, 596–605.CrossRefGoogle Scholar
Thell, A., Randlane, T. and Saag, A. (2005). A new circumscription of the lichen genus Nephromopsis (Parmeliaceae, lichenized Ascomycetes). Mycological Progress, 4, 303–316.CrossRefGoogle Scholar
Thell, A., Westberg, M. and Kärnefelt, I. (2004 b). Biogeography of the lichen family Parmeliaceae in the Nordic countries with taxonomic remarks. Symbolae Botanicae Upsalienses, 34, 429–452.Google Scholar
Thomas, J., Wolk, C. P., Shaffer, P. W., Austin, S. M. and Galonsky, A. (1975). The initial organic product of fixation of 15N-labeled nitrogen gas by the blue-green alga Anabaena cylindrica. Biochemical Biophysical Research Communication, 67, 501–507.CrossRefGoogle Scholar
Thomas, M. A. (1999). Effect of sulfur dioxide and ozone on glutathione reductase and superoxide dismutase in lichens. Ph.D. dissertation. Tempe: Arizona State University.
Thomas, M. A., Romagni, J. G., Gries, C. and Nash, T. H. III (1997). The effects of sulfur dioxide exposure on glutathione reductase activity in the cyanolichen Peltigera canina. Supplement to Plant Physiology, 114, 57.Google Scholar
Thomson, J. W. (1972). Distributional patterns in American Arctic lichens. Canadian Journal of Botany, 50, 1135–1156.CrossRefGoogle Scholar
Tibell, L. (1984). A reappraisal of the taxonomy of Caliciales. Nova Hedwigia, Beiheft, 79, 597–713.Google Scholar
Tibell, L. (1998). Practice and prejudice in lichen classification. Lichenologist, 30, 439–453.CrossRefGoogle Scholar
Tibell, L. (1999). Calicioid lichens and fungi. Nordic Lichen Flora, 1, 20–71.Google Scholar
Tibell, L. (2001). Photobiont association and molecular phylogeny of the lichen genus Chaenotheca. Bryologist, 104, 191–198.CrossRefGoogle Scholar
Tibell, L. (2003). Tholurna dissimilis and generic delimitations in Caliciaceae inferred from nuclear IST and LSU rDNA phylogenies (Lecanorales, lichenized Ascomycetes). Mycological Research, 107, 1403–1418.CrossRefGoogle Scholar
Tibell, L. and Beck, A. (2001). Morphological variation, photobiont association and ITS phylogeny of Chaenotheca phaeocephala and C. subroscida (Coniocybaceae, lichenized Ascomycetes). Nordic Journal of Botany, 22, 651–660.CrossRefGoogle Scholar
Timdal, E. (1991). A monograph of the genus Toninia (Lecideaceae, Ascomycetes). Opera Botanica, 110, 1–137.Google Scholar
Timdal, E. and Tønsberg, T. (2006). Psoroma paleaceum comb. nov. the only hairy Psoroma in northern Europe. Graphis Scripta, 18, 54–57.Google Scholar
Timoney, K. P. and Marsh, J. E. (2004). Lichen trimlines in northern Alberta: establishment, growth rates, and historic water levels. Bryologist, 107, 429–440.CrossRefGoogle Scholar
Tomassini, F. D., Lavoie, P., Puckett, K. J., Nieboer, E. and Richardson, D. H. S. (1977). The effect of time of exposure to sulphur dioxide on potassium loss from and photosynthesis in the lichen, Cladina rangiferina (L.) Harm. New Phytologist, 79, 147–155.CrossRefGoogle Scholar
Tomassini, F. D., Puckett, K. J., Nieboer, E., Richardson, D. H. S. and Grace, B. (1976). Determination of copper, iron, nickel, and sulphur by x-ray fluorescence in lichens from the Mackenzie Valley, Northwest Territories, and the Sudbury District, Ontario. Canadian Journal of Botany, 54, 1591–1603.CrossRefGoogle Scholar
Tomitani, A., Knoll, A. H., Cavanaugh, C. M. and Ohna, T. (2006). The evolutionary diversification of cyanobacteria: molecular-phylogenetic and paleontological perspectives. Proceedings of the National Academy of Sciences, USA, 103, 5442–5447.CrossRefGoogle ScholarPubMed
Tønsberg, T. (1992). The sorediate and isidiate, corticolous, crustose lichens in Norway. Sommerfeltia, 14, 1–331.Google Scholar
Tønsberg, T. and Holtan-Hartwig, J. (1983). Phycotype pairs in Nephroma, Peltigera and Lobaria in Norway. Nordic Journal of Botany, 3, 681–688.CrossRefGoogle Scholar
Topham, P. B. (1977). Colonization, growth, succession and competition. In Lichen Ecology, ed. Seaward, M. R. D., pp. 31–68. London: Academic Press.Google Scholar
Tormo, R., Recio, D., Silva, I. and Muñoz, A. (2001). A quantitative investigation of airborne algae and lichen soredia obtained from pollen traps in south-west Spain. European Journal of Phycology, 36, 385–390.CrossRefGoogle Scholar
Trembley, M. L., Ringli, C. and Honegger, R. (2002 a). Differential expression of hydrophobins DGH1, DGH2 and DGH3 and immunolocalization of DGH1 in strata of the lichenized basidiocarp of Dictyonema glabratum. New Phytologist, 154, 185–195.CrossRefGoogle Scholar
Trembley, M. L., Ringli, C. and Honegger, R. (2002 b). Hydrophobins DGH1, DGH2, and DGH3 in the lichen-forming basidiomycete Dictyonema glabratum. Fungal Genetics and Biology, 35, 247–259.CrossRefGoogle ScholarPubMed
Trembley, M. L., Ringli, C. and Honegger, R. (2002 c). Morphological and molecular analysis of early stages in the resynthesis of the lichen Baeomyces rufus. Mycological Research, 106, 768–776.CrossRefGoogle Scholar
Tretiach, M. and Carpanelli, A. (1992). Chlorophyll content and morphology as factors influencing the photosynthetic rate of Parmelia caperata. Lichenologist, 24, 81–90.Google Scholar
Triebel, D. and Rambold, G. (1988). Cecidonia und Phacopsis (Lecanorales): zwei lichenicole Pilzgattungen mit cecidogenen Arten. Nova Hedwigia, 47, 279–309.Google Scholar
Tschermak, E. (1941). Untersuchungen über die Beziehungen von Pilz und Alge im Flechtenthallus. Österreichische Botanische Zeitschrift, 90, 233–307.CrossRefGoogle Scholar
Tschermak-Woess, E. (1976). Algal taxonomy and the taxonomy of lichens: the phycobiont of Verrucaria adriatica. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 79–87. Orlando: Academic Press.Google Scholar
Tschermak-Woess, E. (1978). Myrmecia reticulata as a phycobiont and free-living Trebouxia – the problem of Stenocybe septata. Plant Systematics and Evolution, 129, 185–208.CrossRefGoogle Scholar
Tschermak-Woess, E. (1980). Chaenothecopsis consociata – kein parasitischer oder parasymbiotischer Pilz, sondern lichenisiert mit Dictyochloropsis symbiontica, spec. nova. Plant Systematics and Evolution, 136, 287–306.CrossRefGoogle Scholar
Tschermak-Woess, E. (1988). The algal partner. In CRC Handbook of Lichenology, Vol. 1, ed. Galun, M., pp. 39–92. Boca Raton: CRC Press.Google Scholar
Tschermak-Woess, E. and Poelt, J. (1976). Vezdea, a peculiar lichen genus, and its phycobiont. In Lichenology: Progress and Problems, ed. Brown, D. H., Hawksworth, D. L. and Bailey, R. H., pp. 89–105. Orlando: Academic Press.Google Scholar
Tuba, Z., Csintalan, Z., Szente, K., Nagy, Z. and Grace, J. (1998). Carbon gains by desiccation-tolerant plants at elevated CO2. Functional Ecology, 12, 39–44.CrossRefGoogle Scholar
Tuba, Z., Proctor, M. C. F. and Takács, Z. (1999). Desiccation-tolerant plants under elevated air CO2: a review. Zeitschrift für Naturforschung, Section C, 54, 788–796.Google Scholar
Tupa, D. D. (1974). An investigation of certain chaetophoralean algae. Nova Hedwigia, 46, 1–155.Google Scholar
Turgeon, B. and Yoder, O. (2000). Proposed nomenclature for mating type genes in filamentous ascomycetes. Fungal Genetics and Biology, 31, 1–5.CrossRefGoogle ScholarPubMed
Türk, R. and Christ, R. (1980). Untersuchungen des CO2-Gaswechsels von Flechtenexplantaten zur Indikation von SO2-Belastung im Stadtgebiet von Salzburg. In Bioindikation auf subzellularen und zellular Ebene (Bioindikation 2), ed. Schubert, R. and Schuh, J., pp. 39–45. Halle-Wittenberg: Martin-Luther-Universitat.Google Scholar
Türk, R. and Wirth, V. (1975). The pH dependence of SO2 damage to lichens. Oecologia, 19, 285–291.CrossRefGoogle Scholar
Türk, R., Wirth, V. and Lange, O. L. (1974). CO2-Gaswechsel-Untersuchungen zur SO2-Resistenz von Flechten. Oecologia, 15, 33–64.Google Scholar
Turner, S., Huang, T. C., and Chaw, S.-M. (2001). Molecular phylogeny of nitrogen-fixing unicellular cyanobacteria. Botanical Bulletin of Academia Sinica, 42, 181–186.Google Scholar
Turner, W. B. and Aldridge, D. C. (1983). Fungal Metabolites II. London: Academic Press.Google Scholar
US EPA (1996). Air Quality Criteria for Ozone and Related Photochemical Oxidants. Vol. II. EPA/600/p-93/004bF. Research Triangle Park, N.C. National Center for Environmental Assessment, Office of Research and Development.
Vainio, E. A. (1890). Étude sur classification naturelle et morphologie des lichens du Brésil. Acta Societatis pro Fauna et Flora Fennica, Helsinki, 7, 1–256.Google Scholar
Valladares, F., Sancho, L. G. and Ascaso, C. (1997). Water storage in the lichen family Umbilicariaceae. Botanica Acta, 111, 1–9.Google Scholar
Hoek, C., Jahns, H. M. and Mann, D. G. (1993). Algen, 3. Auflage. Stuttgart: Thieme.Google Scholar
Eerden, L., Vries, W. and Dobben, H. (1998). Effects of ammonia deposition on forests in The Netherlands. Atmospheric Environment, 32, 525–532.CrossRefGoogle Scholar
van Dobben, H. (1993) Vegetation as a monitor for deposition of nitrogen and acidity. Ph.D. dissertation, University of Utrecht.
Dobben, H. F. (1996). Decline and recovery of epiphytic lichens in an agricultural area in The Netherlands (1900–1988). Nova Hedwigia, 62, 477–485.Google Scholar
Dobben, H. F. and Bakker, A. J. (1996). Re-mapping epiphytic lichen biodiversity in The Netherlands: effects of decreasing SO2 and increasing NH3. Acta Botanica Neerlandica, 45, 55–71.CrossRefGoogle Scholar
Dobben, H. F. and Braak, C. J. F. (1998). Effects of atmospheric NH3 on epiphytic lichens in The Netherlands: the pitfalls of biological monitoring. Atmospheric Environment, 32, 551–557.CrossRefGoogle Scholar
Dobben, H. F. and Braak, C. J. F. (1999). Ranking of epiphytic lichen sensitivity to air pollution using survey data: a comparison of indicator scales. Lichenologist, 31, 27–39.Google Scholar
van Herk, C. M. (2002). Epiphytes on wayside trees as an indicator of eutrophication in the Netherlands. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 285–289. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Herk, C. M., Aptroot, A. and Dobben, H. F. (2002). Long-term monitoring in the Netherlands suggests that lichens respond to global warming. Lichenologist, 34, 141–154.CrossRefGoogle Scholar
Vězda, A. (1979). Flechtensystematische Studien. XI. Beiträge zur Kenntnis der Familie Asterothyriaceae (Discolichenes). Folia Geobotanica Phytotaxonomica Bohemoslovaca, Praha, 14, 43–94.CrossRefGoogle Scholar
Vězda, A. (1980). Foliikole Flechten aus Zaire. Die Arten der Sammelgattungen Catillaria und Bacidia. Folia Geobotanica Phytotaxonomica Bohemoslovaca, Praha, 15, 75–94.CrossRefGoogle Scholar
Villeneuve, J. P. and Holm, E. (1984). Atmospheric background of chlorinated hydrocarbons studied in Swedish lichens. Chemosphere, 13, 1133–1138.CrossRefGoogle Scholar
Virtala, M. (1992). Optimal harvesting of a plant–herbivore system: lichen and reindeer in northern Finland. Ecological Modelling, 60, 233–255.CrossRefGoogle Scholar
Vitikainen, O. (1998). Taxonomic notes on neotropical species of Peltigera. In Lichenology in Latin America: History, Current Knowledge and Applications, ed. Marcelli, M. P. and Seaward, M. R. D., pp. 135–139. São Paulo: CETESB.Google Scholar
Vitikainen, O. (2001). William Nylander (1822–1899) and lichen chemotaxonomy. Bryologist, 104, 263–267.CrossRefGoogle Scholar
Vitousek, P. M., Walker, L. R., Whittaker, L. D., Mueller-Dombois, D. and Matson, P. A. (1987). Biological invasion by Myrica faga alters ecosystem development in Hawaii. Science, 238, 802–804.CrossRefGoogle Scholar
Vobis, G. and Hawksworth, D. L. (1981). Conidial lichen-forming fungi. In The Biology of Conidial Fungi, ed. Cole, G. T. and Kendrick, B., pp. 245–273. New York: Academic Press.Google Scholar
Vogel, H. J. (1964). Distribution of lysine pathways among fungi: evolutionary implications. American Naturalist, 98, 435–446.CrossRefGoogle Scholar
Vogel, S. (1955). “Niedere Fensterpflanzen” in der südafrikanischen Wüste. Eine ökologische Schilderung. Beiträge zur Biologie der Pflanzen, 31, 45–135.Google Scholar
Arb, C., Mueller, C., Ammann, K. and Brunold, C. (1990). Lichen physiology and air pollution. II. Statistical analysis of the correlation between SO2, NO2, NO and O3, and chlorophyll content, net photosynthesis, sulphate uptake and protein synthesis of Parmelia sulcata Taylor. New Phytologist, 115, 431–437.CrossRefGoogle Scholar
Vráblíková, H., McEvoy, M., Solhaug, K. A., Barták, M. and Gauslaa, Y. (2006). Annual variation in photoacclimation and photoprotection of the photobiont in the foliose lichen Xanthoria parietina. Journal of Photochemistry and Photobiology B: Biology, 83, 151–162.CrossRefGoogle Scholar
Wachtmeister, C. A. (1956). Identification of lichen acids by paper chromatography, Botaniser Notiser, 109, 313–324.Google Scholar
Wainright, P. O., Hinkle, G., Sogin, M. L. and Stickel, S. K. (1993). Monophyletic origins of the Metazoa: an evolutionary link with fungi. Science, 260, 340–342.CrossRefGoogle ScholarPubMed
Walker, D. A., Webber, P. J., Everett, K. R. and Brown, J. (1978). Effects of crude and diesel oil spills on plant communities at Prudhoe Bay, Alaska, and the derivation of oil spill sensitivity maps. Arctic, 31, 242–259.CrossRefGoogle Scholar
Walker, F. J. (1985). The lichen genus Usnea subgenus Neuropogon. Bulletin of the British Museum (Natural History), 13, 1–130.Google Scholar
Walker, T. R., Crittenden, P. D. and Young, S. D. (2003). Regional variation in the chemical composition of winter snow pack and terricolous lichens in relation to sources of acid emissions in the Usa River basin, northeast European Russia. Environmental Pollution, 125, 401–412.CrossRefGoogle ScholarPubMed
Walser, J.-C., Holderegger, R., Gugerli, F., Hoebee, S. E. and Scheidegger, C. (2005). Microsatellites reveal regional population differentiation and isolation in Lobaria pulmonaria, an epiphytic lichen. Molecular Ecology, 14, 457–468.CrossRefGoogle ScholarPubMed
Walser, J.-C., Sperisen, C., Soliva, M. and Scheidegger, C. (2003). Fungus-specific microsatellite primers of lichens: application for the assessment of genetic variation on different spatial scales in Lobaria pulmonaria. Fungal Genetics and Biology, 40, 72–82.CrossRefGoogle ScholarPubMed
Warén, H. (1918–19) [1920]. Reinkulturen von Flechtengonidien. Öfversigt af Finska Vetenskaps-Societetens Förhandlingar (Helsingfors), 61, 1–79.Google Scholar
Waring, R. H. and Schlesinger, W. H. (1985). Forest Ecosystems: Concepts and Management. Orlando: Academic Press.Google Scholar
Warren, S. D. and Eldridge, D. J. (2001). Biological soil crusts and livestock in arid ecosystems: are they compatible?. In Biological Soil Crusts: Structure, Function and Management, ed. Belnap, J. and Lange, O. L., pp. 401–415. Berlin: Springer.Google Scholar
Washburn, S. (2005 [2006]). The Epiphytic Macrolichens of the Greater Cincinnati, Ohio, Metropolitan Area. M.S. thesis. Cincinnati: University of Cincinnati.
Washburn, S. (2006). Ozone exposure indices correlated with lichen abundance data in greater Cincinnati metropolitan area, Ohio. Botany 2006 Abstracts, p. 51. St. Louis: Botanical Society of America.
Watanabe, A. (1960). List of algal strains in collection at the Institute of Applied Microbiology, University of Tokyo. Journal of General and Applied Microbiology, 6, 283–292.CrossRefGoogle Scholar
Waterbury, J. B. and Stanier, R. Y. (1978). Patterns of growth and development in pleurocapsalean cyanobacteria. Microbiological Reviews, 42, 2–44.Google ScholarPubMed
Waters, J. M. and Craw, D. (2006). Goodby Gondwana? New Zealand biogeography, geology, and the problem of circularity. Systematic Biology, 55, 351–356.CrossRefGoogle Scholar
Weber, W. A. (1977). Environmental modification and lichen taxonomy. In Lichen Ecology, ed. Seaward, M. R. D., pp. 9–29. London: Academic Press.Google Scholar
Weber, W. A. (2003). The middle Asian element in the southern Rocky Mountain flora of the western United States: a critical biogeographical review. Journal of Biogeography, 30, 649–685.CrossRefGoogle Scholar
Wedin, M. (1995). The lichen family Sphaerophoraceae (Caliciales, Ascomycotina) in temperate areas of the Southern Hemisphere. Symbolae Botanicae Upsalienses, 31, 1–102.Google Scholar
Wedin, M. and Döring, H. (1999). The phylogenetic relationship of the Spaerophoraceae, Austropeltum and Neophyllis (lichenized Ascomycota) inferred by SSU rDNA sequences. Mycological Research, 109, 1131–1137.CrossRefGoogle Scholar
Wedin, M. and Wiklund, E. (2004). The phylogenetic relationships of Lecanorales suborder Peltigerineae revisited. Symbolae Botanicae Upsalienses, 34, 469–475.Google Scholar
Wedin, M., Döring, H. and Ekman, S. (2000 a). Molecular phylogeny of the lichen families Cladoniaceae, Sphaerophoraceae, and Stereocaulaceae (Lecanorales, Ascomycotina). Lichenologist, 32, 171–187.CrossRefGoogle Scholar
Wedin, M., Döring, H. and Gilenstam, G. (2004). Saprotrophy and lichenization as options for the same fungal species on different substrata: environmental plasticity and fungal lifestyles in the Stictis-Conotrema complex. New Phytologist, 164, 459–465.CrossRefGoogle Scholar
Wedin, M., Döring, H., Könberg, K. and Gilenstam, G. (2005 a). Generic delimitations in the family Stictidaceae (Ostropales, Ascomycota): the Stictis-Conotrema problem. Lichenologist, 37, 67–75.CrossRefGoogle Scholar
Wedin, M., Döring, H., Nordin, A. and Tibell, L. (2000 b). Small subunit rDNA phylogeny shows the lichen families Caliciaceae and Physciaceae (Lecanorales, Ascomycotina) to form a monophyletic group. Canadian Journal of Botany, 78, 246–254.CrossRefGoogle Scholar
Wedin, M., Wiklund, E., Crewe, A., et al. (2005 b). Phylogenetic relationships of Lecanoromycetes (Ascomycota) as revealed by analyses of mtSSU and nLSU rDNA sequence data. Mycological Research, 109, 159–172.CrossRefGoogle ScholarPubMed
Wedin, M., Wiklund, E. and Jørgensen, P. M. (2007). Massalongiaceae, fam. nov., an overlooked monophyletic group among the cyanobacterial lichens (Peltigerales, Lecanoromycetes, Ascomycota). Lichenologist, 39, 61–67.CrossRefGoogle Scholar
Wein, R. W. and Speer, J. E. (1975). Lichen biomass in Acadian and boreal forests of Cape Breton Island, Nova Scotia. Bryologist, 78, 328–333.CrossRefGoogle Scholar
Weissman, J. C. and Benemann, J. R. (1977). Hydrogen products by nitrogen-starved cultures of Anabaena cylindrica. Applied Environmental Microbiology, 33, 123–131.Google Scholar
Weissman, L., Garty, J. and Hochman, A. (2005 a). Rehydration of the lichen Ramalina lacera results in production of reactive oxygen species and nitric oxide and a decrease in antioxidants. Applied and Environmental Microbiology, 71, 2121–2129.CrossRefGoogle Scholar
Weissman, L., Garty, J. and Hochman, A. (2005 b). Characterization of enzymatic antioxidants in the lichen Ramalina lacera and their response to rehydration. Applied and Environmental Microbiology, 71, 6508–6514.CrossRefGoogle ScholarPubMed
Werth, S., Wagner, H. H., Holderegger, J. M., Kalwij, J. and Scheidegger, C. (2005). Genetic diversity of an old-forest associated lichen is affected by stand-replacing disturbances. In Dispersal and Persistence of an Epiphytic Lichen in a Dynamic Pasture-woodland Landscape. Ph.D. dissertation, by S. Werth, pp. 23–48. Bern: Universität Bern.Google Scholar
Wessels, D. C. J. and Schoeman, P. (1988). Mechanism and rate of weathering of Clarens sandstone by an endolithic lichen. South African Journal of Science, 84, 274–277.Google Scholar
Wessels, D. C. J. and Wessels, L. A. (1991). Erosion of biogenically weathered Clarens sandstone by lichenophagous bagworm larvae (Lepidoptera: Psychidae). Lichenologist, 23, 283–291.CrossRefGoogle Scholar
Wessels, J. G. H. (1999). Fungi in their own right. Fungal Genetics and Biology, 27, 134–145.CrossRefGoogle ScholarPubMed
Wetmore, C. M. (1973). Multiperforate septa in lichens. New Phytologist, 72, 535–538.CrossRefGoogle Scholar
White, F. J. and James, P. W. (1985). A new guide to microchemical techniques for the identification of lichen substances. British Lichen Society Bulletin, 57 (supp.), 1–41.Google Scholar
Whiteford, J. and Spanu, P. (2002). Hydrophobins and the interactions between fungi and plants. Molecular Plant Pathology, 3, 391–400.CrossRefGoogle ScholarPubMed
Whiton, J. C. and Lawrey, J. D. (1984). Inhibition of crustose lichen spore germination by lichen acids. Bryologist, 87, 42–43.CrossRefGoogle Scholar
Wiklund, E. and Wedin, M. (2003). The phylogenetic relationships of the cyanobacterial lichens in the Lecanorales suborder Peltigerineae. Cladistics, 19, 419–431.CrossRefGoogle Scholar
Will-Wolf, S. (1980). Structure of corticolous lichen communities before and after exposure to emissions from a “clean” coal-fired generating station. Bryologist, 83, 281–295.CrossRefGoogle Scholar
Will-Wolf, S. (2002). Monitoring regional status and trends in forest health with lichen communities: the United States Forest Service approach. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 353–357. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Wilmotte, A. and Golubic, S. (1991). Morphological and genetic criteria in the taxonomy of Cyanophyta/Cyanobacteria. Algological Studies, 64, 1–24.Google Scholar
Winchester, V. and Harrison, S. (1994). A development of the lichenometric method applied to the dating of glacially influenced debris flows in southern Chile. Earth Surface Processes and Landforms, 19, 137.CrossRefGoogle Scholar
Winkworth, R. C., Wagstaff, S. J., Glenny, D. and Lockhart, P. J. (2002). Plant dispersal N.E.W.S. from New Zealand. Trends in Ecology and Evolution, 17, 514–520.CrossRefGoogle Scholar
Winkworth, R. C., Wagstaff, S. J., Glenny, D. and Lockhart, P. J. (2005). Evolution of the New Zealand mountain flora: origins, diversification and dispersal. Organisms, Diversity and Evolution, 5, 237–247.CrossRefGoogle Scholar
Winner, W. E., Atkinson, C. J. and Nash, T. H. III. (1988). Comparisons of SO2 absorption capacities of mosses, lichens, and vascular plants in diverse habitats. Bibliotheca Lichenologica, 30, 217–230.Google Scholar
Wirth, V. (1972). Die Silikatflechten – Gemeinschaften im ausseralpinen Zentraleuropa. Dissertationes Botanicae, 17, 1–306.Google Scholar
Wirth, V. (1987). The influence of water relations on lichen SO2-resistance. In Progress and Problems in Lichenology in the Eighties, ed. Peveling, E., pp. 347–350. Bibliotheca Lichenologica 25. Berlin-Stuttgart: J. Cramer.Google Scholar
Wirth, V. (1993). Trenwende bei der Ausbreitung der anthropogen geförderten Flechte Lecanora conizaeoides? Phytocoenologia, 23, 625–636.CrossRefGoogle Scholar
Wirth, V. (1995) Die Flechten Baden-Würtembergs. Verbreitungsatlas. Vols. I and II. Stuttgart: Eugen Ulmer.Google Scholar
Wirth, V. (2001). Zeiberwerte von Flechten. Scripta Geobotanica, 18, 221–243.Google Scholar
Wirtz, N., Lumbsch, H. T., Green, T. G. A., et al. (2003). Lichen fungi have low cyanobiont selectivity in maritime Antarctica. New Phytologist, 160, 177–183.CrossRefGoogle Scholar
Wirtz, N., Printzen, C., Sancho, L. G. and Lumbsch, H. T. (2006). The phylogeny and classification of Neuropogon and Usnea (Parmeliaceae, Ascomycota) revisited. Taxon, 55, 367–376.CrossRefGoogle Scholar
Wise, M. J. and Tunnacliffe, A. (2004). POPP the question: what do LEA proteins do? Trends in Plant Science, 9, 13–17.CrossRefGoogle Scholar
Wolf, J. H. D. (1993). Diversity patterns and biomass of epiphytic bryophytes and lichens along an altitudinal gradient in the northern Andes. Annals of the Missouri Botanical Garden, 80, 928–960.CrossRefGoogle Scholar
Wolk, C. P., Ernst, A. and Elhai, J. (1994). Heterocyst metabolism and development. In Molecular Genetics of Cyanobacteria, ed. Bryant, D., pp. 769–823. Dordrecht: Kluwer Academic Publishing.CrossRefGoogle Scholar
Wolken, G. J. (2006). High-resolution multispectral techniques for mapping former Little Ice Age terrestrial ice cover in the Canadian High Arctic. Remote Sensing of Environment, 101, 104–114.CrossRefGoogle Scholar
Wolken, G. J., England, J. H. and Dyke, A. S. (2005). Re-evaluating the relevance of vegetation trimlines in the Canadian Arctic as an indicator of Little Ice Age paleoenvironments. Arctic, 58, 341–353.Google Scholar
Wolseley, P. A. (1995). A global perspective on the status of lichens and their conservation. Mitteilungen der Eidgenössischen Forshungsanstalt für Wald, Schnee und Landschaft, 70, 11–22.Google Scholar
Wolseley, P. A. (2002). Using corticolous lichens of tropical forests to assess environmental changes. In Monitoring with Lichens – Monitoring Lichens, ed. Nimis, P. L., Scheidegger, C. and Wolseley, P. A., pp. 373–378. Nato Science Series IV: Earth and Environmental Sciences. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Wolseley, P. A., James, P. W., Theobald, M. R. and Sutton, M. A. (2006). Detecting changes in epiphytic lichen communities at sites affected by atmospheric ammonia from agricultural sources. Lichenologist, 38, 161–176.CrossRefGoogle Scholar
Wösten, H. A. B. (2001). Hydrophobins: multipurpose proteins. Annual Reviews in Microbiology, 55, 625–646.CrossRefGoogle ScholarPubMed
Wösten, H. A. B., Devries, O. M. H. and Wessels, J. G. H. (1993). Interfacial self-assembly of a fungal hydrophobin into a hydrophobic rodlet layer. Plant Cell, 5, 1567–1574.CrossRefGoogle ScholarPubMed
Yahr, R., Vilgalys, R. and DePriest, P. T. (2004). Strong fungal specificity and selectivity for algal symbionts in Florida scrub Cladonia lichens. Molecular Ecology, 13, 3367–3378.CrossRefGoogle ScholarPubMed
Yahr, R., Vilgalys, R. and DePriest, P. T. (2006). Geographic variation in algal partners of Cladonia subtenuis (Cladoniaceae) highlights the dynamic nature of a lichen symbiosis. New Phytologist, 171, 847–860.CrossRefGoogle ScholarPubMed
Yamamoto, Y. (1990). Studies of Cell Aggregates and the Production of Natural Pigments in Plant Cell Culture. Osaka: Nippon Paint Publication.Google Scholar
Yamamoto, Y., Kinoshita, Y. and Yoshimura, I. (2002). Culture of thallus fragments and redifferentiation of lichens. In Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 34–46. Berlin: Springer.CrossRefGoogle Scholar
Yoshimura, I. (1998). Lobaria in Latin America: taxonomic, geographic and evolutionary aspects. In Lichenology in Latin America: History, Current Knowledge and Applications, ed. Marcelli, M. P. and Seaward, M. R. D., pp. 129–134. São Paulo: CETESB.Google Scholar
Yoshimura, I. and Arvidsson, L. (1994). Taxonomy and chemistry of the Lobaria crenulata group in Ecuador. Acta Botanica Fennica, 150, 223–233.Google Scholar
Yoshimura, I., Yamamoto, Y., Nakano, T. and Finnie, J. (2002). Isolation and culture of lichen photobionts and mycobionts. Protocols in Lichenology: Culturing, Biochemistry, Ecophysiology and Use in Biomonitoring (Springer Lab Manual), ed. Kranner, I., Beckett, R. P. and Varma, A. K., pp. 3–33. Berlin: Springer:CrossRefGoogle Scholar
Young, K. (2005). Hardy lichen shown to survive in space. New Scientist, November 10th, 2005.
Yuan, X., Xiao, S. and Taylor, T. N. (2005). Lichen-like symbiosis 600 million years ago. Science, 308, 1017–1020.CrossRefGoogle ScholarPubMed
Yun, S., Berbee, M., Yoder, O. and Turgeon, B. (1999). Evolution of the fungal self-fertile reproductive life style from self-sterile ancestors. Proceedings of the National Academy of Sciences, USA, 96, 5592–5597.CrossRefGoogle ScholarPubMed
Yurchenko, E. and Golubkov, V. (2003). The morphology, biology, and geography of a necrotrophic basidiomycete Athelia arachnoidea in Belarus. Mycological Progress, 2, 275–284.CrossRefGoogle Scholar
Zachariassen, K. E. and Kristiansen, E. (2000). Ice nucleation and antinucleation in nature. Cryobiology, 41, 257–279.CrossRefGoogle ScholarPubMed
Zahlbruckner, A. (1907). Die naturlichen Pflanzenfamilien, I Teil, 1 Apt, ed. Engler, A. and Prantl, K.. Leipzig: Borntraeger.Google Scholar
Zahlbruckner, A. (1926). Die naturlichen Pflanzenfamilien, Vol. 8, ed. Engler, A. and Prantl, K.. Leipzig: Borntraeger.Google Scholar
Zambrano, A., Nash, T. H. III and Gries, C. (2000). Responses of Ramalina farinacea (L.) Ach. to transplanting in southern California and to gaseous formaldehyde. Bibliotheca Lichenologica, 75, 219–230.Google Scholar
Zavarzina, A. and Zavarzin, A. (2006). Laccase and tyrosinase activity in lichens. Microbiology (Rus), 75, 630–641.Google ScholarPubMed
Zeitler, I. (1954). Untersuchungen über die Morphologie, Entwicklungsgeschichte und Systematik von Flechtengonidien. Österreichische Botanische Zeitschrift, 101, 453–487.CrossRefGoogle Scholar
Zhu, Y. Y., Machleder, E. M., Chenchik, A., Li, R. and Siebert, P. M. (2001). Reverse transcriptase template switching: a SMART™ approach for full-length cDNA library construction. BioTechniques, 30, 892–897.Google Scholar
Ziegler, I. (1977). Sulfite action on ribulosediphosphate carboxylase in the lichen Pseudevernia furfuracea. Oecologia, 29, 63–66.CrossRefGoogle ScholarPubMed
Zoller, S. and Lutzoni, F. (2003). Slow algae, fast fungi: exceptionally high nucleotide substitution rate differences between lichenized fungi Omphalina and their symbiotic green algae Coccomyxa. Molecular Phylogenetics and Evolution, 29, 629–640.CrossRefGoogle ScholarPubMed
Zoller, S., Lutzoni, F. and Scheidegger, C. (1999). Genetic variation within and among populations of the threatened lichen Lobaria pulmonaria in Switzerland and implications for its conservation. Molecular Ecology, 8, 2049–2059.CrossRefGoogle ScholarPubMed
Zopf, W. (1897). Über Nebensymbiose (Parasymbiose). Berichte der deutschen Botanischen Gesellschaft, 15, 90–92.Google Scholar
Zopf, W. (1907). Die Flechtenstoffe in chemischer, botanischer, pharmakologischer, und technischer Beziehung. Jena: Gustav Fischer.Google Scholar
Zotz, G. (1999). Altitudinal changes in diversity and abundance of non-vascular epiphytes in the tropics – an ecophysiological explanation. Selbyana, 20, 256–260.Google Scholar
Zotz, G. and Winter, K. (1994). Photosynthesis and carbon gain of the lichen, Leptogium azureum, in a lowland tropical forest. Flora, 189, 179–186.CrossRefGoogle Scholar
Zotz, G., Büdel, B., Meyer, A., Zellner, H. and Lange, O. L. (1998). In situ studies of water relations and CO2 exchange of the tropical macrolichen, Sticta tomentosa. New Phytologist, 139, 525–535.CrossRefGoogle Scholar
Zotz, G., Schultz, S. and Rottenberger, S. (2003). Are tropical lowlands a marginal habitat for macrolichens? Evidence from a field study with Parmotrema endosulphureum in Panama. Flora, 198, 71–77.CrossRefGoogle Scholar
Zukal, H. (1895). Morphologische und biologische Untersuchungen über die Flechten (II. Abhandlung). Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften Wien, Math.-naturw. Cl., 104/1, 1–68.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Edited by Thomas H. Nash, III, Arizona State University
  • Book: Lichen Biology
  • Online publication: 05 September 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511790478.020
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Edited by Thomas H. Nash, III, Arizona State University
  • Book: Lichen Biology
  • Online publication: 05 September 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511790478.020
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Edited by Thomas H. Nash, III, Arizona State University
  • Book: Lichen Biology
  • Online publication: 05 September 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511790478.020
Available formats
×