Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-t6hkb Total loading time: 0 Render date: 2024-07-11T04:54:32.257Z Has data issue: false hasContentIssue false

2 - Bryophyte Physiological Processes in a Changing Climate: an Overview

Published online by Cambridge University Press:  05 October 2012

Nancy G. Slack
Affiliation:
Sage Colleges, New York
Lloyd R. Stark
Affiliation:
University of Nevada, Las Vegas
Get access

Summary

Climate change as a result of global warming is predicted to be most pronounced at high latitudes. It is known from experimental studies that Sphagnum species respond to enhanced UV radiation by decreasing their growth (Huttunen et al. 2005). Some polar bryophytes reproduce sexually and form sporulating sporophytes, e.g., Polytrichum hyperboreum on the Svalbard tundra. Antarctic mosses tend to reproduce sexually more often at higher Antarctic latitudes (Lewis-Smith & Convey 2002). Simultaneously, increased emissions of nitrogenous air pollutants cause increased nitrogen deposition over the northern hemisphere (Bouwman et al. 2002).

Mires, i.e., wetlands actively accumulating organic material, are believed to play an important role in the global biogeochemical carbon cycle, potentially serving as major long-term carbon sinks. Vegetation structure, however, strongly influences the carbon sink capacity of mires (Malmer & Wallén 2005). In general, carbon accumulation is greater in Sphagnum-dominated mires than in sedge-dominated mires (Dorrepaal et al. 2005).

Graminoid dominance, however, leads to increased methane (CH4) emissions, due both to increased root exudation of precursor compounds for CH4 formation and to plant-mediated transport of CH4 to the atmosphere through aerenchyma tissue, bypassing oxidation in the acrotelm. Because of such feedbacks on global carbon budgets, investigations of vegetation dynamics in response to ongoing pollution and warming are crucial for estimates of future scenarios of global change (Wiedermann et al. 2007).

There is great uncertainty in the estimations of the impacts and possible outcome of global climate changes.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2011

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

Asakawa, Y. (1995). In Progress in the Chemistry of Organic Natural Products, ed. Herz, W., Kirby, G. W., Moore, R. E., Steglich, W. & Tamm, C., vol. 65, p. 1. Vienna: Springer.CrossRef
Barker, D. H., Stark, L. R., Zimpfer, J. F., McLetchie, N. D. & Smith, S. D. (2005). Evidence of drought-induced stress on biotic crust moss in the Mojave Desert. Plant, Cell and Environment 28: 939–47.Google Scholar
Beckett, R. P., Csintalan, Z. & Tuba, Z. (2000). ABA treatment increases both the desiccation tolerance of photosynthesis, and nonphotochemical quenching in the moss Atrichum undulatum. Plant Ecology 151: 65–71.Google Scholar
Beckett, R. P., Marschall, M. & Laufer, Zs. (2005). Hardening enhances photoprotection in the moss Atrichum androgynum during rehydration by increasing fast rather than slow-relaxing quenching. Journal of Bryology 27: 7–12.Google Scholar
Berner, R. A. (1998). The carbon cycle and CO2 over Phanerozoic time: the role of land plants. Philosophical Transactions of the Royal Society of London B 353: 75–82.Google Scholar
Bonine, M. L. (2004). Growth, reproductive phenology, and population structure in Syntrichia caninervis. M.S. thesis, University of Nevada, Las Vegas.
Bouwman, A. F., Vuuren, D. P., Derwent, R. G. & Posch, M. (2002). A global analysis of acidification and eutrophication of terrestrial ecosystems. Water, Air, and Soil Pollution 141: 349–82.Google Scholar
Breeuwer, A., Heijmans, M. M. P. D., Robroek, B. J. M. & Berendse, F. (2008). The effect of temperature on growth and competition between Sphagnum species. Oecologia 156: 155–67.Google Scholar
Carballeira, A., Díaz, S., Vázquez, M. D. & López, J. (1998). Inertia and resilience in the response of the aquatic bryophyte Fontinalis antipyretica Hedw. to thermal stress. Archives of Environmental Contamination and Toxicology 34: 343–9.Google Scholar
Chalker-Scott, L. (1999). Environmental significance of anthocyanins in plant stress responses. Photochemistry & Photobiology 70: 1–9.Google Scholar
Crosby, M. R., Magill, R. E., Allen, B. & He, S. (2000). A checklist of the mosses. St. Louis, MO: Missouri Botanical Garden. www.mobot.org/MOBOT/tropicos/most/checklist.shtml.
Csintalan, Z., Tuba, Z. & Laitat, E. (1995). Slow chlorophyll fluorescence, net CO2 assimilation and carbohydrate responses in Polytrichum formosum to elevated CO2 concentrations. In Photosynthesis from Light to Biosphere, ed. Mathis, P., Vol. V, pp. 925–8. Dordrecht: Kluwer Academic Publishers.CrossRef
Csintalan, Z., Tuba, Z., Takács, Z.et al. (2001). Responses of nine bryophyte and one lichen species from different microhabitats to elevated UV-B radiation. Photosynthetica 39: 317–20.Google Scholar
Davidson, A. J., Harborne, J. B. & Longton, R. E. (1989). Identification of hydroxycinnamic acid and phenolic acids in Mnium hornum and Brachythecium rutabulum and their possible role in protection against herbivory. Journal of the Hattori Botanical Laboratory 67: 415–22.Google Scholar
Deltoro, V. I., Calatayud, A., Gimeno, C. & Barreno, E. (1998). Water relations, chlorophyll fluorescence, and membrane permeability during desiccation in bryophytes from xeric, mesic, and hydric environments. Canadian Journal of Botany 76(11): 1923–9.Google Scholar
Dorrepaal, E., Cornelissen, J. H. C., Aerts, R., Wallen, B. & Logtestijn, R. S. P. (2005). Are growth forms consistent predictors of leaf litter quality and decomposability across peatlands along a latitudinal gradient? Journal of Ecology 93: 817–28.Google Scholar
Dunn, J. L. & Robinson, S. A. (2006). UV-B screening potential is higher in two cosmopolitan moss species than in a co-occurring Antarctic endemic moss – implications of continuing ozone depletion. Global Change Biology 12(12): 2282–96.Google Scholar
Gunnarsson, U., Granberg, G. & Nilsson, M. B. (2004). Growth, production and interspecific competition in Sphagnum: effects of temperature, nitrogen and sulfur treatments on a boreal mire. New Phytologist 163: 349–59.Google Scholar
Hada, H., Hidema, J., Maekawa, M.et al. (2003) Higher amounts of anthocyanins and UV absorbing compounds effectively lowered CPD photorepair in purple rice (Oryza sativa L.). Plant Cell and Environment 26: 1691–1701.CrossRefGoogle Scholar
Harvell, D., Mitchell, C. E., Ward, J. R.et al. (2002). Climate warming and disease risks for terrestrial and marine biota. Science 296: 2158–62.Google Scholar
Hearnshaw, G. F. & Proctor, M. C. F. (1982). The effect of temperature on the survival of dry bryophytes. New Phytologist 90: 221–8.Google Scholar
Heijmans, M. M. P., Klees, D. & Berendse, F. (2002).Competition between Sphagnum magellanicum and Eriophorum angustifolium as affected by raised CO2 and increased N deposition. Oikos 97: 415–25.Google Scholar
Houghton, J. T., Jenkins, G. J. & Ephramus, J. J. (1990). Climate Change: The IPCC Scientific Assessment. Cambridge: Cambridge University Press.
Huttunen, S., Lappalainen, N. M. & Turunen, J. (2005). UV-absorbing compounds in subarctic herbarium bryophytes. Environmental Pollution 133: 303–14.Google Scholar
Hyvönen, R., Ågren, G. I., Linder, S., et al. (2006). The likely impact of elevated [CO2], nitrogen deposition, increased temperature and management on carbon sequestration in temperate and boreal forest ecosystems: a literature review. New Phytologist 173: 463–80.Google Scholar
,IPCC (2001). Climate Change 2001: The Scientific Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernment Panel on Climate Change, ed. Houghton, J. T., Ding, Y., Griggs, D. J.et al. Cambridge and New York: Cambridge University Press.
Jauhiainen, J., Silvola, J. & Vasander, H. (1998). Effects of increased carbon dioxide and nitrogen supply on mosses. In Bryology for the Twenty-first Century, ed. Bates, J. W., Ashton, N. W. & Duckett, J. G., pp. 343–60. Leeds: Maney Publishing and the British Bryological Society.
Körner, C. & Bazzaz, F. A. (1996). Carbon Dioxide, Populations, and Communities. San Diego, CA: Academic Press.
Lewis-Smith, R. I. & Convey, P. (2002). Enhanced sexual reproduction in bryophytes at high latitudes in the maritime Antarctic. Journal of Bryology 24: 107–17.Google Scholar
Limpens, J. & Berendse, F. (2003a). Growth reduction of Sphagnum magellanicum subjected to high nitrogen deposition: the role of amino acid nitrogen concentration. Oecologia 135: 339–45.Google Scholar
Limpens, J. & Berendse, F. (2003b). How litter quality affects mass loss and N loss from decomposing Sphagnum. Oikos 103: 537–47.Google Scholar
Lud, D., Moerdijk, T. C. W., Poll, W. H.et al. (2002). DNA Damage and photosynthesis in Antarctic and Arctic Sanionia uncinata (Hedw.) Loeske under ambient and enhanced levels of UV-B radiation. Plant, Cell and Environment 25: 1579–89.Google Scholar
Lud, D., Schlensog, M., Schroeter, B. & Huiskes, A. H. L. (2003). The influence of UV-B radiation on light-dependent photosynthetic performance in Sanionia uncinata (Hedw.) Loeske in Antarctica. Polar Research 26: 225–32.Google Scholar
Lüttge, U., Meirelles, S. T. & Mattos, E. A. (2008). Strong quenching of chlorophyll fluorescence in the desiccated state in three poikilohydric and homoiochlorophyllous moss species indicates photo-oxidative protection on highly light-exposed rocks of a tropical inselberg. Journal of Plant Physiology 165: 172–81.Google Scholar
Malmer, N. & Wallén, B. (2005). Nitrogen and phosphorus in mire plants: variation during 50 years in relation to supply rate and vegetation type. Oikos 109: 539–54.Google Scholar
Marschall, M. & Beckett, R. P. (2005). Photosynthetic responses in the inducible mechanisms of desiccation tolerance of a liverwort and a moss. Acta Biologica Szegediensis 49 (1–2): 155–6.Google Scholar
Marschall, M. & Proctor, M. C. F. (2004). Are bryophytes shade plants? Photosynthetic light responses and proportions of chlorophyll a, chlorophyll b and total carotenoids. Annals of Botany doi:10.1093/aob/mch178, available online at www.aob.oupjournals.org.CrossRef
Mayaba, N. & Beckett, R. P. (2003). Increased activities of superoxide dismutase and catalase are not the mechanism of desiccation tolerance induced by hardening in the moss Atrichum androgynum. Journal of Bryology 25: 281–6.Google Scholar
McLetchie, D. N. & Stark, L. R. (2006). Sporophyte and gametophyte generations differ in their thermotolerance response in the moss Microbryum. Annals of Botany 97: 505–11.Google Scholar
Medlyn, B. E., Berbigier, P., Clement, R.et al. (2005). The carbon balance of coniferous forests growing in contrasting climatic conditions: a model-based analysis. Agricultural and Forest Meteorology 131: 97–124.Google Scholar
Meyer, H. & Santarius, K. A. (1998). Short-term thermal acclimation and heat tolerance of gametophytes of mosses. Oecologia 115: 1–8.Google Scholar
Mori, M., Yoshida, K., Ishigaki, Y.et al. (2005). UV-B protective effect of a polyacylated anthocyanin, HBA, in flower petals of the blue morning glory, Ipomoea tricolor cv. Heavenly Blue. Bioorganic & Medicinal Chemistry 13: 2015–20.Google Scholar
Mues, R. (2000). Chemical constituents and biochemistry. In Bryophyte Biology, ed. Shaw, A. J. & Goffinet, B., pp. 150–81. Cambridge: Cambridge University Press.CrossRef
Nabe, H., Funabiki, R., Kashino, Y., Koike, H. & Satoh, K. (2007). Responses to desiccation stress in bryophytes and an important role of dithiothreitol-insensitive non-photochemical quenching against photoinhibition in dehydrated states. Plant and Cell Physiology 48: 1548–57.Google Scholar
Newsham, K. K. (2003). UV-B radiation arising from stratospheric ozone depletion influences the pigmentation of the Antarctic moss Andreaea regularis. Oecologia 135: 327–31.Google Scholar
Nobel, P. S., Geller, G. N., Kee, Z. C. & Zimmerman, A. D. (1986). Temperatures and thermal tolerances for cacti exposed to high temperatures near the soil surface. Plant, Cell and Environment 9: 279–87.Google Scholar
Nordbakken, J. F., Ohlson, M. & Högberg, P. (2003). Boreal bog plants: nitrogen sources and uptake of recently deposited nitrogen. Environmental Pollution 126: 191–200.Google Scholar
Oechel, W. C. & Svejnbjörnsson, B. (1978). Primary production processes in arctic bryophytes at Barrow, Alaska. In Vegetation and Production Ecology of an Alaskan Arctic Tundra, ed. Tieszen, L. L., pp. 269–98. New York: Springer.CrossRef
Oliver, M. J., Dowd, S. E., Zaragoza, J., Mauget, S. A. & Payton, P. R. (2004). The rehydration transcriptome of the desiccation-tolerant bryophyte Tortula ruralis: transcript classification and analysis. BMC Genomics 2004 5: 89.Google Scholar
Oliver, M. J., Tuba, Z. & Mishler, B. D. (2000). The evolution of vegetative desiccation tolerance in land plants. Plant Ecology 151: 85–100.Google Scholar
Oliver, M. J., Velten, J. J. & Mishler, B. D. (2005). Desiccation tolerance in bryophytes: a reflection of the primitive strategy for plant survival in dehydrating habitats? Integrative and Comparative Biology 45: 788–99.Google Scholar
Ötvös, E. & Tuba, Z. (2005). Ecophysiology of mosses under elevated air CO2 concentration: overview. Physiology and Molecular Biology of Plants 11: 65–70.Google Scholar
Proctor, M. C. F. (1990). The physiological basis of bryophyte production. Botanical Journal of the Linnean Society 104: 61–77.Google Scholar
Proctor, M. C. F. (2004). How long must a desiccation-tolerant moss tolerate desiccation? Some results of 2 years' data logging on Grimmia pulvinata. Physiologia Plantarum 122: 21–7.Google Scholar
Proctor, M. C. F. & Pence, V. C. (2002). Vegetative tissues: bryophytes, vascular ‘resurrection plants’ and vegetative propagules. In Desiccation and Plant Survival, ed. Pritchard, H. & Black, M., pp. 207–37. Wallingford, UK: CABI Publishing.CrossRef
Proctor, M. C. F. & Smirnoff, N. (2000). Rapid recovery of photosystems on rewetting desiccation-tolerant mosses: chlorophyll fluorescence and inhibitor experiments. Journal of Experimental Botany 51: 1695–1704.Google Scholar
Proctor, M. C. F. & Tuba, Z. (2002). Poikilohydry and homoihydry: antithesis or spectrum of possibilities? Tansley review no. 141. New Phytologist 156: 27–49.Google Scholar
Robinson, S. A., Turnbull, J. D. & Lovelock, C. E. (2005). Impact of changes in natural ultraviolet radiation on pigment composition, physiological and morphological characteristics of the Antarctic moss, Grimmia antarctici. Global Change Biology 11: 476–89.Google Scholar
Robson, M. T., Pancotto, V. A., Flint, S. D., et al. (2003). Six years of solar UV-B manipulations affect growth of Sphagnum and vascular plants in a Tierra del Fuego peatland. New Phytologist 160: 379–89.Google Scholar
Roy, H. E., Brown, P. M. J., Rothery, P., Ware, R. L. & Majerus, M. E. N. (2008). Interactions between the fungal pathogen Beauveria bassiana and three species of coccinellid: Harmonia axyridis, Coccinella septempunctata and Adalia bipunctata. BioControl 53: 265–76.Google Scholar
Rozema, J., Boelen, P. & Blokker, P. (2005). Depletion of stratospheric ozone over the Antarctic and Arctic: responses of plants of polar terrestrial ecosystems to enhanced UV-B, an overview. Environmental Pollution 137: 428–42.Google Scholar
Rozema, J., Björn, L. O., Bornman, J. F.et al. (2002). The role of UV-B radiation in aquatic and terrestrial ecosystems – an experimental and functional analysis of the evolution of UV-absorbing compounds. Journal of Photochemistry and Photobiology B 66: 2–12.Google Scholar
Schlesinger, W. H. (1997). Biogeochemistry: an Analysis of Global Change, 2nd edn. San Diego, CA: Academic Press.
Searles, P. S., Flint, S. D., Diaz, S. B.et al. (2002). Plant response to solar ultraviolet-B radiation in a southern South American Sphagnum peatland. Journal of Ecology 90: 704–13.Google Scholar
Shaw, A. J. & Beer, S. C. (1997). Gametophyte–sporophyte variation and covariation in mosses. Advances in Bryology 6: 35–63.Google Scholar
Stark, L. R. (1997). Phenology and reproductive biology of Syntrichia inermis (Bryopsida, Pottiaceae) in the Mojave Desert. Bryologist 100: 13–27.Google Scholar
Stark, L. R. (2002). Skipped reproductive cycles and extensive sporophyte abortion in the desert moss Tortula inermis correspond to unusual rainfall patterns. Canadian Journal of Botany 80: 533–42.Google Scholar
Stark, L. R., NicholsII, L., McLetchie, D. N. & Bonine, M. L. (2005). Do the sexes of the desert moss Syntrichia caninervis differ in desiccation tolerance? A leaf regeneration assay. International Journal of Plant Sciences 166: 21–9.Google Scholar
Stark, L. R., Oliver, M. J., Mishler, B. D. & McLetchie, D. N. (2007). Generational differences in response to desiccation stress in the desert moss Tortula inermis. Annals of Botany 99: 53–60.Google Scholar
Strengbom, J., Nordin, A., Näsholm, T. & L. Ericson, L. (2002). Parasitic fungus mediates change in nitrogen-exposed boreal forest vegetation. Journal of Ecology 90: 61–7.Google Scholar
Taalas, P., Koskela, T., Kyro, E., Damski, J. & Supperi, A. (1995). Ultraviolet Radiation in Finland. In The Finnish Research Programme on Climate Change, Final Report, ed. Roos, J.. Publications of The Academy of Finland4: 83–91.
Taipale, T. & Huttunen, S. (2002). Moss flavonoids and their ultrastructural localization under enhanced UV-B radiation. Polar Research 38: 211–18.Google Scholar
Takács, Z., Lichtenthaler, H. K. & Tuba, Z. (2000). Fluorescence emission spectra of desiccation-tolerant cryptogamic plants during a rehydration – desiccation cycle. Journal of Plant Physiology 156: 375–9.Google Scholar
Throop, H. L. & Lerdau, M. T. (2004). Effects of nitrogen deposition on insect herbivory: implications for community and ecosystem processes. Ecosystems 7: 109–33.Google Scholar
Tuba, Z. (1987). Light, temperature and desiccation responses of CO2-exchange in desiccation tolerant moss, Tortula ruralis. In Proceedings of the IAB Conference of Bryoecology, ed. Pócs, T., Simon, T., Tuba, Z. & Podani, J., pp. 137–50. Symp. Biol. Hung. Vol. 35, Part A. Budapest: Akadémiai Kiadó.
Tuba, Z. (2008). Notes on the poikilochlorophyllous desiccation-tolerant plants. Acta Biologica Szegediensis 52 (1): 111–13.Google Scholar
Tuba, Z., Csintalan, Zs. & Proctor, M. C. F. (1996). Photosynthetic responses of a moss, Tortula ruralis (Hedw.) Gaertn. et al. ssp. ruralis, and the lichens Cladonia convoluta (Lam.) P. Cout. and C. furcata (Huds.) Schrad. to water deficit and short periods of desiccation, and their eco-physiological significance: a baseline study at present-day CO2 concentration. New Phytologist 133: 353–61.Google Scholar
Tuba, Z., Csintalan, Zs., Szente, K., Nagy, Z. & Grace, J. (1998). Carbon gains by desiccation tolerant plants at elevated CO2. Functional Ecology 12: 39–44.Google Scholar
Tuba, Z., Proctor, M. C. F. & Takács, Z. (1999). Desiccation-tolerant plants under elevated air CO2: a review. Zeitschrift für Naturforschung 54: 788–96.Google Scholar
Velten, J. & Oliver, M. J. (2001). Tr288, a rehydrin with a dehydrin twist. Plant Molecular Biology 45: 713–22.Google Scholar
Wasley, J., Robinson, S. A., Lovelock, C. E.et al. (2006). Some like it wet – an endemic Antarctic bryophyte likely to be threatened under climate change induced drying. Functional Plant Biology 33: 443–55.Google Scholar
Wiedermann, M. M., Nordin, A., Gunnarsson, U., Nilsson, M. B. & Ericson, L. (2007). Global change shifts vegetation and plant – parasite interactions in a boreal mire. Ecology 88: 454–64.Google Scholar
,WMO. (2006). WMO Greenhouse Gas Bulletin 1. Geneva: World Meteorological Organization/Global Atmosphere Watch. http://www.wmo.ch/web/arep/gaw/ghg/ghg-bulletin-en-03–06.pdf
Zotz, G., Schweikert, A., Jetz, W. & Westerman, H. (2000). Water relations and carbon gain are closely related to cushion size in the moss Grimmia pulvinata. New Phytologist 148: 59–67.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.

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.

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.

Available formats
×