Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-05-19T11:11:39.652Z Has data issue: false hasContentIssue false

8 - Constraints on Nutrient Dynamics in Terrestrial Vegetation

from Part III - Coupling Hillslope Geomorphology, Soils, Hydrology, and Ecosystems

Published online by Cambridge University Press:  27 October 2016

David Robinson
Affiliation:
University of Aberdeen
Edward A. Johnson
Affiliation:
University of Calgary
Yvonne E. Martin
Affiliation:
University of Calgary
Get access

Summary

Introduction

Every limit is a beginning as well as an ending.

George Eliot, Middlemarch.

Nutrients, along with carbon, water and solar radiation, are required by virtually all plants to produce new biomass and, eventually, new offspring. Terrestrial plant growth is almost always limited strongly by the availability of one or more nutrients. On some soils growth is restricted by the excessive availability of potentially toxic elements. Plant biomass production and turnover have a major influence on nutrient cycling processes because many are microbial, fueled by plant-derived carbon (Bardgett, 2005; Mulder et al., 2013). This interdependence between nutrient availability and growth is manifested at large spatial and over long temporal scales as variations in ecosystem productivity, vegetation composition and soil development (Anderson-Teixeira et al., 2008; Fernández-Martínez et al., 2014; Hayes et al., 2014), but it originates at the scale of a soil pore, root cell or fungal hypha. It is at this scale that nutrients enter vegetation, but also at which nutrient capture is limited by important physical, chemical and biological factors. These small-scale constraints are the focus of this chapter.

The aim of this chapter is to explain how these constraints operate, how plants have evolved ways to subvert them, and to examine some of their consequences at the larger spatial and longer temporal scales at which ecological and biogeochemical processes are often studied. To achieve this aim, I use information that originated in mechanistic studies of nutrient dynamics in agriculture (Tinker and Nye, 2000; Gregory, 2006), but which is now recognized as a fundamental aspect of both plant ecology and terrestrial biogeochemistry (Fitter and Hay, 2002; Chapin et al., 2011; Schlesinger and Bernhardt, 2013).

Nutrients in Vegetation

Fundamentals

Macronutrients (N, P, K, S, Ca, Mg and Si) are those present in plant tissues in the largest quantities; micronutrients (Na, Cl, Mn, Fe, Cu, Zn, Mo, B and, in some species, Ni and Co) are required in much smaller abundances. Both macro- and micronutrients are essential for plant metabolism; some (Na, Si, Co and Ni) are essential only for certain species or in particular circumstances.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2016

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

Allen, M. F. and Kitajima, K. (2013). In situ high-frequency observations of mycorrhizas. New Phytologist, 200, 222–8.CrossRefGoogle Scholar
Anderson-Teixeira, K. J., Vitousek, P. M. and Brown, J. H. (2008). Amplified temperature dependence in ecosystems developing on the lava flows of Mauna Loa, Hawai'i. Proceedings of the National Academy of Sciences (USA), 105, 228–33.CrossRefGoogle Scholar
Baldwin, J. P. (1975). A quantitative analysis of the factors affecting plant nutrient uptake from some soils. Journal of Soil Science, 26, 195–206.CrossRefGoogle Scholar
Bardgett, R. D. (2005). The Biology of Soil: A Community and Ecosystem Approach. Oxford: Oxford University Press.CrossRef
Barraclough, P. B. (1989). Root growth, macro-nutrient uptake dynamics and soil fertility requirements of a high-yielding winter oilseed rape crop. Plant and Soil, 119, 59–70.CrossRefGoogle Scholar
Batjes, N. H. (1996). Total carbon and nitrogen in the soils of the world. European Journal of Soil Science, 47, 151–63.CrossRefGoogle Scholar
Bonan, G. B. (2002). Ecological Climatology: Concepts and Applications. Cambridge: Cambridge University Press.
Bouldin, D. R. (1989). A multiple ion uptake model. Journal of Soil Science, 40, 309–19.CrossRefGoogle Scholar
Broadley, M., Brown, P., Çakmak, I., Rengel, Z. and Zhao, F. (2012a). Function of nutrients: micronutrients. In Marschner's Mineral Nutrition of Higher Plants, ed. Marschner, P.. London: Elsevier, pp. 191–248.
Broadley, M., Brown, P., Çakmak, I. et al. (2012b). Beneficial elements. In Marschner's Mineral Nutrition of Higher Plants, ed. Marschner, P.. London: Elsevier, pp. 249–69.
Butnor, J. R., Barton, C., Day, F. P. et al. (2012). Using ground-penetrating radar to detect tree roots and estimate biomass. In Measuring Roots, ed. Mancuso, S.. Berlin: Springer-Verlag, pp. 213–45.CrossRef
Chapin, F. S., Matson, P. A. and Vitousek, P. M. (2011). Principles of Terrestrial Ecosystem Ecology,. New York: Springer.CrossRef
Chen, G., Yang, Y. and Robinson, D. (2014). Allometric constraints on, and trade-offs in, belowground carbon allocation and their control of soil respiration across global forest ecosystems. Global Change Biology, 20, 1674–84.CrossRefGoogle Scholar
Clark, F. E. (1977). Internal cycling of 15nitrogen in shortgrass prairie. Ecology, 58, 1322–33.CrossRefGoogle Scholar
Cooper, J. E. and Scherer, H. W. (2012). Nitrogen fixation. In Marschner's Mineral Nutrition of Higher Plants, ed. Marschner, P.. London: Elsevier, pp. 389–408.CrossRef
Dakora, F. D. and Phillips, D. A. (2002). Root exudates as mediators of mineral acquisition in low-nutrient environments. Plant and Soil, 245, 35–47.CrossRefGoogle Scholar
Dijkshoorn, W. (1969). The relation of growth to the chief ionic constituents of the plant. In Ecological Aspects of the Mineral Nutrition of Plants, ed. Rorison, I. H.. Oxford: Blackwell Scientific Publications, pp. 201–13.
Downie, H., Adu, M. O., Schmidt, S. et al. (2015). Challenges and opportunities for quantifying roots and rhizosphere interactions through imaging and image analysis. Plant, Cell and Environment, 38, 1213–32.CrossRefGoogle Scholar
Drew, M. C., Nye, P. H. and Vaidyanathan, L. (1969). The supply of nutrient ions by diffusion to plant roots in soil. I. Absorption of potassium by cylindrical roots of onion and leek. Plant and Soil, 30, 252–70.CrossRefGoogle Scholar
Eissenstat, D. M. and Yanai, R. D. (1997). The ecology of root lifespan. Advances in Ecological Research, 27, 1–60.CrossRefGoogle Scholar
Enquist, B. J. and Bentley, L. P. (2012). Land plants: new theoretical directions and empirical prospects. In Metabolic Ecology: A Scaling Approach, ed. Sibly, R. M., Brown, J. H. and Kodric-Brown, A.. Chichester: John Wiley and Sons Ltd., pp. 164–87.CrossRef
Enquist, B. J. and Niklas, K. J. (2002). Global allocation rules for patterns of biomass partitioning in seed plants. Science, 295, 1517–20.CrossRefGoogle Scholar
Fernández-Martínez, M., Vicca, S., Janssens, I. A. et al. (2014). Nutrient availability as the key regulator of global forest carbon balance. Nature Climate Change, 4, 471–6.CrossRefGoogle Scholar
Field, C. B. and Mooney, H. A. (1986). The photosynthesis-nitrogen relationship in wild plants. In On the Economy of Plant Form and Function, ed. Givnish, T. J.. Cambridge: Cambridge University Press, pp. 25–55.
Fitter, A. H. and Hay, R. K. M. (2002). Environmental Physiology of Plants,. San Diego: Academic Press.
Fraústo da Silva, J. J. R. and Williams, R. J. P. (2001). The Biological Chemistry of the Elements: The Inorganic Chemistry of Life,. Oxford: Oxford University Press.
Gardner, W. R. (1965). Movement of nitrogen in soil. In Soil Nitrogen, ed. Bartholomew, W. V. and Clark, F. E.. Madison, WI: American Society of Agronomy, pp. 550–72.
Gerber, S. and Brookshire, E. N. J. (2014). Scaling of physical constraints at the root-soil interface to macroscopic patterns of nutrient retention in ecosystems. American Naturalist, 183, 418–30.Google Scholar
Giehl, R. F. H., Gruber, B. D. and von Wirén, N. (2014). It's time to make some changes: modulation of root system architecture by nutrient signals. Journal of Experimental Botany, 65, 769–78.CrossRefGoogle Scholar
Gill, R. A. and Jackson, R. B. (2000). Global patterns of root turnover for terrestrial ecosystems. New Phytologist, 147, 13–31.CrossRefGoogle Scholar
Gordon, W. S. and Jackson, R. B. (2000). Nutrient concentrations in fine roots. Ecology, 81, 275–80.CrossRefGoogle Scholar
Gregory, P. J. (2006). Plant Roots: Growth, Activity and Interaction with Soils. Oxford: Blackwell Publishing.CrossRef
Grime, J. P., Thompson, K., Hunt, R. et al. (1997). Integrated screening validates primary axes of specialisation in plants. Oikos, 79, 259–81.CrossRefGoogle Scholar
Grinsted, M. J., Hedley, M. J., White, R. E. and Nye, P. H. (1982). Plant-induced changes in the rhizosphere of rape (Brassica napus var. Emerald) seedlings. I. pH change and the increase in P concentration in the soil solution. New Phytologist, 91, 19–29.CrossRefGoogle Scholar
Harper, J. L. (1977). Population Biology of Plants. London: Academic Press.
Hawkesford, M., Horst, W., Kichey, T. et al. (2012). Functions of macronutrients. In Marschner's Mineral Nutrition of Higher Plants, ed. Marschner, P.. London: Elsevier, pp. 135–89.CrossRef
Hayes, P., Turner, B. L., Lambers, H. and Laliberté, E. (2014). Foliar nutrient concentrations and resorption efficiency in plants of contrasting nutrient-acquisition strategies along a 2-million-year dune chronosequence. Journal of Ecology, 102, 396–410.CrossRefGoogle Scholar
Heming, S. D. and Rowell, D. L. (1985). Soil structure and potassium supply. II. The effect of ped size and root density on the supply to plants of exchangeable and non-exchangeable potassium from a chalky boulder clay soil. Journal of Soil Science, 36, 61–9.CrossRefGoogle Scholar
Hodge, A. (2004). The plastic plant: root responses to heterogeneous supplies of nutrients. New Phytologist, 162, 9–24.CrossRefGoogle Scholar
Hodge, A. and Storey, K. (2015). Arbuscular mycorrhiza and nitrogen: implications for individual plants through to ecosystems. Plant and Soil, 386, 1–19.CrossRefGoogle Scholar
Hodson, M. J., White, P. J., Mead, A. and Broadley, M. R. (2005). Phylogenetic variation in the silicon composition of plants. Annals of Botany, 96, 1027–46.CrossRefGoogle Scholar
Hoffland, E., Bloemhof, H. S., Leffelaar, P. A., Findenegg, G. R. and Nelemans, J. A. (1990). Simulation of nutrient uptake by a growing root system considering increasing root density and inter-root competition. Plant and Soil, 124,149–55.CrossRefGoogle Scholar
Huston, M. A. (2012). Precipitation, soils, NPP, and biodiversity: resurrection of Albrecht's curve. Ecological Monographs, 82, 277–96.CrossRefGoogle Scholar
Huston, M. A. and Wolverton, S. (2009). The global distribution of net primary production: resolving the paradox. Ecological Monographs, 79, 343–77.CrossRefGoogle Scholar
Jackson, R. B., Canadell, J., Ehleringer, J. R. et al. (1996). A global analysis of root distributions for terrestrial biomes. Oecologia, 108, 389–411.CrossRefGoogle Scholar
Jobbágy, E. G. and Jackson, R. B. (2001). The distribution of soil nutrients with depth: global patterns and the imprint of plants. Biogeochemistry, 53, 51–77.CrossRefGoogle Scholar
Jones, D. L., Hodge, A. and Kuzyakov, Y. (2004). Plant and mycorrhizal regulation of rhizodeposition. New Phytologist, 163, 459–80.CrossRefGoogle Scholar
Jones, D. L., Healey, J. R., Willett, V. B., Farrar, J. F. and Hodge, A. (2005). Dissolved organic nitrogen uptake by plants—an important N uptake pathway? Soil Biology and Biochemistry, 37, 413–23.Google Scholar
Kaspari, M. (2012). Stoichiometry. In Metabolic Ecology: A Scaling Approach, ed. Sibly, R. M., Brown, J. H. and Kodric-Brown, A.. Chichester: John Wiley and Sons Ltd, pp. 34–47.CrossRef
Kaspari, M., Clay, N. A., Donoso, D. A. and Yanoviak, S. P. (2014). Sodium fertilization increases termites and enhances decomposition in an Amazonian forest. Ecology, 95, 795–800.CrossRefGoogle Scholar
Kautsky, J., Barley, K. P. and Fiddaman, D. K. (1969). Ion uptake from soils by plant roots, subject to the Epstein-Hagen relation. Australian Journal of Soil Research 6, 159–67.Google Scholar
Kiers, E. T., Duhamel, M., Beesetty, Y. et al. (2012). Reciprocal rewards stabilize cooperation in the mycorrhizal symbiosis. Science, 333, 880–2.CrossRefGoogle Scholar
Knyazikhin, Y., Schull, M. A., Stenberg, P. et al. (2013). Hyperspectral remote sensing of foliar nitrogen content. Proceedings of the National Academy of Sciences (USA), 110, E185–92.CrossRefGoogle Scholar
Kutschera, L. (1960). Wurzelatlas Mitteleuropäischer Ackerunkräuter und Kulturpflanzen. Frankfurt am Main: DLG-Verlag.
Laliberté, E., Zemunik, G. and Turner, B. L. (2014). Environmental filtering explains variation in plant diversity along resource gradients. Science, 345, 1602–5.Google Scholar
Lambers, H., Chapin, F. S III and Pons, T. L. (2008). Plant Physiological Ecology,. New York: Springer Science+Business Media.
Lambers, H., Raven, J. A., Shaver, G. R. and Smith, S. E. (2008). Plant nutrient-acquisition strategies change with soil age. Trends in Ecology and Evolution, 23, 95–103.Google Scholar
Lambers, H., Finnegan, P. M., Laliberté, E. et al. (2011). Phosphorus nutrition of Proteaceae in severely phosphorus-impoverished soils: are there lessons to be learned for future crops? Plant Physiology, 156, 1058–66.Google Scholar
Lawrence, D., D'Odorico, P., Diekmann, L. et al. (2007). Ecological feedbacks following deforestation create the potential for a catastrophic ecosystem shift in tropical dry forest. Proceedings of the National Academy of Sciences (USA), 104, 20696–701.CrossRefGoogle Scholar
LeBauer, D. S. and Treseder, K. K. (2008). Nitrogen limitation of net primary productivity in terrestrial ecosystems is globally distributed. Ecology, 89, 371–9.CrossRefGoogle Scholar
Leigh, R. A. and Johnston, A. E. (1983). The effects of fertilizers and drought on the concentrations of potassium in the dry matter and tissue water of field-grown spring barley. Journal of Agricultural Science, 101, 741–8.CrossRefGoogle Scholar
Leigh, R. A. and Wyn Jones, R. G. (1984). A hypothesis relating critical potassium concentrations form growth to the distribution and functions of this ion in the plant cell. New Phytologist, 91, 1–13.CrossRefGoogle Scholar
Li, X.-L., George, E. and Marschner, H. (1991). Phosphorus depletion and pH decrease at the root-soil and hyphae-soil interfaces of VA mycorrhizal white clover fertilized with ammonium. New Phytologist, 119, 397–404.CrossRefGoogle Scholar
Mairhofer, S., Zappala, S., Tracy, S. R. et al. (2012). RooTrak: automated recovery of three-dimensional plant root architecture in soil from x-ray microcomputed tomography images using visual tracking. Plant Physiology, 158, 561–9.Google Scholar
McCully, M. E. (1999). Roots in soil: unearthing the complexities of roots and their rhizospheres. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 695–718.CrossRefGoogle Scholar
Merryweather, J. and Fitter, A. (1995). Phosphorus and carbon budgets: mycorrhizal contribution in Hyacinthoides non-scripta (L.) Chouard ex Rothm. under natural conditions. New Phytologist, 129, 619–27.CrossRefGoogle Scholar
Michaletz, S. T., Cheng, D., Kerkhoff, A. J. and Enquist, B. J. (2014). Convergence of terrestrial plant production across global climate gradients. Nature, 512, 39–43.CrossRefGoogle Scholar
Millard, P. and Grelet, G.-A. (2010). Nitrogen storage and remobilization by trees: ecophysiological relevance in a changing world. Tree Physiology, 30, 1083–95.CrossRefGoogle Scholar
Mulder, C., Ahrestani, F. S., Bahn, M. et al. (2013). Connecting the green and brown worlds: allometric and stoichiometric predictability of above- and below-ground networks. Advances in Ecological Research, 49, 69–175.CrossRefGoogle Scholar
Neumann, G. and Römheld, V. (2012). Rhizosphere chemistry in relation to plant nutrition. In Marschner's Mineral Nutrition of Higher Plants, ed. Marschner, P.. London: Elsevier, pp. 347–68.CrossRef
Newman, E. I. (1988). Mycorrhizal links between plants: their functioning and ecological significance. Advances in Ecological Research, 18, 243–70.CrossRefGoogle Scholar
Newman, E. I., Ritz, K. and Jupp, A. P. (1989). The functioning of roots in grassland communities. Aspects of Applied Biology, 22, 263–9.Google Scholar
Northrup, R. R., Dahlgren, R. A. and McColl, J. G. (1998). Polyphenols as regulators of plant-litter-soil interactions in northern California's pygmy forest: a positive feedback? Biogeochemistry, 42,189–220.Google Scholar
Nye, P. H. (1994). The effect of root shrinkage on soil water inflow. Philosophical Transactions of the Royal Society of London, B345, 395–402.CrossRefGoogle Scholar
Ollinger, S. V., Richardson, A. D., Martin, M. E. et al. (2008). Canopy nitrogen, carbon assimilation, and albedo in temperate and boreal forests: functional relations and potential climate feedbacks. Proceedings of the National Academy of Sciences (USA), 105, 19336–41.Google Scholar
Paungfoo-Lonhienne, C., Lonhienne, T. G. A., Rentsch, D. et al. (2008). Plants can use protein as a nitrogen source without assistance from other organisms. Proceedings of the National Academy of Sciences (USA), 105, 4524–9.CrossRefGoogle Scholar
Peek, M. S. (2007). Explaining variation in fine root life span. Progress in Botany, 68, 282–398.CrossRefGoogle Scholar
Poorter, H., Bühler, J., van Dusschoten, D., Climent, J. and Postma, J. A. (2012a). Pot size matters: a meta-analysis of the effects of rooting volume on plant growth. Functional Plant Biology, 39, 839–50.Google Scholar
Poorter, H., Niklas, K. J., Reich, P. B. et al. (2012b). Biomass allocation to leaves, stems and roots: meta-analyses of interspecific variation and environmental control. New Phytologist, 193, 30–50.Google Scholar
Pregitzer, K. S. and King, J. S. (2005). Effects of temperature on nutrient uptake. In Nutrient Acquisition by Plants: An Ecological Perspective, ed. BassiriRad, H.. Heidelberg: Springer-Verlag, pp. 277–310.CrossRef
Read, D. B., Bengough, A. G., Gregory, P. J. et al. (2003). Plant roots release phospholipid surfactants that modify the physical and chemical properties of soil. New Phytologist, 157, 315–26.CrossRefGoogle Scholar
Read, D. J. and Perez-Moreno, J. (2003). Mycorrhizas and nutrient cycling in ecosystems—a journey towards relevance? New Phytologist, 157, 475–92.Google Scholar
Ritz, K. (1995). Growth responses of some fungi to spatially heterogeneous nutrients. FEMS Microbiology Ecology, 26, 269–80.CrossRefGoogle Scholar
Robinson, D. (1994a). Resource capture by single roots. In Resource Capture by Crops, ed. Monteith, J. L., Scott, R. K. and Unsworth, M. H.. Nottingham: Nottingham University Press, pp. 53–76.
Robinson, D. (1994b). The responses of plants to non-uniform supplies of nutrients. New Phytologist, 127, 635–74.Google Scholar
Robinson, D. (1996). Variation, co-ordination and compensation in root systems in relation to soil variability. Plant and Soil, 187, 57–66.CrossRefGoogle Scholar
Robinson, D. (2001). Root proliferation, nitrate inflow and their carbon costs during nitrogen capture by competing plants in patchy soil. Plant and Soil, 232, 41–50.CrossRefGoogle Scholar
Robinson, D. (2004). Scaling the depths: below-ground allocation in plants, forests and biomes. Functional Ecology, 18, 290–5.Google Scholar
Robinson, D. and Fitter, A. (1999). The magnitude and control of carbon transfer between plants linked by a common mycorrhizal network. Journal of Experimental Botany, 50, 9–13.Google Scholar
Robinson, D. and Rorison, I. H. (1987). Root hairs and plant growth at low nitrogen availabilities. New Phytologist, 107, 681–93.CrossRefGoogle Scholar
Robinson, D., Griffiths, B., Ritz, K. and Wheatley, R. (1989). Root-induced nitrogen mineralisation: a theoretical analysis. Plant and Soil, 117, 185–93.CrossRefGoogle Scholar
Robinson, D., Hodge, A., Griffiths, B. S. and Fitter, A. H. (1999). Plant root proliferation in nitrogen-rich patches confers competitive advantage. Proceedings of the Royal Society of London, B266, 431–5.CrossRefGoogle Scholar
Robinson, D., Linehan, D. J. and Caul, S. (1991). What limits nitrate uptake from soil? Plant, Cell and Environment, 14, 77–85.Google Scholar
Ryan, M. H., Tibbett, M., Edmonds-Tibbett, T. et al. (2012). Carbon trading for phosphorus gain: the balance between rhizosphere carboxylates and arbuscular mycorrhizal symbiosis in plant phosphorus acquisition. Plant, Cell and Environment, 35, 2170–80.Google Scholar
Sanders, F. E. and Tinker, P. B. (1973). Phosphate flow into mycorrhizal roots. Pesticide Science, 4, 385–95.CrossRefGoogle Scholar
Schlesinger, W. H. and Bernhardt, E. S. (2013). Biogeochemistry,. Waltham, MA: Elsevier.
Shane, M. W. and Lambers, H. (2005). Cluster roots; a curiosity in context. Plant and Soil, 274, 101–25.Google Scholar
Shane, M. W., Cramer, M. D., Funayama-Noguchi, S. et al. (2004). Developmental physiology of cluster-root carboxylate synthesis and exudation in harsh hakea. Expression of phosphoenolpyruvate carboxylase and the alternative oxidase. Plant Physiology, 135, 549–60.CrossRefGoogle Scholar
Simard, S. W. and Durall, D. M. (2004). Mycorrhizal networks: a review of their extent, function, and importance. Canadian Journal of Botany, 82, 1140–65.Google Scholar
Simard, S. W., Jones, M. and Durall, D. M. (2002). Carbon and nutrient fluxes within and between mycorrhizal plants. In Mycorrhizal Ecology, ed. van der Heijden, M.G.A. and Sanders, I.. Heidelberg: Springer-Verlag, pp. 33–74.
Smith, S. E. and Read, D. J. (2008). Mycorrhizal Symbioses,. London: Academic Press.
Smith, S. E., Jakobsen, I., Grønlund, M. and Smith, F. A. (2011). Roles of arbuscular mycorrhizas in plant phosphorus nutrition: interactions between pathways of phosphorus uptake in arbuscular mycorrhizal roots have important implications for understanding and manipulating plant phosphorus acquisition. Plant Physiology, 156, 1050–7.Google Scholar
Staddon, P. L., Ramsey, C. B., Ostle, N., Ineson, P. and Fitter, A. H. (2003). Rapid turnover of hyphae of mycorrhizal fungi determined by AMS microanalysis of 14C. Science, 300, 1138–40.CrossRefGoogle Scholar
Sterner, R. W. and Elser, J. J. (2002). Ecological Stoichiometry: The Biology of Elements from Molecules to the Biosphere. Princeton, NJ: Princeton University Press.
Stevens, G. N., Jones, R. H. and Mitchell, R. J. (2002). Rapid fine root disappearance in a pine woodland: a substantial carbon flux. Canadian Journal of Forest Research, 32, 2225–30.CrossRefGoogle Scholar
Sun, J., Bankston, J. R., Payandeh, J. et al. (2014). Crystal structure of the plant dual-affinity nitrate transporter NRT1.1. Nature, 507, 73–7.Google Scholar
Swift, M. J., Heal, O. W. and Anderson, J. M. (1979). Decomposition in Terrestrial Ecosystems. Oxford: Blackwell Scientific Publications.
Teste, F. P. Veneklaas, E. J., Dixon, K. W. and Lambers, H. (2014). Complementary plant nutrient-acquisition strategies promote growth of neighbour species. Functional Ecology, 28, 819–28.CrossRefGoogle Scholar
Thompson, K., Parkinson, J. A., Band, S. R. and Spencer, R. E. (1997). A comparative study of leaf nutrient concentrations in a regional herbaceous flora. New Phytologist, 136, 679–89.Google Scholar
Tinker, P. B. (1976). Transport of water to plant roots in soil. Philosophical Transactions of the Royal Society of London, B273, 445–61.CrossRefGoogle Scholar
Tinker, P. B. and Nye, P. H. (2000). Solute Movement in the Rhizosphere. New York: Oxford University Press.
Trinder, C. J., Brooker, R., Davidson, H. and Robinson, D. (2013). Dynamic trajectories of growth and nitrogen capture by competing plants. New Phytologist, 193, 948–58.CrossRefGoogle Scholar
Van Vuuren, M. M. I., Robinson, D. and Griffiths, B. S. (1996). Nutrient inflow and root proliferation during the exploitation of a temporally and spatially discrete source of nitrogen in soil. Plant and Soil, 178, 185–92.CrossRefGoogle Scholar
Vergutz, L., Manzoni, S., Porporato, A., Novais, R. F. and Jackson, R. B. (2012a). A global database of carbon and nutrient concentrations of green and senesced leaves. Data set available online [http://daac.ornl.gov] from Oak Ridge National Laboratory Distributed Active Archive Center, Oak Ridge, Tennessee, U.S.A. http://dx.doi.org/10.3334/ORNLDAAC/1106.
Vergutz, L., Manzoni, S., Porporato, A., Novais, R. F. and Jackson, R. B. (2012b). Global resorption efficiencies and concentrations of carbon and nutrients in leaves of terrestrial plants. Ecological Monographs, 82, 205–20.Google Scholar
Walder, F., Niemann, H., Natarajan, M. et al. (2012). Mycorrhizal networks: common goods of plants shared under unequal terms of trade. Plant Physiology, 159, 789–97.CrossRefGoogle Scholar
Wang, Y.-H., Garvin, D. F. and Kochian, L. V. (2002). Rapid induction of regulatory and transporter genes in response to phosphorus, potassium, and iron deficiencies in tomato roots. Evidence for cross talk and root/rhizosphere-mediated signals. Plant Physiology, 130, 1361–70.CrossRefGoogle Scholar
Watanabe, T., Broadley, M. R., Jansen, S. et al. (2007). Evolutionary control of leaf element composition in plants. New Phytologist, 174, 516–23.Google Scholar
Watson, C.A., Ross, J. M., Bagnaresi, U. et al. (2000). Environment-induced modifications to root longevity in Lolium perenne and Trifolium repens . Annals of Botany, 85, 397–401.CrossRefGoogle Scholar
Weremijewicz, J. and Janos, D. P. (2013). Common mycorrhizal networks amplify size inequality in Andropogon gerardii monocultures. New Phytologist, 198, 203–13.Google Scholar
West, G., Brown, J. H. and Enquist, B. J. (1999). A general model for the structure and allometry of plant vascular systems. Nature, 400, 664–7.CrossRefGoogle Scholar
White, P. J. (2012). Ion uptake mechanisms of individual cells and roots: short-distance transport. In Marschner's Mineral Nutrition of Higher Plants, ed. Marschner, P., pp. 7–47. London: Elsevier.CrossRef
White, P. J., Bowen, H. C., Marshall, B. and Broadley, M. R. (2007). Extraordinarily high leaf selenium to sulfur ratios define “Se-accumulator” plants. Annals of Botany, 100, 111–8.CrossRefGoogle Scholar
Wright, I. J., Reich, P. B., Westoby, M. et al. (2004). The worldwide leaf economics spectrum. Nature, 428, 821–7.Google Scholar
Yanai, J., Linehan, D. J., Robinson, D. et al. (1996). Effects of inorganic nitrogen application on the dynamics of the soil solution composition in the root zone of maize. Plant and Soil, 180, 1–9.Google Scholar
Yanai, R. D. (1994). A steady-state model of nutrient uptake accounting for newly grown roots. Soil Science Society of America Journal, 58, 1562–71.CrossRefGoogle 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
×