Hostname: page-component-7c8c6479df-r7xzm Total loading time: 0 Render date: 2024-03-28T12:01:27.889Z Has data issue: false hasContentIssue false

ARE LICHEN GROWTH FORM CATEGORIES SUPPORTED BY CONTINUOUS FUNCTIONAL TRAITS: WATER-HOLDING CAPACITY AND SPECIFIC THALLUS MASS?

Published online by Cambridge University Press:  06 August 2019

S. Wan
Affiliation:
Royal Botanic Garden Edinburgh, 20A Inverleith Row, Edinburgh EH3 5LR, Scotland, UK.
C. J. Ellis*
Affiliation:
Royal Botanic Garden Edinburgh, 20A Inverleith Row, Edinburgh EH3 5LR, Scotland, UK.
*
E-mail for correspondence: c.ellis@rbge.org.uk
Get access

Abstract

The focus of community ecology has shifted from the description of taxonomic composition towards an understanding of community assembly based on species’ ‘functional traits’. The functional trait approach is well developed for vascular plants, utilising variability of continuous phenotypic characters that affect ecological fitness, such as specific leaf area, tissue nitrogen concentration or seed mass, to explain community structure. In contrast, community assembly studies for poikilohydric cryptogamic plants and fungi, such as lichens, remain focused on broad categorical traits such as growth form difference: fruticose, foliose or crustose. This study examined intra- and interspecific variability for two highly promising continuous phenotypic measurements that affect lichen physiology and ecological fitness: water-holding capacity (WHC) and specific thallus mass (STM). Values for WHC and STM were compared within and among species, and within and among key macrolichen growth forms (fruticose and green-algal and cyanolichen foliose species), asking whether these widely used categories adequately differentiate the continuous variables (WHC and STM). We show large intra- and interspecific variability that does not map satisfactorily onto growth form categories, and on this basis provide recommendations and caveats in the future use of lichen functional traits.

Type
Articles
Copyright
© Trustees of the Royal Botanic Garden Edinburgh (2019) 

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

Ackerly, D. D. & Cornwell, W. K. (2007). A trait-based approach to community assembly: partitioning of species trait values into within- and among-community components. Ecol. Letters 10(2): 135145.CrossRefGoogle ScholarPubMed
Alam, M.A., Gauslaa, Y. & Solhaug, K. A. (2015). Soluble carbohydrates and relative growth rates in chloro-, cyano- and cephalolichens: effects of temperature and nocturnal hydration. New Phytol. 208(3): 750762.CrossRefGoogle ScholarPubMed
Asplund, J. & Wardle, D. A. (2017). How lichens impact on terrestrial community and ecosystem properties. Biol. Rev. 92(3): 17201738.CrossRefGoogle ScholarPubMed
Bidussi, M., Gauslaa, Y. & Solhaug, K. A. (2013). Prolonging the hydration and active metabolism from light periods into nights substantially enhances lichen growth. Planta 237(5): 13591366.CrossRefGoogle ScholarPubMed
Cadotte, M. W., Mai, D. V., Jantz, S., Collins, M. D., Keele, M. & Drake, J. A. (2006). On testing the competition–colonization trade-off in a multispecies assemblage. Amer. Naturalist 168(5): 704709.CrossRefGoogle Scholar
Cornwell, W. K. & Ackerly, D. D. (2009). Community assembly and shifts in plant trait distributions across an environmental gradient in coastal California. Ecol. Monogr. 79(1): 109126.CrossRefGoogle Scholar
Ehrlén, J. & van Groenendael, J. M. (1998). The trade-off between dispersability and longevity – an important aspect of plant species diversity. Appl. Veg. Sci. 1(1): 2936.CrossRefGoogle Scholar
Elbert, W., Weber, B., Burrows, S., Steinkamp, J., Bűdel, B., Andreae, M. O. & Pöschl, U. (2012). Contribution of cryptogamic covers to the global cycles of carbon and nitrogen. Nature Geosci. 5: 459462.CrossRefGoogle Scholar
Ellis, C. J. (2013). A risk-based model of climate change threat: hazard, exposure, and vulnerability in the ecology of lichen epiphytes. Botany 91(1): 111.CrossRefGoogle Scholar
Ellis, C. J. & Coppins, B. J. (2006). Contrasting functional traits maintain lichen epiphyte diversity in response to climate and autogenic succession. J. Biogeogr. 33(9): 16431656.CrossRefGoogle Scholar
Ellis, C. J. & Ellis, S. C. (2013). Signatures of autogenic succession for an aspen chronosequence. J. Veg. Sci. 24(4): 688701.CrossRefGoogle Scholar
Ellis, C. J., Eaton, S., Theodoropoulos, M. & Elliott, K. (2015). Epiphyte Communities and Indicator Species. An Ecological Guide for Scotland’s Woodlands. Edinburgh: Royal Botanic Garden Edinburgh.Google Scholar
Esseen, P.-A., Olsson, T., Coxson, D. & Gauslaa, Y. (2015). Morphology influences water storage in hair lichens from boreal forest canopies. Fungal Ecol. 18: 2635.CrossRefGoogle Scholar
Gauslaa, Y. (2014). Rain, dew, and humid air as drivers of morphology, function and spatial distribution in epiphytic lichens. Lichenologist 46(1): 116.CrossRefGoogle Scholar
Gauslaa, Y. & Coxson, D. (2011). Interspecific and intraspecific variations in water storage in epiphytic old forest lichens. Botany 89(11): 787798.CrossRefGoogle Scholar
Gauslaa, Y., Lie, M., Solhaug, K. A. & Ohlson, M. (2006). Growth and ecophysiological acclimation of the foliose lichen Lobaria pulmonaria in forests with contrasting light climates. Oecologia 147: 406416.CrossRefGoogle ScholarPubMed
Gauslaa, Y., Palmqvist, K., Solhaug, K. A., Hilmo, O., Holien, H., Nybakken, L. & Ohlson, M. (2009). Size-dependent growth of two old-growth associated macrolichen species. New Phytol. 181(3): 683692.CrossRefGoogle ScholarPubMed
Gauslaa, Y., Solhaug, K. A. & Longinotti, S. (2017). Functional traits prolonging photosynthetically active periods in epiphytic cephalolichens during desiccation. Environm. Exp. Bot. 141: 8391.CrossRefGoogle Scholar
Grime, J. P. (1974). Vegetation classsification by reference to strategies. Nature 250, 2631.CrossRefGoogle Scholar
Grime, J. P. (1977). Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. Amer. Naturalist 111(982): 11691194.CrossRefGoogle Scholar
Grime, J. P., Thompson, K., Hunt, R., Hodgson, J. G., Cornelissen, J. H. C., Rorison, I. H., Hendry, G. A. F., Ashenden, T. W., Askew, A. P., Band, S. R., Booth, R. E., Bossard, C. C., Campbell, B. D., Cooper, J. E. L., Davison, A. W., Gupta, P. L., Hall, W., Hand, D. W., Hannah, M. A., Hillier, S. H., Hodkinson, D. J., Jalili, A., Liu, Z., Mackey, J. M. L., Matthews, N., Mowforth, M. A., Neal, A. M., Reader, R. J., Reiling, K., Ross-Fraser, W., Spencer, R. E., Sutton, F., Tasker, D. E., Thorpe, P. C. & Whitehouse, J. (1997). Integrated screening validates primary axes of specialisation in plants. Oikos 79(2): 259281.CrossRefGoogle Scholar
Hoffmann, L., Franco, A. C., Moreira, M. Z. & Haridasan, M. (2005). Specific leaf area explains differences in leaf traits between congeneric savanna and forest trees. Funct. Ecol. 19(6): 932940.CrossRefGoogle Scholar
John, E. (1992). Distribution patterns and interthalline interactions of epiphytic foliose lichens. Canad. J. Bot. 70(4): 818823.CrossRefGoogle Scholar
Jung, V., Violle, C., Mondy, C., Hoffmann, L. & Muller, S. (2010). Intraspecific variability and trait-based community assembly. J. Ecol. 98(5): 11341140.CrossRefGoogle Scholar
Kraft, N. J. B., Valencia, R. & Ackerly, D. D. (2008). Functional traits and niche-based tree community assembly in an Amazonian forest. Science 322(5901): 580582.CrossRefGoogle Scholar
Lakatos, M., Rascher, U. & Bűdel, B. (2006). Functional characteristics of corticolous lichens in the understory of a tropical lowland rain forest. New Phytol. 172: 679695.CrossRefGoogle ScholarPubMed
Lange, O. L., Kilian, E. & Ziegler, H. (1986). Water vapor uptake and photosynthesis of lichens: performance differences in species with green and blue-green algae as phycobionts. Oecologia 71(1): 104110.CrossRefGoogle ScholarPubMed
Lange, O. L., Bűdel, B., Meyer, A. & Kilian, E. (1993). Further evidence that activation of net photosynthesis by dry cyanobacterial lichens requires liquid water. Lichenologist 25(2): 175189.CrossRefGoogle Scholar
Lange, O. L., Green, T. G. A., Reichenberger, H. & Meyer, A. (1996). Photosynthetic depression at high thallus water contents in lichens: concurrent use of gas exchange and flourescence techniques with a cyanobacterial and a green algal Peltigera species. Bot. Acta 109(1): 4350.CrossRefGoogle Scholar
Larsson, P., Solhaug, K. A. & Gauslaa, Y. (2012). Seasonal partitioning of growth into biomass and area expansion in a cephalolichen and a cyanolichen of the old forest genus Lobaria . New Phytol. 194(4): 9911000.CrossRefGoogle Scholar
Lavorel, S. & Garnier, E. (2002). Predicting changes in community composition and ecosystem functioning from plant traits: revisiting the Holy Grail. Funct. Ecol. 16: 545556.CrossRefGoogle Scholar
Lewis, J. E. J. & Ellis, C. J. (2010). Taxon- compared with trait-based analysis of epiphytes, and the role of tree species and tree age in community composition. Plant Ecol. Divers. 3(2): 203210.CrossRefGoogle Scholar
Li, Y., Shipley, B., Price, J. N., Dantas, V. de L.,Tamme, R., Westoby, M., Siefert, A., Schamp, B. S., Spasojevic, M. J., Jung, V., Laughlin, D. C., Richardson, S. J., Bagousse-Pinguet, Y. L. le, Schöb, C., Gazol, A., Prentice, H. C., Gross, N., Overton, J., Cianciaruso, M. V., Louault, F., Kamiyama, C., Nakashizuka, T., Hikosaka, K., Sasaki, T., Katabuchi, M., Dussault, C. F., Gaucherand, S., Chen, N., Vandewalle, M. & Batalha, M. A. (2018). Habitat filtering determines the functional niche occupancy of plant communities worldwide. J. Ecol. 106(3): 10011009.CrossRefGoogle Scholar
Longinotti, S., Solhaug, K. A. & Gauslaa, Y. (2017). Hydration traits in cephalolichen members of the epiphytic old forest genus Lobaria (s. lat). Lichenologist 49(5): 493506.CrossRefGoogle Scholar
MacDonald, A. & Coxson, D. (2013). A comparison of Lobaria pulmonaria population structure between subalpine fir (Abies lasiocarpa) and mountain alder (Alnus incana) host-tree species in British Columbia’s inland temperate rainforest. Botany 91(8): 535544.CrossRefGoogle Scholar
Máguas, C., Griffiths, H. & Broadmeadow, M. S. J. (1995). Gas exchange and carbon isotope discrimination in lichens: evidence for interactions between CO2-concentrating mechanisms and diffusion limitation. Planta 196(1): 95102.CrossRefGoogle Scholar
Marteinsdóttir, B., Svavarsdóttir, K. & Thórhallsdóttir, T. E. (2018). Multiple mechanisms of early plant community assembly with stochasticity driving the process. Ecology 99(1): 91102.CrossRefGoogle ScholarPubMed
Matos, P., Pinho, P., Aragón, G., Martínez, I., Nunes, A., Soares, A. M. V. M. & Branquinho, C. (2015). Lichen traits responding to aridity. J. Ecol. 103(2): 451458.CrossRefGoogle Scholar
McCune, B. (1993). Gradients in epiphyte biomass in three Pseudotsuga–Tsuga forests of different ages in western Oregon and Washington. Bryologist 96(3): 405411.CrossRefGoogle Scholar
McCune, B., Amsberry, K. A., Camacho, F. J., Clery, S., Cole, C., Emerson, C., Felder, G., French, P., Greene, D., Harris, R., Hutten, M., Larson, B., Lesko, M., Majors, S., Markwell, T., Parker, G. G., Pendergrass, K., Peterson, E. B., Peterson, E. T., Platt, J., Proctor, J., Rambo, T. R., Rosso, A., Shaw, D., Turner, R. & Widmer, M. (1997). Vertical profile of epiphytes in a Pacific Northwest old-growth forest. N. W. Sci. 71(2): 145152.Google Scholar
McGill, B. J., Enquist, B. J., Weiher, E. & Westoby, M. (2006). Rebuilding community ecology from functional traits. Trends Ecol. Evol. 21(4): 178185.CrossRefGoogle ScholarPubMed
Merinero, S., Hilmo, O. & Gauslaa, Y. (2014). Size is a main driver for hydration traits in cyano- and cephalolichens of boreal rainforest canopies. Fungal Ecol. 7: 5966.CrossRefGoogle Scholar
Nascimbene, J. & Marini, L. (2015). Epiphytic lichen diversity along elevational gradients: biological traits reveal a complex response to water and energy. J. Biogeogr. 42(7): 12221232.CrossRefGoogle Scholar
Nelson, P. R., McCune, B., Roland, C. & Stehn, S. (2015). Non-parametric methods reveal non-linear functional trait variation of lichens along environmental and fire age gradients. J. Veg. Sci. 26: 848865.CrossRefGoogle Scholar
Palmqvist, K. (1993). Photosynthetic CO2-use efficiency in lichens and their isolated photobionts: the possible role of a CO2-concentrating mechanism. Planta 191(1): 4856.CrossRefGoogle Scholar
Palmqvist, K. & Sundberg, B. (2000). Light use efficiency of dry matter gain in five macro-lichens: relative impact of microclimate conditions and species-specific traits. Pl. Cell Environm. 23(1): 114.CrossRefGoogle Scholar
Phinney, N. H., Solhaug, K. A. & Gauslaa, Y. (2018). Rapid resurrection of chlorolichens in humid air: specific thallus mass drives rehydration and reactivation kinetics. Environm. Exp. Bot. 148: 184191.CrossRefGoogle Scholar
Porada, P., Weber, B., Elbert, W., Pöschl, U. & Kleidon, A. (2014). Estimating impacts of lichens and bryophytes on global biogeochemical cycles. Global Biogeochem. Cycles 28(2): 7185.CrossRefGoogle Scholar
Prieto, M., Martínez, I., Aragón, G. & Verdú, M. (2017). Phylogenetic and functional structure of lichen communities under contrasting environmental conditions. J. Veg. Sci. 28(4): 871881.CrossRefGoogle Scholar
R Development Core Team (2013). R: A Language and Environment for Statistical Computing. Vienna, Austria: R Foundation for Statistical Computing. Downloadable from https://www.R-project.org Google Scholar
Rasband, W. (no date). ImageJ: Image Processing and Analysis in Java. Online. Available: https://imagej.nih.gov/ij/ Google Scholar
Siefert, A., Violle, C., Chalmandrier, L., Albert, C. H., Taudiere, A., Fajardo, A., Aarssen, L. W., Baraloto, C., Carlucci, M. B., Cianciaruso, M. V., Dantas, V. de L., de Bello, F., Duarte, L. D. S., Fonseca, C. R., Freschet, G. T., Gaucherand, S., Gross, N., Hikosaka, K., Jackson, B., Jung, V., Kamiyama, C., Katabuchi, M., Kembel, S. W., Kichenin, E., Kraft, N. J. B., Lagerström, A., Bagousse-Pinguet, Y. L. le, Li, Y., Mason, N., Messier, J., Nakashizuka, T., Overton, J.M., Peltzer, D.A., Pérez-Ramos, I. M., Pillar, V. D., Prentice, H. C., Richardson, S., Sasaki, T., Schamp, B. S., Schöb, C., Shipley, B., Sundqvist, M., Sykes, M. T., Vandewalle, M. & Wardle, D. A. (2015). A global meta-analysis of the relative extent of intraspecific trait variation in plant communities. Ecol. Letters 18(12): 14061419.CrossRefGoogle ScholarPubMed
Smith, C. W., Aptroot, A., Coppins, B. J., Fletcher, A., Gilbert, O. L., James, P. W., Wolseley, P. A. (2009). The Lichens of Britain and Ireland. London: British Lichen Society.Google Scholar
Stanton, D. E. (2015). Small scale fog-gradients change epiphytic lichen shape and distribution. Bryologist 118(3): 241244.CrossRefGoogle Scholar
Stanton, D. E. & Horn, H. S. (2013). Epiphytes as “filter-drinkers”: life-form changes across a fog desert. Bryologist 116(1): 3442.CrossRefGoogle Scholar
Violle, C., Navas, M.-L., Vile, D., Kazakou, E., Fortunel, C., Hummel, I. & Garnier, E. (2007). Let the concept of trait be functional! Oikos 116(5): 882892.CrossRefGoogle Scholar
Weiher, E., Clarke, G. D. P. & Keddy, P. A. (1998). Community assembly rules, morphological dispersion, and the coexistence of plant species. Oikos 81(2): 309322.CrossRefGoogle Scholar