Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-29T06:11:06.697Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 March 2013

Jaume Flexas
Affiliation:
Universitat de les Illes Balears, Palma de Mallorca
Francesco Loreto
Affiliation:
Consiglio Nazionale delle Ricerche (CNR), Firenze
Hipólito Medrano
Affiliation:
Universitat de les Illes Balears, Palma de Mallorca
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
Terrestrial Photosynthesis in a Changing Environment
A Molecular, Physiological, and Ecological Approach
, pp. 546 - 722
Publisher: Cambridge University Press
Print publication year: 2012

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

Aalto, T. and Juurola, E. (2002). A three-dimensional model of CO2 transport in airspaces and mesophyll cells of a silver birch leaf. Plant, Cell and Environment, 25, 1399–1409.CrossRefGoogle Scholar
Aasamaa, K., Sober, A. and Rahi, M. (2001). Leaf anatomical characteristics associated with hydraulic conductance, stomatal conductance and stomatal sensitivity to changes of leaf water status in temperate deciduous trees. Australian Journal of Plant Physiology, 28, 765–774.Google Scholar
Aasamaa, K., Sõber, A., Hartung, W., et al. (2002). Rate of stomatal opening, shoot hydraulic conductance and photosynthesis characteristics in relation to leaf abscisic acid concentration in six temperate deciduous trees. Tree Physiology, 22, 267–276.CrossRefGoogle ScholarPubMed
Aasamaa, K., Sõber, A., Hartung, W., et al. (2004). Drought acclimation of two deciduous tree species of different layers in a temperate forest canopy. Trees: Structure and Function, 18, 93–101.Google Scholar
Abadía, J., Morales, F. and Abadía, A. (1999). Photosystem II efficiency in low chlorophyll, iron-deficient leaves. Plant and Soil, 215, 183–192.CrossRefGoogle Scholar
Abadía, J., Nishio, J.N. and Terry, N. (1986). Chlorophyll-protein and polypeptide composition of Mn-deficient sugar beet thylakoids. Photosynthesis Research, 7, 379–381.CrossRefGoogle ScholarPubMed
Abbink, T.E.M., Peart, J.R., Mos, T.N.M., et al. (2002). Silencing of a gene encoding a protein component of the oxygen-evolving complex of Photosystem II enhances virus replication in plants. Virology, 295, 307–319.CrossRefGoogle ScholarPubMed
Abe, H., Urao, T., Ito, T., et al. (2003). Arabidopsis AtMYC2 (bHLH) and AtMYB2 (MYB) function as transcriptional activators in abscisic acid signaling. Plant Cell, 15, 63–78.CrossRefGoogle ScholarPubMed
Abeledo, J.G., Calderini, D.F. and Slafer, G.A. (2003). Genetic improvement of barley yield potential and its physiological determinants in Argentina (1944–1998). Euphytica, 130, 325–334.CrossRefGoogle Scholar
Abeles, F.B. (1986). Plant Chemiluminescence. Annual Review of Plant Physiology, 37, 49–72.CrossRefGoogle Scholar
Aber, J.D., Reich, P.B. and Goulden, M.L. (1996). Extrapolating leaf CO2 exchange to the canopy: a generalized model of forest photosynthesis compared with measurements by eddy correlation. Oecologia, 106, 257–265.CrossRefGoogle ScholarPubMed
Abrams, M.D. (1996). Distribution, historical development and ecophysiological attributes of oak species in the eastern United States. Annales des Sciences Forestieres, 53, 487–512.CrossRefGoogle Scholar
Ache, P., Bauer, H., Kollist, H., et al. (2010). Stomatal action directly feeds back on leaf turgor: new insights into the regulation of the plant water status from non-invasive pressure probe measurements. The Plant Journal, 62, 1072–1082.Google ScholarPubMed
Ackerly, D. (1997). Allocation, leaf display, and growth in fluctuating light environments. In: Plant Resource Allocation (eds Bazzaz, F. A. and Grace, J.). Academic Press, San Diego.
Adams, K.L. and Wendel, J.F. (2005). Polyploidy and genome evolution in plants. Current Opinion in Plant Biology, 8, 135–141.CrossRefGoogle ScholarPubMed
Adams, M.L., Norvell, W.A., Peverly, J.H., et al. (1993). Fluorescence and reflectance characteristics of manganese deficient soybean leaves: effects of leaf age and choice of leaflet. Plant and Soil, 155/156, 235–238.CrossRefGoogle Scholar
Adams, M.S. and Strain, B.R. (1968). Photosynthesis in stems and leaves of Cercidium floridum: spring and summer diurnal field response and relation to temperature. Oecologia Plantarum, 3, 285–297.Google Scholar
Adams, P., Nelson, D.E., Yamada, S., et al. (1998). Growth and development of Mesembryanthemum crystallinum (Aizoaceae). New Phytologist, 138, 171–190.CrossRefGoogle Scholar
Adams, W.W. III and Demmig-Adams, B. (1994). Carotenoid composition and down regulation of photosystem II in three conifer species during the winter. Physiologia Plantarum, 92, 451–458.CrossRefGoogle Scholar
Adams, W.W. III, Demmig-Adams, B., Rosenstiel, T.N., et al. (2002). Photosynthesis and photoprotection in overwintering plants. Plant Biology, 4, 535–654.Google Scholar
Adams, W.W. III, Demmig-Adams, B., Verhoeven, A.S., et al. (1995). ‘Photoinhibition’ during winter stress: involvement of sustained xanthophyll cycle-dependent energy dissipation. Australian Journal of Plant Physiology, 22, 261–276.CrossRefGoogle Scholar
Adams, W.W. III, Demmig-Adams, B., Winter, K., et al. (1990a). The ratio of variable to maximum chlorophyll fluorescence from photosystem II, measured at room temperature and 77K, as indicator of the photon yield of photosynthesis. Planta, 180, 166–174.CrossRefGoogle Scholar
Adams, W.W. III, Winter, K. and Lanzl, A. (1989). Light and the maintenance of photosynthetic competence in leaves of Populus balsamifera L. during short-term exposures to high concentrations of sulfur dioxide. Planta, 177, 91–97.CrossRefGoogle ScholarPubMed
Adams, W.W. III, Winter, K., Schreiber, U., et al. (1990b). Photosynthesis and chlorophyll fluorescence characteristics in relationship to changes in pigment and element composition of leaves of Platanus occidentalis L. during autumnal leaf senescence. Plant Physiology, 93, 1184–1190.CrossRefGoogle Scholar
Affek, H.P., Krisch, M.J. and Yakir, D. (2006). Effects of intraleaf variations in carbonic anhydrase activity and gas echange on leaf C18OO isoflux in Zea mays. New Phytologist, 169, 321–329.CrossRefGoogle Scholar
Aganchich, B., Wahbi, S., Loreto, F., et al. (2009). Partial root zone drying: regulation of photosynthetic limitations and antioxidant enzymatic activities in young olive (Olea europaea) saplings. Tree Physiology, 29, 685–696.CrossRefGoogle ScholarPubMed
Agati, G., Cerovic, Z.G. and Moya, I. (2000). The effect of decreasing temperature up to chilling values on the in vivo F685/F735 chlorophyll fluorescence ratio in Phaseolus vulgaris and Pisum sativum. The role of the photosystem I contribution to the 735 nm fluorescence band. Photochemistry and Photobiology, 72, 75–84.2.0.CO;2>CrossRefGoogle ScholarPubMed
Agati, G., Cerovic, Z.G., Dalla Marta, A., et al. (2008). Optically assessed preformed flavonoids and susceptibility of grapevine to Plasmopora viticola under different light regimes. Functional Plant Biology, 35, 77–84.CrossRefGoogle Scholar
Agati, G., Galardi, C., Gravano, E., et al. (2002). Flavonoid distribution in tissues of Phillyrea latifolia L. leaves as estimated by microspectrofluorometry and multispectral fluorescence microimaging. Photochemistry and Photobiology, 76, 350–360.2.0.CO;2>CrossRefGoogle ScholarPubMed
Agati, G., Mazzinghi, P., Fusi, F., et al. (1995). The F685/730 chlorophyll fluorescence ratio as a tool in plant physiology: response to physiological and environmental factors. Journal of Plant Physiology, 145, 228–238.CrossRefGoogle Scholar
Agati, G., Meyer, S., Matteini, P., et al. (2007). Assessment of anthocyanins in grape (Vitis vinifera L.) berries using a noninvasive chlorophyll fluorescence method. Journal of Agricultural and Food Chemistry, 55, 1053–1061.CrossRefGoogle ScholarPubMed
Agbariah, K.T. and Roth-Bejerano, N. (1990). The effect of blue light on energy levels in epidermal strips. Physiologia Plantarum, 78, 100–104.CrossRefGoogle Scholar
Ågren, G.I. and Ingestad, T. (1987). Root:shoot ratio as a balance between nitrogen productivity and photosynthesis. Plant Cell Environment, 10, 579–586.Google Scholar
Aguirreolea, J., Irigoyen, J., Sánchez-Díaz, M., et al. (1995). Physiological alterations in pepper during wilt induced by Phytophthora capsici and soil water deficit. Plant Pathology, 44, 587–596.CrossRefGoogle Scholar
Agustí, S., Enríquez, S., Frost-Christensen, H., et al. (1994). Light harvesting among photosynthetic organisms. Functional Ecology, 8, 273–279.CrossRefGoogle Scholar
Agustynowicz, J. and Gabrys, H. (1999). Chloroplast movements in fern leaves: correlation of movement dynamics and environmental flexibility of the species. Plant, Cell and Environment, 22, 1239–1248.CrossRefGoogle Scholar
Aharon, R., Shahak, Y., Wininger, S., et al. (2003). Overexpression of a plasma membrane aquaporin in transgenic tobacco improves plant vigor under favourable growth conditions but not under drought or salt stress. The Plant Cell, 15, 439–447.CrossRefGoogle ScholarPubMed
Ahn, T.K., Avenson, T.J., Ballottari, M., et al. (2008). Architecture of a charge-transfer state regulating light harvesting in a plant antenna protein. Science, 320, 794–797.CrossRefGoogle Scholar
Ainsworth, E., Davey, P., Bernacchi, C., et al. (2002). A meta-analysis of elevated [CO2] effects on soybean (Glycine max) physiology, growth and yield. Global Change Biology, 8, 695–709.CrossRefGoogle Scholar
Ainsworth, E., Leakey, A., Ort, D., et al. (2008b). FACE-ing the facts: inconsistencies and interdependence among field, chamber and modeling studies of elevated [CO2] impacts on crop yield and food supply. New Phytologist, 179, 5–9.CrossRefGoogle ScholarPubMed
Ainsworth, E.A. (2008). Rice production in a changing climate: a meta-analysis of responses to elevated carbon dioxide and elevated ozone concentration. Global Change Biology, 14, 1642–1650.CrossRefGoogle Scholar
Ainsworth, E.A. and Long, S.P. (2005). What have we learned from 15 years of free-air CO2 enrichment (FACE)? A meta-analytic review of the responses of photosynthesis, canopy properties and plant production to rising CO2. New Phytologist, 165, 351–372.CrossRefGoogle ScholarPubMed
Ainsworth, E.A. and Rogers, A. (2007). The response of photosynthesis and stomatal conductance to rising [CO2]: mechanisms and environmental interactions. Plant, Cell and Environment, 30, 258–270.CrossRefGoogle Scholar
Ainsworth, E.A., Davey, P.A., Hymus, G.J., et al. (2003). Is stimulation of leaf photosynthesis by elevated carbon dioxide concentration maintained in the long term? A test with Lolium perenne grown for 10 years at two nitrogen fertilization levels under free air CO2 enrichment (FACE). Plant, Cell and Environment, 26, 705–714.CrossRefGoogle Scholar
Ainsworth, E.A., Rogers, A. and Leakey, A.D.B. (2008a). Targets for crop biotechnology in a future high CO2 and HighO3 world. Plant Physiology, 147, 13–19.CrossRefGoogle Scholar
Ainsworth, E.A., Rogers, A.Nelson, R., et al. (2004). Testing the “source-sink” hypothesis of down-regulation of photosynthesis in elevated CO2 in the field with single gene substitutions in Glycine max. Agricultural and Forest meteorology, 122, 85–94.CrossRefGoogle Scholar
Akhani, H., Barroca, J., Koteeva, N., et al. (2005). Bienertia sinuspersici (Chenopodiaceae): A new species from southwest Asia and discovery of a third terrestrial C4 plant without Kranz anatomy. Systematic Botany, 30, 290–301.CrossRefGoogle Scholar
Akhani, H., Trimborn, P. and Ziegler, H. (1997). Photosynthetic pathways in Chenopodiaceae from Africa, Asia and Europe with their ecological, phytogeographical and taxonomical importance. Plant Systematics and Evolution, 206, 187–221.CrossRefGoogle Scholar
Al-Abbas, A.H., Barr, R., Hall, J.D., et al. (1974). Spectra of normal and nutrient deficient maize leaves. Agronomy Journal, 66, 16–20.CrossRefGoogle Scholar
Alboresi, A., Gerotto, C., Giacometti, G.M., et al. (2010). Physcomitrella patens mutants affected on heat dissipation clarify the evolution of photoprotection mechanisms upon land colonization. Proceedings of the National Academy of Sciences USA, 107, 11128–11133.CrossRefGoogle ScholarPubMed
Aldea, M., Frank, T.D. and DeLucia, E.H. (2006a). A method for quantitative analysis of spatially variable physiological processes across leaf surfaces. Photosynthesis Research, 90, 161–172.CrossRefGoogle ScholarPubMed
Aldea, M., Hamilton, J.G., Resti, J.P., et al. (2005). Indirect effects of insect herbivory on leaf gas exchange in soybean. Plant, Cell and Environment, 28, 402–411.CrossRefGoogle Scholar
Aldea, M., Hamilton, J.G., Resti, J.P., et al. (2006b). Comparison of photosynthetic damage from arthropod herbivory and pathogen infection in understory hardwood saplings. Oecologia, 149, 221–232.CrossRefGoogle ScholarPubMed
Al-Hazmi, M., Lakso, A.N. and Denning, S.S. (1997). Whole canopy versus single leaf gas exchange responses to water stress in Cabernet Sauvignon grapevines. In: Proceedings of the IVth International Symposium on Cool Climate Viticulture and Enology (eds Edson, C., Wolf, T., Pool, R., Reynolds, A., Henick-Kling, T., Acree, T., Reisch, B. and Harkness, E.), Eastern Section, Am. Soc. Enol. Vitic., Geneva, NY, USA, pp. II.47-II.48.Google Scholar
Ali, B., Hasan, S.A., Hayat, S., et al. (2008). A role for brassinosteroids in the amelioration of aluminium stress through antioxidant system in mung bean (Vigna radiata L. Wilczek). Environmental and Experimental Botany, 62, 153–159.CrossRefGoogle Scholar
Allard, V., Ourcival, J.M., Rambal, S., et al. (2008). Seasonal and annual variation of carbon exchange in an evergreen Mediterranean forest in southern France. Global Change Biology, 14, 1–12.CrossRefGoogle Scholar
Allen, D.J. and Ort, D.R. (2001). Impacts of chilling temperatures on photosynthesis in warm-climate plants. Trends Plant Sciences, 6, 36–42.CrossRefGoogle ScholarPubMed
Allen, D.J., Ratner, K., Giller, Y.E., et al. (2000). An overnight chill induces a delayed inhibition of photosynthesis at midday in mango (Mangifera indica L.). Journal of Experimental Botany, 51, 1893–1902.CrossRefGoogle Scholar
Allen, D.K., Shachar-Hill, Y. and Ohlrogge, J.B. (2007). Compartment-specific labeling information in 13C metabolic flux analysis of plants. Phytochemistry, 68, 2197–2210.CrossRefGoogle Scholar
Allen, J.F. (1992). How does protein phosphorylation regulate photosynthesis?Trends in Biochemical Sciences, 17, 12–17.CrossRefGoogle ScholarPubMed
Allen, J.F. (2003). Cyclic, pseudocyclic and noncyclic photophosphorylation: new links in the chain. Trends Plant Sciences, 8, 15–19.CrossRefGoogle ScholarPubMed
Allen, J.F. and Forsberg, J. (2001). Molecular recognition in thylakoid structure and function. Trends Plant Sciences, 6, 317–326.CrossRefGoogle ScholarPubMed
Allen, J.F. and Nilsson, A. (1997). Redox signalling and the structural basis of regulation of photosynthesis by protein phosphorylation. Physiologia Plantarum, 100, 863–868.CrossRefGoogle Scholar
Allen, L.H., Pan, D., Boote, K.J., et al. (2003). Carbon dioxide and temperature effects on evapotranspiration and water use efficiency of soybean. Agronomy Journal, 95, 1071–1081.CrossRefGoogle Scholar
Allen, M.F. (1991). The Ecology of Mycorrhizae. Cambridge University Press, Cambridge, UK.Google Scholar
Almási, A., Harsányi, A. and Gáborjányi, R. (2001). Photosynthetic alterations on virus infected plants. Acta Phytopathologica et Entomologica Hungarica, 36, 15–29.CrossRefGoogle Scholar
Alo, C.A. and Wang, G.L. (2008). Potential future changes of the terrestrial ecosystem based on climate projections by eight general circulation models. Journal of Geophysical Research – Biogeosciences, 113, Article number G01004.CrossRefGoogle Scholar
Alongi, D.M., Ayukai, T., Brunskill, G.J., et al. (1998). Sources, sinks and export of organic carbon through a tropical, semi-enclosed delta (Hinchinbrook Channel, Australia). Mangrove and Salt Marshes, 2, 237–242.CrossRefGoogle Scholar
Alonso, A.A. and Machado, S.R. (2007). Morphological and developmental investigations of the underground system of Erythroxylum species from Brazilian cerrado. Australian Journal of Botany, 55, 749–758.CrossRefGoogle Scholar
Alterio, G., Giorio, P. and Sorrentino, G. (2006). Open-system chamber for measurements of gas exchanges at plant level. Environmental Science and Technology, 40, 1950–1955.CrossRefGoogle ScholarPubMed
Altesor, A., Ezcurra, E. and Silva, C. (1992). Changes in the photosynthetic metabolism during early ontogeny of four cactus species. Acta Oecologica, 13, 777–785.Google Scholar
Althawadi, A.M. and Grace, J. (1986). Water use by the desert cucurbit Citrullus colocynthis (L.) Schrad. Oecologia, 70, 475–480.CrossRefGoogle ScholarPubMed
Amthor, J.S. (1991). Respiration in a future, higher-CO2 world. Plant Cell and Environment, 14, 13–20.CrossRefGoogle Scholar
Amthor, J.S. (1994). Scaling CO2-photosynthesis relationships from the leaf to the canopy. Photosynthesis Research, 39, 321–350.CrossRefGoogle ScholarPubMed
Amthor, J.S. (2000). The McCree–de Wit–Penning de Vries–Thornley respiration paradigms: 30 years later. Annals of Botany, 86, 1–20.CrossRefGoogle Scholar
Amthor, J.S. (2007). Improving photosynthesis and yield potential. In: Improvement of Crop Plants for Industrial End Uses, (ed. Ramalli, P.), Springer, NY, USA, pp. 27–58.Google Scholar
Amthor, J.S., Goulden, M.L., Munger, J.W., et al. (1994). Testing a mechanistic model of forest-canopy mass and energy exchange using eddy correlation: carbon dioxide and ozone uptake by a mixed oak-maple stand. Australian Journal of Plant Physiology, 21, 623–651.CrossRefGoogle Scholar
Ananyev, G.S., Kolber, Z.S., Klimov, D., et al. (2005). Remote sensing of heterogeneity in photosynthetic efficiency, electron transport and dissipation of excess light in Populus deltoides stands under ambient and elevated CO2 concentrations, and in a tropical forest canopy, using a new laser-induced fluorescence transient device. Global Change Biology, 11, 1195–1206.CrossRefGoogle Scholar
Anderson, D.E. and Verma, S.B. (1986). Carbon dioxide, water vapor and sensible heat exchanges of a grain sorghum canopy. Boundary Layer Meteorology, 34, 317–331.CrossRefGoogle Scholar
Anderson, J.M., Chow, W.S., and Goodchild, D.J. (1988). Thylakoid membrane organization in sun shade acclimation. Australian Journal of Plant Physiology, 15, 11–26.CrossRefGoogle Scholar
Anderson, L.J., Maherali, H., Johnson, H.B., et al. (2001). Gas exchange and photosynthetic acclimation over subambient to elevated CO2 in a C3-C4 grassland. Global Change Biology, 7, 693–707.CrossRefGoogle Scholar
Andersson, A., Keskitalo, J., Sjödin, A., et al. (2004). A transcriptional timetable of autumn senescence. Genome Biology, 5, R24.CrossRefGoogle ScholarPubMed
Andersson, B. and Barber, J. (1996). Mechanisms of photodamage and protein degradation during photoinhibition of photosystem II. In: Advances in Photosynthesis Vol.5: Photosynthesis and the Environment (ed. Baker, N.R.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 101–121.Google Scholar
Andrade, J.L. and Nobel, P.S. (1996). Habitat, CO2 uptake and growth for the CAM epiphytic cactus Epiphyllum phyllanthus in a Panamanian tropical forest. Journal of Tropical Ecology, 12, 291–306.CrossRefGoogle Scholar
Andralojc, P.J., Keys, A.J., Kossmann, J., et al. (2002). Elucidating the biosynthesis of 2-carboxyarabinitol 1-phosphate through reduced expression of chloroplastic fructose 1,6-bisphosphate phosphatase and radiotracer studies with 14CO2. Proceedings of National Academy of Sciences of USA, 99, 4742–4747.CrossRefGoogle ScholarPubMed
Andralojc, P.J., Keys, A.J., Martindale, W., et al. (1996). Conversion of Δ-hamamelose into 2-carboxy D-arabinitol and 2-carboxy D-arabinitol 1-phosphate in leaves of Phaseolus vulgaris L. Journal of Biological Chemistry, 271, 26803–26809.CrossRefGoogle ScholarPubMed
Andrews, J.R., Bredenkamp, G.J. and Baker, N.R. (1993). Evaluation of the role of State transitions in determining the efficiency of light utilisation for CO2 assimilation in leaves. Photosynthesis Research, 38, 15–26.CrossRefGoogle ScholarPubMed
Andrews, T.J. and Muller, G.J. (1985). Photosynthetic gas exchange of the mangrove, Rhizophora stylosa, in its natural environment. Oecologia, 65, 449–455.CrossRefGoogle ScholarPubMed
Angelopulos, K., Dichio, B. and Xiloyannis, C. (1996). Inhibition of photosynthesis in olive trees (Olea europaea L.) during water stress and rewatering. Journal of Experimental Botany, 47, 1093–1100.CrossRefGoogle Scholar
Ankele, E., Kindgren, P., Pesquet, E., et al. (2007). In vivo visualisation of Mg-protoporphyrin IX, a coordinator of photosynthetic gene expression in the nucleus and the chloroplast. The Plant Cell, 19, 1964–1979.CrossRefGoogle Scholar
Antlfinger, A.E. and Wendel, L.F. (1997). Reproductive effort and floral photosynthesis in Spiranthes cernua (Orchidaceae). American Journal of Botany, 84, 769–780.CrossRefGoogle Scholar
Apel, K. and Hirt, H. (2004). Reactive oxygen species: Metabolism, oxidative stress and signal transduction. Annual Reviews of Plant Biology, 55, 373–399.CrossRefGoogle ScholarPubMed
Aphalo, P.J. and Jarvis, P.G. (1991). Do stomata respond to relative humidity. Plant, Cell and Environment, 14, 127–132.CrossRefGoogle Scholar
Aphalo, P.J. and Jarvis, P.G. (1993). An analysis of Ball’s empirical model of stomatal conductance. Annals of Botany, 72, 321–327.CrossRefGoogle Scholar
Apostol, S., Szalai, G., Sujbert, L., et al. (2006). Non-invasive monitoring of the light-induced cyclic photosynthetic electron flow during cold hardening in wheat leaves. Zeitschrift für Naturforschung, 61c, 734–740.Google Scholar
Aranda, I., Pardo, F., Gil, L., et al. (2004). Anatomical basis of the change in leaf mass per area and nitrogen investment with relative irradiance within the canopy of eight temperate tree species. Acta Oecologica, 25, 187–195.CrossRefGoogle Scholar
Aranda, I., Pardos, M., Puértolas, J., et al. (2007). Water-use efficiency in cork oak (Quercus suber) is modified by the interaction of water and light availability. Tree Physiology, 27, 671–677.CrossRefGoogle Scholar
Araujo, W.L., Dias, P.C., Moraes, G.A.B.K., et al. (2008). Limitations to photosynthesis in coffee leaves from different canopy positions. Plant Physiology and Biochemistry, 46, 884–890.CrossRefGoogle ScholarPubMed
Araus, J.L. (2004). The problems of sustainable water use in the Mediterranean and research requirements agriculture. Annals of Applied Biology, 144, 259–272.Google Scholar
Araus, J.L., Bort, J., Steduto, P., et al. (2003). Breeding cereals for Mediterranean conditions: ecophysiological clues for biotechnology application. Annals of Applied Biology, 142, 129–141.CrossRefGoogle Scholar
Araus, J.L., Brown, H.R., Febrero, A., et al. (1993). Ear photosynthesis, carbon isotope discrimination and the contribution of respiratory CO2 to differences in grain mass in Durum Wheat. Plant, Cell and Environment, 16, 383–392.CrossRefGoogle Scholar
Araus, J.L., Sánchez, C. and Cabrera-Bosquet, LL. (2010). Is heterosis in maize mediated through better water use?New Phytologist, 187, 392–406.CrossRefGoogle ScholarPubMed
Araus, J.L., Slafer, G.A., Royo, C., et al. (2008). Breeding for yield potential and stress adaptation in cereals. Critical Reviews in Plant Science, 27, 377–412.CrossRefGoogle Scholar
Arbaugh, M., Bytnerowicz, A., Grulke, N., et al. (2003). Photochemical smog effects in mixed conifer forests along a natural gradient of ozone and nitrogen deposition in the San Bernardino Mountains. Environment International, 29, 401–406.CrossRefGoogle ScholarPubMed
Arbona, V., López-Climent, M.F., Pérez-Clemente, R.M., et al. (2009). Maintenance of a high photosynthetic performance is linked to flooding tolerance in citrus. Environmental and Experimental Botany, 66, 135–142.CrossRefGoogle Scholar
Archetti, M. (2009). Classification of hypotheses on the evolution of autumn colours. Oikos, 118, 328–333.CrossRefGoogle Scholar
Archetti, M., Doring, T.F., Hagen, S.B., et al. (2009). Unravelling the evolution of autumn colours: an interdisciplinary approach. Trends Ecology Evolution, 24, 166–173.CrossRefGoogle Scholar
Ares, A., Fownes, J.H. and Sun, W. (2000). Genetic differentiation of intrinsic water-use efficiency in the Hawaiian native Acacia koa. International Journal of Plant Sciences, 161, 909–915.CrossRefGoogle Scholar
Ariana, D., Guyer, D.E. and Shrestha, B. (2006). Integrating multispectral reflectance and fluorescence imaging for defect detection on apples. Computers and Electronics in Agriculture, 50, 148–161.CrossRefGoogle Scholar
Arimura, G.I., Ozawa, R., Nishioka, T., et al. (2002). Herbivore-induced volatiles induce the emission of ethylene in neigbouring lima bean plants. Plant Journal, 29, 87–98.CrossRefGoogle Scholar
Armond, P.A., Björkman, O. and Staehelin, H.A. (1980) Dissociation of supramolecular complexes in chloroplast membranes. A manifestation of heat damage to the photosynthetic apparatus. Biochimica et Biophysica Acta, 601, 433–442.CrossRefGoogle ScholarPubMed
Armstrong, J.K., Williams, K., Huenneke, L.F., et al. (1988). Topographic position effects on growth depression of California Sierra Nevada pines during the 1982–83 El Niño. Arctic and Alpine Research, 20, 352–357.CrossRefGoogle Scholar
Armstrong, W. and Armstrong, J. (2005). Stem photosynthesis not pressurized ventilation is responsible for light-enhanced oxygen supply to submerged roots of alder (Alnus glutinosa). Annals of Botany, 96, 591–612.CrossRefGoogle Scholar
Arneth, A., Niinemets, Ü. and Pressley, S. (2007). Process-based estimates of terrestrial ecosystem isoprene emissions: incorporating the effects of a direct CO2-isoprene interaction. Atmospheric Chemistry and Physics, 7, 31–53.CrossRefGoogle Scholar
Arnon, D.I. (1949). Copper enzymes in isolated chloroplasts. Polyphenoloxidase in Beta vulgaris. Plant Physiology, 24, 1–15.CrossRefGoogle ScholarPubMed
Arnon, D.I. (1959). Conversion of light into chemical energy in photosynthesis. Nature, 184, 10–21.CrossRefGoogle ScholarPubMed
Aro, E.M., Virgin, I. and Andersson, B. (1993). Photoinhibition of photosystem II. Inactivation, protein damage and turnover. Biochicima et Biophysica Acta. (Bioenegetics), 1143, 113–134.CrossRefGoogle ScholarPubMed
Aronne, G. and De Micco, V. (2001). Seasonal dimorphism in the Mediterranean Cistus incanus L. subsp. incanus. Annals of Botany, 87, 789–794.CrossRefGoogle Scholar
Arp, W.J. (1991). Effects of source-sink relations on photosynthetic acclimation to elevated CO2. Plant Cell and Environment, 14, 869–875.CrossRefGoogle Scholar
Arquero, O., Barranco, D. and Benlloch, M. (2006). Potassium starvation increases stomatal conductance in olive trees. Hortscience, 41, 433–436.Google Scholar
Arulanantham, A., Rao, I. and Terry, N. (1990). Limiting factors in photosynthesis. VI. Regeneration of ribulose 1, 5-bisphosphate limits photosynthesis at low photochemical capacity. Plant Physiology, 93, 1466–1475.CrossRefGoogle ScholarPubMed
Asada, K. (1996). Radical production and scavenging in the chloroplasts. In: Advances in Photosynthesis Vol.5: Photosynthesis and the Environment (ed. Baker, N.R.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 123–150.Google Scholar
Asada, K. (1999). The water-water cycle in chloroplasts: scavenging of active oxygen and dissipation of excess photons. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 601–639.CrossRefGoogle Scholar
Asada, K. (2000). The water–water cycle as alternative photon and electron sinks. Philosophical Transactions of the Royal Society of London B, 355, 11419–1432.CrossRefGoogle ScholarPubMed
Aschan, G. and Pfanz, H. (2003). Non-foliar photosynthesis: a strategy of additional carbon acquisition. Flora, 198, 81–97.CrossRefGoogle Scholar
Aschan, G., Pfanz, H., Vodnik, D., et al. (2005). Photosynthetic performance of vegetative and reproductive structures of green hellebore (Helleborus viridis L. agg.). Photosynthetica, 43, 55–64.CrossRefGoogle Scholar
Ashley, D.A. and Boerma, H.R. (1989). Canopy photosynthesis and its association with seed yield in advanced generations of a soybean cross. Crop Science, 29, 1042–1045.CrossRefGoogle Scholar
Ashmore, M.R. (2002). Effects of oxidants at the whole plant and community level. In: Air pollution and Plant Life 2nd edn (eds Bell, J.N.B. and Treshow, M.), John Wiley and Sons, Chichester, USA, pp. 89–118.Google Scholar
Ashmore, M.R. (2005). Assessing the future global impacts of ozone on vegetation. Plant, Cell and Environment, 28, 949–964.CrossRefGoogle Scholar
Asmus, G.L. and Ferraz, L.C.C.B. (2002). Effect of population densities of Heterodera glycines race 3 on leaf area, photosynthesis and yield of soybean. Fitopatologia Brasileira, 27, 273–278.CrossRefGoogle Scholar
Asner, G.P. and Wessman, C.A. (1997). Scaling PAR absorption from the leaf to landscape level in spatially heterogeneous ecosystems. Ecological Modelling, 103, 81–97.CrossRefGoogle Scholar
Asner, G.P., Wessman, C.A., Bateson, A.C., et al. (2000). Impact of tissue, canopy, and landscape factors on the hyperspectral reflectance variability of arid ecosystems. Remote Sensing of Environment, 74, 69–84.CrossRefGoogle Scholar
Asokanthan, P., Johnson, R.W., Griffith, M., et al. (1997). The photosynthetic potential of canola embryos. Physiologia Plantarum, 101, 353–360.CrossRefGoogle Scholar
Athanasiou, K., Dyson, B.C., Webster, R.E. et al. (2010). Dynamic acclimation of photosynthesis increases plant fitness in changing environments. Plant Physiology, 152, 366–373.CrossRefGoogle ScholarPubMed
Atkin, O.K. and Macherel, D. (2009). The crucial role of plant mitochondria in orchestrating drought tolerance. Annals of Botany, 103, 581–597.CrossRefGoogle ScholarPubMed
Atkin, O.K. and Tjoelker, M.G. (2003). Thermal acclimation and the dynamic response of plant respiration to temperature. Trends Plant Science, 8, 343–351.CrossRefGoogle Scholar
Atkin, O.K., Botman, B. and Lambers, H. (1996). The causes of inherently slow growth in alpine plants: an analysis based on the underlying carbon economies of alpine and lowland Poa species. Functional Ecology, 10, 698–707.CrossRefGoogle Scholar
Atkin, O.K., Evans, J.R. and Siebke, K. (1998). Relationship between the inhibition of leaf respiration by light and enhancement of leaf dark respiration following light treatment. Australian Journal of Plant Physiology, 25, 437–443.CrossRefGoogle Scholar
Atkin, O.K., Scheurwater, I. and Pons, T.L. (2006). High thermal acclimation potential of both photosynthesis and respiration in two lowland Plantago species in contrast to an alpine congeneric. Global Change Biology, 12, 500–515.CrossRefGoogle Scholar
Atkin, O.K., Scheurwater, I. and Pons, T.L. (2007). Respiration as a percentage of daily photosynthesis in whole plants is homeostatic at moderate, but not high, growth temperatures. New Phytologist, 174, 367–380.CrossRefGoogle Scholar
Atkin, O.K., Evans, J.R., Ball, M.C., et al. (2000b). Leaf respiration of snow gum in the light and dark. Interactions between temperature and irradiance. Plant Physiology, 122, 915–923.CrossRefGoogle ScholarPubMed
Atkin, O.K., Millar, A.H., Gärdestrom, P., et al. (2000a). Photosynthesis, carbohydrate metabolism and respiration in leaves of higher plants. In: Photosynthesis, Physiology and Metabolism (eds Leegood, R. C., Sharkey, T. D. and von Caemmerer, S.) Kluwer Academic Publisher, London.Google Scholar
Atkins, C.A., Kuo, J., Pate, J.S., et al. (1977). Photosynthetic pod wall of pea (Pisum sativum L.). Distribution of carbon dioxide-fixing enzymes in relation to pod structure. Plant Physiology, 60, 779–786.CrossRefGoogle ScholarPubMed
Attiwill, P.M. and Adams, M.A. (1993). Nutrient cycling in forests. New Phytologist, 124, 561–582.CrossRefGoogle Scholar
Attiwill, P.M. and Clough, B.F. (1980). Carbon dioxide and water vapor exchange in the white mangrove (Avicennia marina). Photosynthetica, 14, 40–47.Google Scholar
Au, S.F. (1969). Internal leaf surface and stomatal abundance in arctic and alpine populations of Oxyria digyna. Ecology, 50, 131–134.CrossRefGoogle Scholar
Aubert, S., Assard, N., Boutin, J.P., et al. (1999). Carbon metabolism in the subantarctic Kerguelen cabbage Pringlea antiscorbutica R.Br.: environmental controls over carbohydrates and proline contents and relation to phenology. Plant Cell and Environment, 22, 243–254.CrossRefGoogle Scholar
Aubert, S., Choler, P., Pratt, J., et al. (2004). Methyl-beta-D-glucopyranoside in higher plants: accumulation and intraccellular localization in Geum montanum L. leaves and in model systems studied by 13C nuclear magnetic resonance. Journal of Experimental Botany, 55, 2179–2189.CrossRefGoogle Scholar
Aubinet, M., Heinesch, B. and Yernaux, M. (2003). Horizontal and vertical CO2 advection in a sloping forest. Boundary Layer Meteorology, 108, 397–417.CrossRefGoogle Scholar
Aubinet, M., Berbigier, P., Bernhofer, Ch., et al. (2005). Comparing CO2 storage and advection conditions at night at different Carboeuroflux sites. Boundary Layer Meteorology, 116, 63–94.CrossRefGoogle Scholar
Aubinet, M., Grelle, A., Ibrom, A., et al. (2000). Estimates of the annual net carbon and water exchange of forests: the EUROFLUX methodology. Advances in Ecological Research, 30, 113–175.CrossRefGoogle Scholar
Augspurger, C.K. and Bartlett, E.A. (2003). Differences in leaf phenology between juvenile and adult trees in a temperate deciduous forest. Tree Physiology, 23, 517–525.CrossRefGoogle Scholar
Augspurger, C.K., Cheeseman, J.M. and Salk, C.F. (2005). Light gains and physiological capacity of understory woody plants during physiological avoidance of canopy shade. Functional Ecology, 19, 537–546.CrossRefGoogle Scholar
Augusti, A. and Schleucher, J. (2007). The ins and outs of stable isotopes in plants. New Phytologist, 174, 473–475.CrossRefGoogle ScholarPubMed
Austin II, J. and Webber, A.N. (2005). Photosynthesis in Arabidopsis thaliana mutants with reduced chloroplast number. Photosynthesis Research, 85, 373–384.Google Scholar
Avenson, T.J., Cruz, J.A., Kanazawa, A., et al. (2005a). Regulating the proton budget of higher plant photosynthesis. Proceedings of the National Academy of Sciences (USA), 102, 9709–9713.CrossRefGoogle ScholarPubMed
Avenson, T.J., Kanazawa, A., Cruz, J.A., et al. (2005b). Integrating the proton circuit into photosynthesis: progress and challenges. Plant, Cell and Environment, 28, 97–109.CrossRefGoogle Scholar
Awramik, S.M. (1992). The oldest records of photosynthesis. Photosynthesis Research, 33, 75–89.CrossRefGoogle ScholarPubMed
Axelrod, D.I. (1966). Origin of deciduous and evergreen habits in temperate forests. Evolution, 20, 1–15.CrossRefGoogle ScholarPubMed
Baas, W.J. (1989). Secondary plant compounds, their ecological significance and consequences for the carbon budget. In: Causes and Consequences of Variation in Growth Rate and Productivity of Higher Plants (ed. Lambers, H.), SPB Academic Publishing, The Hague, pp. 313–340.Google Scholar
Bacelar, E.A., Mountibho-Pereira, J.M., Gonçalves, B.C., et al. (2007). Changes in growth, gas exchange, xylem hydraulic properties and water use efficiency of three olive cultivars under contrasting water availability regimes. Environmental and Experimental Botany, 60, 183–192.CrossRefGoogle Scholar
Bachereau, F., Marigo, G. and Asta, J. (1998). Effect of solar radiation (UV and visible) at high altitude on CAM-cycling and phenolic compound biosynthesis in Sedum album. Physiologia Plantarum, 104, 203–210.CrossRefGoogle Scholar
Bachmann, A., Fernándes-López, J., Ginsburg, J., et al. (1994). Stay green genotypes of Phaseolus vulgaris L.: chloroplast proteins and chlorophyll catabolites during foliar senescence. New Phytologist, 126, 593–600.CrossRefGoogle Scholar
Backhausen, J., Kitzmann, C., Horton, P., et al. (2000). Electron acceptors in isolated intact spinach chloroplasts act hierarchically to prevent over-reduction and competition for electrons. Photosynthesis Research, 64, 1–13.CrossRefGoogle ScholarPubMed
Badeck, F.W. (1995). Intra-leaf gradient of assimilation rate and optimal allocation of canopy nitrogen: a model on the implications of the use of homogeneous assimilation functions. Australian Journal of Plant Physiology, 22, 425–439.CrossRefGoogle Scholar
Badger, M.R. and Collatz, G.J. (1977). Studies on the kinetic mechanism of ribulose-1,5-bisphosphate carboxylase and oxygenase reactions, with particular reference to the effect of temperature on kinetic parameters. Carnegie Institution of Washington Year Book, 76, 355–361.Google Scholar
Badger, M.R., Björkman, O. and Armond, P.A. (1982). An analysis of photosynthetic response and adaptation to temperature in higher plants: temperature acclimation in the desert evergreen Nerium oleander L. Plant, Cell and Environment, 5, 85–99.Google Scholar
Badger, M.R., Sharkey, T.D. and von Caemmerer, S. (1984). The relationship between steady state gas exchange of bean leaves and the levels of carbon-reduction cycle intermediates. Planta, 160, 305–313.CrossRefGoogle ScholarPubMed
Badger, M.R., Spalding, M.H., Leegood, R.C., et al. (2000a). CO2 acquisition, concentration and fixation in cyanobacteria and algae. In: Advances in Photosynthesis, vol. 9: Photosynthesis, Physiology and Metabolism (ed. Leegood, R.C., Sharkey, T.D. and von Caemmerer, S.), Kluwer Academic, Dordrecht, Netherlands, pp. 369–397.Google Scholar
Badger, M.R., von Caemmerer, S., Ruuska, S., et al. (2000b). Electron flow to oxygen in higher plants and algae: rates and control of direct photoreduction (Mehler reaction) and Rubisco oxygenase. Philosophical Transactions of The Royal Society, 355, 1433–1446.CrossRefGoogle ScholarPubMed
Bahn, M., Rodeghiero, M., Anderson, M., et al. (2008). Soil respiration in European grasslands in relation to climate and assimilate supply. Ecosystems, 11, 1352–1367.CrossRefGoogle ScholarPubMed
Bailey Serres, J. and Voesenek, L.A.C.J. (2008). Flooding stress: acclimations and genetic diversity. Annual Review of Plant Biology, 59, 313–339.CrossRefGoogle ScholarPubMed
Baker, J.T., Kimb, S.H., Gitz, D.C., et al. (2004). A method for estimating carbon dioxide leakage rates in controlled-environment chambers using nitrous oxide. Environmental Experimental Botany, 51, 103–110.CrossRefGoogle Scholar
Baker, J.T., Van Pelt, S., Gitz, D.C., et al. (2009). Canopy gas exchange measurements of cotton in an open system. Agronomy Journal, 101, 52–59.CrossRefGoogle Scholar
Baker, N.R. (2008). Chlorophyll fluorescence: a probe of photosynthesis in vivo. Annual Reviews of Plant Biology, 59, 89–113.CrossRefGoogle ScholarPubMed
Baker, N.R. and Ort, D.R. (1992). Light and crop photosynthetic performance. In Topics in Photosynthesis. Crop Photosynthesis: Spatial and Temporal Determinants. (ed. Baker, N.R. and Thomas, H.), Elsevier, UK, pp. 289–312.Google Scholar
Baker, N.R. and Rosenqvist, E. (2004). Applications of chlorophyll fluorescence can improve crop production strategies: an examination of future possibilities. Journal of Experimental Botany, 55, 1607–1621.CrossRefGoogle ScholarPubMed
Baker, N.R., Harbinson, J. and Kramer, D.M. (2007). Determining the limitations and regulation of photosynthetic energy transduction in leaves. Plant, Cell and Environment, 30, 1107–1125.CrossRefGoogle ScholarPubMed
Baker, N.R., Oxborough, K., Lawson, T., et al. (2001). High resolution imaging of photosynthetic activities of tissues, cells and chloroplasts in leaves. Journal of Experimental Botany, 52, 615–621.CrossRefGoogle ScholarPubMed
Balachandran, S. and Osmond, C.B. (1994). Susceptibility of tobacco leaves to photoinhibition following infection with two strains of tobacco mosaic virus under different light and nitrogen nutrition regimes. Plant Physiology, 104, 1051–1057.CrossRefGoogle ScholarPubMed
Balachandran, S., Hurry, V.M., Kelley, S.E., et al. (1997). Concepts of plant biotic stress: some insights into the stress physiology of virus-infected plants, from the perspective of photosynthesis. Physiologia Plantarum, 100, 203–213.CrossRefGoogle Scholar
Balaguer, L., Manrique, E., de 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., Pugnaire, F. I., Martínez-Ferri, E., et al. (2002). Ecophysiological significance of chlorophyll loss and reduced photochemical efficiency under extreme aridity in Stipa tenacissima L. Plant and Soil, 240, 343–352.CrossRefGoogle Scholar
Baldocchi, D. (1994). An analytical solution for coupled leaf photosynthesis and stomatal conductance models. Tree Physiology, 14, 1069–1079.CrossRefGoogle ScholarPubMed
Baldocchi, D. (2003). Assessing the eddy covariance technique for evaluating carbon dioxide exchange rates of ecosystems: past, present and future. Global Change Biology, 9, 479–492.CrossRefGoogle Scholar
Baldocchi, D. and Collineau, S. (1994). The physical nature of solar radiation in heterogeneous canopies: spatial and temporal attributes. In: Exploitation of Environmental Heterogeneity by Plants (eds Caldwell, M.M. and Pearcy, R.W.), Academic Press, San Diego, California, pp. 21–71.Google Scholar
Baldocchi, D.D. (1993). Scaling water vapour and carbon dioxide exchanges from leaves to canopy: rules and tool. In: Scaling Physiological Processes: Leaf to Global (eds Ehleringer, J.R. and Field, C.B.), Academic Press, San Diego, USA, pp. 77–114.Google Scholar
Baldocchi, D.D. (2008). ‘Breathing’ of the terrestrial biosphere: lessons learned from a global network of carbon dioxide flux measurement systems. Australian Journal of Botany, 56, 1–26.CrossRefGoogle Scholar
Baldocchi, D.D. and Amthor, J.S. (2001). Canopy photosynthesis: history, measurements, and models. In: Terrestrial Global Productivity: Past, Present, and Future (eds Mooney, H.A., Saugier, B. and Roy, J.), Academic Press, Inc, San Diego. pp. 9–31.Google Scholar
Baldocchi, D.D. and Harley, P.C. (1995). Scaling carbon dioxide and water vapour exchange from leaf to canopy in a deciduous forest. II. Model testing and application. Plant Cell and Environment, 18, 1157–1173.CrossRefGoogle Scholar
Baldocchi, D.D. and Valentini, R. (associated eds.) (1996). Thematic issue: Strategies for monitoring and modelling CO2 and water vapour fluxes over terrestrial ecosystems. Global Change Biology, 2, 159–318.CrossRefGoogle Scholar
Baldocchi, D.D., Wilson, K.B. and Gu, L. (2002). How the environment, canopy structure and canopy physiological functioning influence carbon, water and energy fluxes of a temperate broad-leaved deciduous forest – an assessment with the biophysical model CANOAK. Tree Physiology, 22, 1065–1077.CrossRefGoogle Scholar
Baldocchi, D.D., Xu, L. and Kiang, N.Y. (2004). How plant functional-type, weather, seasonal drought, and soil physical properties alter water and energy fluxes of an oak-grass savanna and an annual grassland. Agricultural and Forest Meteorology, 123, 13–39.CrossRefGoogle Scholar
Baldocchi, D.D., Falge, E., Gu, L., et al. (2001). FLUXNET: a new tool to study the temporal and spatial variability of ecosystem-scale carbon dioxide, water vapor and energy flux densities. Bulletin of the American Meteorological Society, 82, 2415–2434.2.3.CO;2>CrossRefGoogle Scholar
Baldocchi, D.D., Finnigan, J.J., Wilson, K., et al. (2000). On measuring net ecosystem carbon exchange over tall vegetation on complex terrain. Boundary Layer Meteorology, 96, 257–291.CrossRefGoogle Scholar
Baldocchi, D.D., Hicks, B.B. and Meyers, T.P. (1988). Measuring biosphere-atmosphere exchanges of biologically related gases with micrometeorological methods. Ecology, 69, 1331–1340.CrossRefGoogle Scholar
Baldocchi, D.D., Valentini, R., Running, S., et al. (1996). Strategies for measuring and modelling carbon dioxide and water vapour fluxes over terrestrial ecosystems. Global Change Biology, 2, 159–168.CrossRefGoogle Scholar
Baldwin, A., Egnotovich, M., Ford, M., et al. (2001). Regeneration in fringe mangrove forests damaged by Hurricane Andrew. Plant Ecology, 157, 149–162.CrossRefGoogle Scholar
Baldwin, I.T. (2001). An ecologically motivated analysis of plant-herbivore interactions in native tobacco. Plant Physiology, 127, 1449–1458.CrossRefGoogle ScholarPubMed
Baldwin, I.T. and Callahan, P. (1993). Autotoxicity and chemical defense: nicotine accumulation and carbon gain in solanaceous plants. Oecologia, 94, 534–541.CrossRefGoogle ScholarPubMed
Ball, J.T., Woodrow, I.E. and Berry, J.A. (1987). A model predicting stomatal conductance and its contribution to the control of photosynthesis under different environmental conditions. In: Progress in Photosynthesis Research, Vol. IV (ed. Biggins, I.), Martinus-Nijhoff Publishers, Dordrecht, Netherlands, pp. 221–224.Google Scholar
Ball, M.C. (2002). Interactive effects of salinity and irradiance on growth: implications for mangrove forest structure along salinity gradients. Trees, 16, 126–139.CrossRefGoogle Scholar
Ball, M.C. and Critchley, C. (1982). Photosynthetic responses to irradiance by the grey mangrove, Avicennia marina, grown under different light regimes. Plant Physiology, 70, 1101–1106.CrossRefGoogle ScholarPubMed
Ballare, C.L., Scopel, A.L., Stapleton, A.E., et al. (1996). Solar ultraviolet-B radiation affects seedling emergence, DNA integrity, plant morphology, growth rate, and attractiveness to herbivore insects in Datura ferox. Plant Physiology, 112, 161–170.CrossRefGoogle ScholarPubMed
Baltzer, J.L. and Thomas, S.C. (2007a). Determinants of whole-plant light requirements in Bornean rain forest tree saplings. Journal of Ecology, 95, 1208–1221.CrossRefGoogle Scholar
Baltzer, J.L. and Thomas, S.C. (2007b). Physiological and morphological correlates of whole-plant light compensation point in temperate deciduous tree seedlings. Oecologia, 153, 209–223.CrossRefGoogle ScholarPubMed
Baltzer, J.L., Davies, S.J., Bunyavejchewin, S., et al. (2008). The role of desiccation tolerance in determining tree species distributions along the Malay Thai Peninsula. Functional Ecology, 22, 221–231.CrossRefGoogle Scholar
Baltzer, J.L., Thomas, S.C., Nilus, R., et al. (2005). Edaphic specialization in tropical trees: physiological correlates and responses to reciprocal transplantation. Ecology, 86, 3063–3077.CrossRefGoogle Scholar
Bansal, S. and Germino, M.J. (2008). Carbon balance of conifer seedlings at timberline: relative changes in uptake, storage, and utilization. Oecologia, 158, 217–227.CrossRefGoogle ScholarPubMed
Barak, P. and Helmke, P.A. (1993). The chemistry of zinc. In: Zinc in Soils and Plants, Developments in Plants and Soil Sciences (ed. Robson, A). Kluwer Academic Press, New York, USA, pp. 1–13.Google Scholar
Barbehenn, R.V., Karowe, D.N. and Zhong, C. (2004).Performance of a generalist grasshopper on a C3 and a C4 grass: compensation for the effects of elevated CO2 on plant nutritional quality. Oecologia, 140, 96–103.CrossRefGoogle Scholar
Barber, J. (2008a). Photosynthetic generation of oxygen. Philosophical Transactions of the Royal Society B, 363, 2665–2674.CrossRefGoogle ScholarPubMed
Barber, J. (2008b). Crystal structure of the oxygen-evolving complex of Photosystem II. Inorganic Chemistry, 47, 1700–1710.CrossRefGoogle ScholarPubMed
Barbour, M.M. (2007). Stable oxygen isotope composition of plant tissue: a review. Functional Plant Biology, 34, 83–94.CrossRefGoogle Scholar
Barbour, M.M. and Buckley, T.N. (2007). The stomatal response to evaporative demand persists at night in Ricinus communis plants with high nocturnal conductance. Plant, Cell and Environment, 30, 711–721.CrossRefGoogle ScholarPubMed
Barbour, M.M. and Farquhar, G.D. (2000). Relative humidity and ABA-induced variation in carbon and oxygen isotope ratios of cotton leaves. Plant, Cell and Environment, 23, 473–485.CrossRefGoogle Scholar
Barbour, M.M., Cernusak, L.A. and Farquhar, G.D. (2005). Factors affecting the oxygen isotope ratio of plant organic material. In: Stable Isotopes and Biosphere-Atmosphere Interactions (eds Flanagan, L.B., Ehleringer, J.R. and Pataki, D.E.), Elsevier Academic Press, San Diego, pp. 9–28.Google Scholar
Barbour, M.M., Fischer, R.A., Sayre, K.D., et al. (2000b). Oxygen isotope ratio of leaf and grain material correlates with stomatal conductance and grain yield in irrigated wheat. Australian Journal of Plant Physiology, 27, 625–637.Google Scholar
Barbour, M.M., McDowell, N.G., Tcherkez, G., et al. (2007). A new measurement technique reveals rapid post-illumination changes in the carbon isotope composition of leaf-respired CO2. Plant, Cell and Environment, 30, 469–482.CrossRefGoogle ScholarPubMed
Barbour, M.M., Shurr, U., Henry, B.K., et al. (2000a). Variation in the oxygen isotope ratio of phloem sap sucrose from castor bean: evidence in support of the Péclet effect. Plant Physiology, 123, 671–679.CrossRefGoogle ScholarPubMed
Barbour, M.M., Warren, C.R., Farquhar, G.D., et al. (2010). Variability in mesophyll conductance between barley genotypes, and effects on transpiration efficiency and carbon isotope discrimination. Plant, Cell and Environment, 33, 1176–1185.Google ScholarPubMed
Bar-Even, A., Noor, E., Lewis, N.E., et al. (2010). Design and analysis of synthetic carbon fixation pathways. Proceedings of the National Academy of Sciences USA, 107, 8889–8894.CrossRefGoogle ScholarPubMed
Bariac, T., Gonzales-Dunia, J., Tardieu, F., et al. (1994). Spatial variation of the isotopic composition of water (18O, 2H) in organs of aerophytic plants. 1. Assessment under laboratory conditions. Chemical Geology, 115, 307–315.CrossRefGoogle Scholar
Barker, D.A., Seaton, G.G.R. and Robinson, S.A. (1997). Internal and external photoprotection in developing leaves of the CAM plant Cotyledon orbiculata. Plant Cell and Environment, 20, 617–624.CrossRefGoogle Scholar
Barker, D.H., Vanier, C., Naumburg, E., et al. (2006). Enhanced monsoon precipitation and nitrogen deposition affect leaf traits and photosynthesis differently in spring and summer in the desert shrub Larrea tridentata. New Phytologist, 169, 799–808.CrossRefGoogle ScholarPubMed
Barker, E.R., Press, M.C., Scholes, J.D., et al. (1996). Interactions between the parasitic angiosperm Orobanche aegyptiaca and its tomato host: growth and biomass allocation. New Phytologist, 133, 637–642.CrossRefGoogle Scholar
Barnes, B.B., Zak, D.R., Denton, S.R., et al. (1998). Forest Ecology, 4th edn. John Wiley and Sons Inc., New York, USA.Google Scholar
Barnes, J., Davison, A., Balaguer, L., et al. (2007). Resistance to air pollutants: from cell to community. In: Functional Plant Ecology (eds Pugnaire, F.I. and Valladares, F.) 2nd edn: CRC Press, USA, pp. 601–626.Google Scholar
Barnes, P.W. and Archer, S. (1996). Influence of an overstorey tree (Prosopis glandulosa) on associated shrubs in a savanna parkland: implications for patch dynamics. Oecologia, 105, 493–500.CrossRefGoogle Scholar
Baroli, I. and Niyogi, K.K. (2000). Molecular genetics of xanthophyll-dependent photoprotection in green algae and plants. Philosophical Transactions of the Royal Society B, 355, 1385–1394.CrossRefGoogle ScholarPubMed
Baroli, I., Price, G.D., Badger, M.R., et al. (2008). The contribution of photosynthesis to the red light response of stomatal conductance. Plant Physiology, 146, 737–747.CrossRefGoogle ScholarPubMed
Barón, M., Arellano, J.B. and López Gorgé, J. (1995). Copper and photosystem II: a controversial relationship. Physiologia Plantarum, 94, 174–180.CrossRefGoogle Scholar
Barrantes, O., Moliner, E., Plaza, M., et al. (1997). Preliminary results of an SO2 experiment with Pinus halepensis Mill. seedlings in open top chambers. In: Impact of Global Change on Tree Physiology and Forest Ecosystems (eds Mohren, G.M.J., Kramer, K. and Sabaté, S.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 111–118.Google Scholar
Barry, J.C., Morgan, M.E., Flynn, L.J., et al. (2002). Faunal and environmental change in the late Miocene Siwaliks of northern Pakistan. Paleobiology Memoirs, 28, 1–55.CrossRefGoogle Scholar
Barry, R.G. (2008). Mountain, weather and climate. 3rd edn. Cambridge University Press, Cambridge, UK.
Barta, C. and Loreto, F. (2006). The relationship between the methyl-erythritol phosphate pathway leading to emission of volatile isoprenoids and abscisic acid content in leaves. Plant Physiology, 141, 1676–1683.CrossRefGoogle ScholarPubMed
Barták, M., Gloser, J. and Hájek, J. (2005). Visualized photosynthetic characteristics of the lichen Xanthoria elegans related to daily courses of light, temperature and hydration: a field study from Galindez Island, maritime Antarctica. Lichenologist, 37, 433–443.CrossRefGoogle Scholar
Barták, M.L., Váczi, P., Hájek, J., et al. (2006). Low-temperature limitation of primary photosynthetic processes in Antarctic lichens Umbilicaria antarctica and Xanthoria elegans. Polar Biology, 31, 47–51.CrossRefGoogle Scholar
Bartelink, H.H. (1998). A model of dry matter partitioning in trees. Tree Physiology, 18, 91–101.CrossRefGoogle ScholarPubMed
Bartels, D. and Salamini, F. (2001). Desiccation tolerance in the resurrection plant Craterostigma plantagineum: a contribution to the study of drought tolerance at the molecular level. Plant Physiology, 127, 1346–1353.CrossRefGoogle Scholar
Bartoli, C.G., Gómez, F., Gergoff, G., et al. (2005). Up-regulation of the mitochondrial alternative oxidase pathway enhances photosynthetic electron transport under drought conditions. Journal of Experimental Botany, 56, 1269–1276.CrossRefGoogle ScholarPubMed
Barton, A.M., Fetcher, N. and Redhead, S. (1989). The relationship between treefall gap size and light-flux in a Neotropical rain-forest in Costa-Rica. Journal of Tropical Ecology, 5, 437–439.CrossRefGoogle Scholar
Baruch, Z. and Bilbao, B. (1999). Effects of fire and defoliation on the life history of native and invader C-4 grasses in a Neotropical savanna. Oecologia, 119, 510–520.CrossRefGoogle Scholar
Basak, U.C., Das, A.B. and Das, P. (1996). Chlorophylls, carotenoids, proteins and secondary metabolites in leaves of 14 species of mangrove. Bulletin of Marine Science, 58, 654–659.Google Scholar
Basha, E., Lee, G.J., Breci, L.A., et al. (2004). The identity of proteins associated with a small heat shock protein during heat stress in vivo indicates that these chaperones protect a wide range of cellular functions. Journal of Biological Chemistry, 279, 7566–7575.CrossRefGoogle ScholarPubMed
Basile, B., Reidel, E.J., Weinbaum, S.A., et al. (2003). Leaf potassium concentration, CO2 exchange and light interception in almond trees (Prunus dulcis (Mill) D.A. Webb). Scientia Horticulturae, 98, 185–194.CrossRefGoogle Scholar
Bassham, J.A. and Calvin, M. (1957). The Path of Carbon in Photosynthesis. Prentice-Hall, N.J., USA, p. 104.Google Scholar
Baszynski, T., Wajda, L., Krol, M., et al. (1980). Photosynthetic activities of cadmium-treated tomato plants. Physiologia Plantarum, 48, 365–370.CrossRefGoogle Scholar
Bauer, H., Nagele, M., Comploj, M., et al. (1994). Photosynthesis in cold-acclimated leaves of plants with various degrees of freezing tolerance. Physiologia Plantarum, 91, 403–412.CrossRefGoogle Scholar
Bauerle, W.L., Hinckley, T.M., Chermák, J., et al. (1999). The canopy water relations of old-growth Douglas-fir trees. Trees, 13, 211–217.CrossRefGoogle Scholar
Baxter, C.J., Redestig, H., Schauer, N., et al. (2007). The metabolic response of heterotrophic Arabidopsis cells to oxidative stress. Plant Physiology, 143, 312–325.CrossRefGoogle ScholarPubMed
Bazzaz, F.A., Carlson, R.W. and Harper, J.L. (1979). Contribution to reproductive effort by photosynthesis of flowers and fruits. Nature, 279, 554–555.CrossRefGoogle Scholar
Beadle, N.C.W. (1966). Soil phosphate and its role in molding segments of the Australian flora and vegetation with special reference to xeromorphy and sclerophylly. Ecology, 47, 992–1007.CrossRefGoogle Scholar
Bean, R.C., Porter, G.G. and Barr, K.B. (1963). Photosynthesis and respiration in developing fruits. III. Variations in photosynthesis capacity during colour change. Plant Physiology, 38, 285–290.CrossRefGoogle Scholar
Beauchamp, J., Wisthaler, A., Hansel, A., et al. (2005). Ozone induced emissions of biogenic VOC from tobacco: relations between ozone uptake and emission of LOX products. Plant, Cell and Environment, 28, 1334–1343.CrossRefGoogle Scholar
Beauford, W., Barber, J. and Barringer, A.R. (1977). Uptake and distribution of mercury within higher plants. Physiologia Plantarum, 39, 261–265.CrossRefGoogle Scholar
Becher, M., Talke, I.N., Krall, L., et al. (2004). Cross-species microarray transcript profiling reveals high constitutive expression of metal homeostasis genes in shoots of the zinc hyperaccumulator Arabidopsis halleri. Plant Journal, 37, 251–268.CrossRefGoogle ScholarPubMed
Beck, C. (2005). Signaling pathways from the chloroplast to the nucleus. Planta, 222, 743–756.CrossRefGoogle ScholarPubMed
Beck, E., Senser, M., Scheibe, R., et al. (1982). Frost avoidance and freezing tolerance in Afroalpine ‘giant rosette’ plants. Plant Cell and Environment, 5, 215–222.Google Scholar
Bednarz, C.W. and van Iersel, M.W. (1999). Continuous whole plant carbon dioxide exchange rates in cotton treated with pyrithiobac. The Journal of Cotton Science, 3, 53–59.Google Scholar
Bednarz, C.W. and van Iersel, M.W. (2001). Temperature response of whole-plant CO2 exchange rates of four upland cotton cultivars differing in leaf shape and leaf pubescence. Communications in Soil Science and Plant Analysis, 32, 2485–2501.CrossRefGoogle Scholar
Beer, C., Ciais, P., Reichstein, M., et al. (2009). Temporal and among-site variability of inherent water use efficiency at the ecosystem level. Global Biogeochemical Cycles, 23, GB2018.CrossRefGoogle Scholar
Beer, C., Reichstein, M., Ciais, P., et al. (2007). Mean annual GPP of Europe derived from its water balance. Geophysical Research Letters, 34, L05401.CrossRefGoogle Scholar
Beer, C., Reichstein, M., Tomelleri, E. et al. (2010). Terrestrial gross carbon dioxide uptake: global distribution and covariation with climate. Science, 329, 834–838.CrossRefGoogle ScholarPubMed
Beerling, D.J. (2005). Leaf evolution: gases, genes and geochemistry. Annals of Botany, 96, 345–352.CrossRefGoogle ScholarPubMed
Beerling, D.J. and Kelly, C.K. (1996). Evolutionary comparative analyses of the relationship between leaf structure and function. The New Phytologist, 134, 35–51.CrossRefGoogle Scholar
Beerling, D.J. and Osborne, C.P. (2006). The origin of the savanna biome. Global Change Biology, 12, 2023–2031.CrossRefGoogle Scholar
Beerling, D.J. and Rundgren, M. (2000). Leaf metabolic and morphological responses of dwarf willow (Salix herbacea) in the Sub-Arctic to the past 9000 years of global environmental change. New Phytologist, 145, 257–269.CrossRefGoogle Scholar
Beerling, D.J. and Woodward, F.I. (1996). Paleo-ecophysiological perspectives on plant responses to global change. Trends in Ecology and Evolution, 11, 20–23.CrossRefGoogle ScholarPubMed
Beerling, D.J. and Woodward, F.I. (1997). Changes in land plant function over the Phanerozoic: recontructions based on the fossil record. Botanical Journal of the Linnean Society, 124, 137–153.CrossRefGoogle Scholar
Beerling, D.J. and Woodward, F.I. (2001). Vegetation and the Terrestrial Carbon Cycle: Modelling the First 400 Million Years. Cambridge University Press,Cambridge, UK.CrossRefGoogle Scholar
Beerling, D.J., McElwain, J.C. and Osborne, C.P. (1998). Stomatal responses of the ‘living fossil’ Ginkgo biloba L. to changes in atmospheric CO2 concentrations. Journal of Experimental Botany, 49, 1603–1607.Google Scholar
Beerling, D.J., Osborne, C.P., Chaloner, W.G. (2001). Evolution of leaf-form in land plants linked to atmospheric CO2 decline in the Late Paleozoic era. Nature, 410, 352–354.CrossRefGoogle Scholar
Behrensmeyer, A.K., Quade, J., Cerling, T.E., et al. (2007). The structure and rate of late Miocene expansion of C4 plants: evidence from lateral variation in stable isotopes in paleosols of the Siwalik Group, northern Pakistan. Geological Society of America Bulletin, 119, 1486–1505.CrossRefGoogle Scholar
Belkhodja, R., Morales, F., Quílez, R., et al. (1998). Iron deficiency causes changes in chlorophyll fluorescence due to the reduction in the dark of the photosystem II acceptor side. Photosynthesis Research, 56, 265–276.CrossRefGoogle Scholar
Bellingham, P.J., Kohyama, T. and Aiba, S.I. (1996). The effects of a typhoon on Japanese warm temperate rainforests. Ecological Research, 11, 229–247.CrossRefGoogle Scholar
Belsky, A.J., Amundson, R.G.Duxbury, J.M. et al. (1989). The effects of trees on their physical, chemical, and biological environments in a semi-arid savanna in Kenya. Journal of Applied Ecology, 26, 1005–1024.CrossRefGoogle Scholar
Bendall, D.S. and Manasse, R.S. (1995). Cyclic photophosphorylation and electron transport. Biophysica et Biochimica Acta (Bioenergetics), 1229, 23–38.CrossRefGoogle Scholar
Bender, M., Sowers, T. and Labeyrie, L. (1994). The Dole effect and its variations during the last 130,000 years as measured in the Vostok ice core. Global Biogeochemical Cycles, 8, 363–376.CrossRefGoogle Scholar
Bender, M.M. (1968). Mass spectrometric studies of carbon-13 variations in corn and other grasses. Radiocarbon, 10, 468–472.CrossRefGoogle Scholar
Bender, M.M. (1971). Variations in the 13C/12C ratios of plants in relations to the pathway of photosynthetic carbon dioxide fixation. Phytochemistry, 10, 1239–1244.CrossRefGoogle Scholar
Bender, M.M., Rouhani, I., Vines, H.M., et al. (1973). 13C/12C ratio changes in crassulacean acid metabolism plants. Plant Phisiology, 52, 427–430.CrossRefGoogle Scholar
Benkeblia, N., Shinano, T. and Osaki, M. (2007). Metabolite profiling and assessment of metabolome compartmentation of soybean leaves using non-aqueous fractionation and GC-MS analysis. Metabolomics, 2, 297–305.CrossRefGoogle Scholar
Bennoun, P. (2001). Chlororespiration and the process of carotenoid biosynthesis. Biochimica et Biophysica Acta – Bioenergetics, 1506, 133–142.CrossRefGoogle ScholarPubMed
Benowicz, A., Guy, R.D. and El-Kassaby, Y.A. (2000). Geographic pattern of genetic variation in photosynthetic capacity and growth in two hardwood species from British Columbia. Oecologia, 123, 168–174.CrossRefGoogle ScholarPubMed
Benzing, D.H. (1990). Vascular Epiphytes: General Biology and Related Biota. Cambridge University Press, Cambridge, UK.CrossRefGoogle Scholar
Berard, R.G. and Thurtell, G.W. (1990). Respiration measurements of maize plants using a whole-plant enclosure system. Agronomy Journal, 82, 641–643.CrossRefGoogle Scholar
Berenbaum, M.R. and Zangerl, A.R. (2008). Facing the future of plant-insect interaction research: le retour a la “raison d’etre”. Plant Physiology, 146, 804–811.CrossRefGoogle ScholarPubMed
Berger, B., Parent, B. and Tester, M. (2010). High-throughput shoot imaging to study drought responses. Journal of Experimental Botany, 61, 3519–3528.CrossRefGoogle ScholarPubMed
Berger, S., Benediktyová, Z., Matouš, K., et al. (2007). Visualization of dynamics of plant-pathogen interaction by novel combination of chlorophyll fluorescence imaging and statistical analysis: differential effects of virulent an avirulent strains of P. syringae and of oxylipins on A. thaliana. Journal of Experimental Botany, 58, 797–806.CrossRefGoogle ScholarPubMed
Berger, S., Papadopoulos, M., Schreiber, U., et al. (2004). Complex regulation of gene expression, photosynthesis and sugar levels by pathogen infection in tomato. Physiologia Plantarum, 122, 419–428.CrossRefGoogle Scholar
Beringer, J., Hutley, L.B., Tapper, N.J., et al. (2007). Savanna fires and their impact on net ecosystem productivity in North Australia. Global Change Biology, 13, 990–1004.CrossRefGoogle Scholar
Bernacchi, C.J., Calfapietra, C., Davey, P.A., et al. (2003b). Photosynthesis and stomatal conductance responses of poplars to free-air CO2 enrichment (PopFACE) during the first growth cycle and immediately following coppice. New Phytologist, 159, 609–621.CrossRefGoogle Scholar
Bernacchi, C.J., Hollinger, S.E. and Meyers, T. (2005b). The conversion of the corn/soybean ecosystem to no-till agriculture may result in a carbon sink. Global Change Biology, 11, 1867–1872.Google Scholar
Bernacchi, C.J., Kimball, B., Quarles, D., et al. (2007). Decreases in stomatal conductance of soybean under open-air elevation of [CO2] are closely coupled with decreases in ecosystem evapotranspiration. Plant Physiology, 143, 134–144.CrossRefGoogle Scholar
Bernacchi, C.J., Leakey, A.D.B., Heady, L.E., et al. (2006). Hourly and seasonal variation in photosynthesis and stomatal conductance of soybean grown at future CO2 and ozone concentrations for 3 years under fully open-air field conditions. Plant, Cell and Environment, 29, 2077–2090.CrossRefGoogle Scholar
Bernacchi, C.J., Morgan, P.B., Ort, D.R., et al. (2005a). The growth of soybean under free air [CO2] enrichment (FACE) stimulates photosynthesis while decreasing in vivo Rubisco capacity. Planta, 220, 434–446.CrossRefGoogle ScholarPubMed
Bernacchi, C.J., Pimentel, C. and Long, S.P. (2003a). In vivo temperature response functions of parameters required to model RuBP-limited photosynthesis. Plant, Cell and Environment, 26, 1419–1430.CrossRefGoogle Scholar
Bernacchi, C.J., Portis, A.R., Nakano, H., et al. (2002). Temperature response of mesophyll conductance: implications for the determination of Rubisco enzyme kinetics and for limitations to photosynthesis in vivo. Plant Physiology, 130, 1992–1998.CrossRefGoogle ScholarPubMed
Bernacchi, C.J., Singsaas, E.L., Pimentel, C., et al. (2001). Improved temperature response functions for models of Rubisco-limited photosynthesis. Plant Cell and Environment, 24, 253–259.CrossRefGoogle Scholar
Berner, R.A. (1990). Atmospheric carbon dioxide levels over phanerozoic time. Science, 249, 1382–1386.CrossRefGoogle ScholarPubMed
Berner, R.A. and Canfield, D.E. (1989). A new model for atmospheric oxygen over phaenerozoic time. American Journal of Science, 289, 333–361.CrossRefGoogle ScholarPubMed
Berni, J.A., Zarco-Tejada, P.J., Suarez, L., et al. (2009). Thermal and narrowband multispectral remote sensing for vegetation monitoring from an unmanned aerial vehicle. IEEE Transactions on Geoscience and Remote Sensing, 47, 722–738.CrossRefGoogle Scholar
Berninger, F. (1997). Effects of drought and phenology on GPP a simulation study along a geographical gradient. Functional Ecology, 11, 33–42.CrossRefGoogle Scholar
Berova, M., Stoeva, N., Zlatev, Z., et al. (2007). Physiological changes in bean (Phaseolus vulgaris L.) leaves, infected by the most important bean diseases. Journal of Central European Agriculture, 8, 57–62.Google Scholar
Berry, J.A. and Björkman, O. (1980). Photosynthetic response and adaptation to temperature in higher plants. Annual Review of Plant Physiology, 31, 491–543.CrossRefGoogle Scholar
Bertamini, M., Muthuchelian, K. and Nedunchezhian, N. (2004). Effect of grapevine leafroll on the photosynthesis of field grown grapevine plants (Vitis vinifera L. cv. Lagrein). Journal of Phytopathology, 152, 145–152.CrossRefGoogle Scholar
Bertone, P. and Snyder, M. (2007). Prospects and challenges in proteomics. Plant Physiology, 138, 560–562.CrossRefGoogle Scholar
Bertsch, W.F. and Azzi, J.R. (1965). A relative maximum in the decay of long-term delayed light emission from the photosynthetic apparatus. Biochimica et Biophysica Acta, 94, 15–26.CrossRefGoogle ScholarPubMed
Berveiller, D., Kierzkowski, D. and Damesin, C. (2007a). Interspecific variability of stem photosynthesis among tree species. Tree Physiology, 27, 53–61.CrossRefGoogle ScholarPubMed
Berveiller, D., Vidal, J., Degrouard, J., et al. (2007b). Tree stem phosphoenolpyruvate carboxylase (PEPC): lack of biochemical and localization evidence for a C4-like photosynthesis system. New Phytologist, 176, 775–781.CrossRefGoogle ScholarPubMed
Besson-Bard, A., Pugin, A. and Wendehenne, D. (2008). New insights into nitric oxide signaling in plants. Annual Review of Plant Biology, 59, 21–39.CrossRefGoogle ScholarPubMed
Betts, R.A., Boucher, O., Collins, M., et al. (2007). Projected increase in continental runoff due to plant responses to increasing carbon dioxide. Nature, 448, 1037–1041.CrossRefGoogle ScholarPubMed
Betzelberger, A.M., Gillespie, K.M., McGrath, J.M., et al. (2010). Effects of chronic elevated ozone concentration on antioxidant capacity, photosynthesis and seed yield of 10 soybean cultivars. Plant, Cell and Environment, 33, 1569–1581.Google ScholarPubMed
Beuker, E. (1994). Adaptations to climate change of the timing of bud burst of Pinus sylvestris L. And Picea abies Karst. Trees, 14, 961–970.Google Scholar
Beyschlag, W., Kresse, F., Ryel, R.J., et al. (1994). Stomatal patchiness in conifers: experiments with Picea abies (L.) Karst. and Abies alba Mill. Trees: Structure and Function, 8, 132–138.CrossRefGoogle Scholar
Beyschlag, W., Pfanz, H. and Ryel, R.J. (1992). Stomatal patchiness in Mediterranean evergreen sclerophylls. Phenomenology and consequences for the interpretation of the midday depression in photosynthesis and transpiration. Planta, 187, 546–553.CrossRefGoogle ScholarPubMed
Bickford, C.P., Hanson, D.T. and McDowell, N.G. (2010). Influence of diurnal variation in mesophyll conductance on modeled 13C discrimination: results from a field study. Journal of Experimental Botany, 61, 3223–3233.CrossRefGoogle Scholar
Bickford, C.P., McDowell, N.G., Erhardt, E.B., et al. (2009). High-frequency field measurements of diurnal carbon isotope discrimination and internal conductance in a semi-arid species, Juniperus monosperma. Plant, Cell and Environment, 32, 796–810.CrossRefGoogle Scholar
Bieleski, R.L. (1973). Phosphate pools, phosphate transport, and phosphate availability. Annual Review of Plant Physiology, 24, 225–252.CrossRefGoogle Scholar
Bigras, F.J. and Bertrand, A. (2006). Responses of Picea mariana to elevated CO2 concentration during growth, cold hardening and dehardening: phenology, cold tolerance, photosynthesis and growth. Tree Physiology, 26, 875–888.CrossRefGoogle Scholar
Bilger, H.W., Schreiber, U. and Lange, O.L. (1984). Determination of leaf heat resistance: comparative investigation of chlorophyll fluorescence changes and tissue necrosis methods. Oecologia, 63, 256–262.CrossRefGoogle Scholar
Bilger, W. and Björkman, O. (1991). Temperature dependence of violaxanthin de-epoxidation and non-photochemical fluorescence quenching in intact leaves of Gossypium hirsutum L. and Malva parviflora L. Planta, 184, 226–234.CrossRefGoogle ScholarPubMed
Bilger, W., Björkmann, O. and Thayer, S.S. (1989). Light-induced spectral absorbance changes in relation to photosynthesis and the epoxidation state of xanthophyll cycle components in cotton leaves. Plant Physiology, 91, 542–551.CrossRefGoogle ScholarPubMed
Bilger, W., Veit, M., Schreiber, L., et al. (1997). Measurement of leaf epidermal transmittance of UV radiation by chlorophyll fluorescence. Physiologia Plantarum, 101, 754–763.CrossRefGoogle Scholar
Bilgin, D.D., Zavala, J.A., Zhu, J., et al. (2010). Biotic stress globally downregulates photosynthesis genes. Plant, Cell and Environment, 33, 1597–1613.CrossRefGoogle ScholarPubMed
Billesbach, D.P., Fischer, M.L., Torn, M.S., et al. (2004). A portable eddy covariance system for the measurement of ecosystem–atmosphere exchange of CO2, water vapor, and Energy. Journal of Atmospheric and Oceanic Technology, 21, 639–650.2.0.CO;2>CrossRefGoogle Scholar
Billings, W.D. (1973). Arctic and alpine vegetations: similarities, differences and susceptibility to disturbance. BioScience, 37, 58–67.Google Scholar
Billings, W.D. (1974). Adaptations and origins of alpine plants. Arctic and Alpine Research, 6, 129–142.CrossRefGoogle Scholar
Billings, W.D. and Mooney, H.A. (1968). The ecology of arctic and alpine plants. Biological Reviews, 481–529.CrossRefGoogle Scholar
Bird, A.F. (1974). Plant response to root-knot nematode. Annual Review of Phytopathology, 12, 69–85.CrossRefGoogle Scholar
Bird, I.F., Cornelius, M.J. and Keys, A.J. (1982). Affinity of RuBP carboxylases for carbon dioxide and inhibition of the enzymes by oxygen. Journal of Experimental Botany, 33, 1004–1013.CrossRefGoogle Scholar
Birkhold, K.T., Koch, K.E. and Darnell, R.L. (1992). Carbon and nitrogen economy of developing rabbiteye blueberry fruit. Journal of the American Society of Horticultural Sciences, 117, 139–145.Google Scholar
Birth, G.S. and McVey, G.R. (1968). Measuring the colour of growing turf with a reflectance spectrophotometer. Agronomy Journal, 60, 640–643.CrossRefGoogle Scholar
Biswal, B., Smith, A.J. and Rogers, L.J. (1994). Changes in carotenoids but not in D1 protein in response to nitrogen depletion and recovery in a cyanobacterium. FEMS Microbiology Letters, 116, 341–347.CrossRefGoogle Scholar
Biswal, U.C., Biswal, B. and Raval, M.K. (2003). Chloroplast Biogenesis: From Protoplastid to Gerontoplast, Kluwer Academic Publishers, Dordrecht, Netherlands.CrossRefGoogle Scholar
Björkman, O. (1971). Comparative photosynthetic CO2 exchange in higher plants. In: Photosynthesis and Photorespiration (eds Hatch, M.D., Osmond, C.B. and Slatyer, R.O.), Wiley Interscience, New York, USA. pp. 18–32.Google Scholar
Björkman, O. (1981). Responses to different quantum flux densities. In: Physiological Plant Ecology, vol I, Encyclopedia of Plant Physiology, 12A (eds Lange, O.L., Nobel, P.S., Osmond, C.B. and Ziegler, H.), Springer-Verlag, Berlin, pp. 57–107.Google Scholar
Björkman, O. (1994). Responses to long-term drought and high-irradiance stress in natural vegetation. Carnegie Institution of Washington Yearbook, 94, 62–63.Google Scholar
Björkman, O. and Demmig, B. (1987). Photon yield of O2 evolution and chlorophyll fluorescence characteristics at 77K among vascular plants of diverse origins. Planta, 170, 489–504.CrossRefGoogle Scholar
Björkman, O. and Demmig-Adams, B. (1994). Regulation of photosynthetic light energy capture, conversion and dissipation in leaves of higher plants. In: Ecophysiology of Photosynthesis (eds Schulze, E.-D. and Caldwell, M.M.), Springer-Verlag. Berlin. pp. 17–47.Google Scholar
Björkman, O., Badger, M.R. and Armond, P.A. (1980). Response and adaptation of photosynthesis to high temperature. In: Adaptations of Plants to Water and High Temperature Stress (eds Turner, N.C. and Kramer, P.J.), Wiley Interscience, NY, USA, pp. 223–249.Google Scholar
Björkman, O., Pearcy, R.W, Harrison, A.T., et al. (1972). Photosynthetic adaptation to high temperatures: a field study in Death Valley, California. Science, 175, 786–789.CrossRefGoogle ScholarPubMed
Black, C.C. (1973). Photosynthetic carbon fixation in relation to net CO2 uptake. Annual Review of Plant Physiology, 24, 253–286.CrossRefGoogle Scholar
Black, K., Davis, P., McGrath, J., et al. (2005). Interactive effects of irradiance and water availability on the photosynthetic performance of Picea sitchensis seedlings: implications for seedling establishment under different management practices. Annals of Forest Science, 62, 413–422.CrossRefGoogle Scholar
Black, T.A., Gaumont-Guay, D., Jassla, R., et al. (2005). Measurement of CO2 exchange between the boreal forest and the atmosphere. In: The Carbon Balance of Forest Biomes (eds Griffiths, H and Jarvis, P.G., Taylor & Francis, Oxon.Google Scholar
Blackburn, G.A. (2007). Hyperspectral remote sensing of plant pigments. Journal of Experimental Botany, 58, 855–867.CrossRefGoogle ScholarPubMed
Blanke, M.M. (2002). Photosynthesis of strawberry fruit. Acta Horticulturae, 567, 373–376.CrossRefGoogle Scholar
Blanke, M.M. and Cooke, D.T. (2004). Effects of flooding and drought on stomatal activity, transpiration, photosynthesis, water potential and water channel activity in strawberry stolons and leaves. Plant Growth Regulation, 42, 153–160.CrossRefGoogle Scholar
Blanke, M.M. and Lenz, F. (1989). Fruit photosynthesis. Plant, Cell and Environment, 12, 31–46.CrossRefGoogle Scholar
Blankenship, R.E. (1992). Origin and early evolution of photosynthesis. Photosynthesis Research, 33, 91–111.CrossRefGoogle ScholarPubMed
Blankenship, R.E., Tiede, D.M., Barber, J. et al. (2011). Comparing photosynthetic and photovoltaic efficiencies and recognizing the potential for improvement. Science, 332, 805–809.CrossRefGoogle Scholar
Bläsing, O.E., Ernst, K., Streubel, M., et al. (2002). The non-photosynthetic phosphoenolpyruvate carboxylases of the C4 dicot Flaveria trinervia: implications for the evolution of C4 photosynthesis. Planta, 215, 448–456.Google ScholarPubMed
Blasius, B., Neff, R., Beck, F., et al. (1999). Oscillatory model of crassulacean acid metabolism with a dynamic hysteresis switch. Proceedings of the Royal Society of London Series B, 266, 93–101.CrossRefGoogle Scholar
Bligny, R. and Douce, R. (2001). NMR and plant metabolism. Current Opinion in Plant Biology, 4, 191–196.CrossRefGoogle ScholarPubMed
Bliss, L.C. (1962). Adaptations of arctic and alpine plants to environmental conditions. Arctic, 15, 117–144.CrossRefGoogle Scholar
Blödner, C., Majcherczyk, A., Kües, U., et al. (2007). Early drought induced changes to the needle proteome of Norway spruce. Tree Physiology, 27, 1423–1431.CrossRefGoogle ScholarPubMed
Blom, C.W.P.M. and Voesenek, L.A.C.J. (1996). Flooding: the survival strategies of plants. Trends in Ecology and Evolution, 11, 290–295.CrossRefGoogle Scholar
Bloom, A.J., Smart, D.R., Nguyen, D.T., et al. (2002). Nitrogen assimilation and growth of wheat under elevated carbon dioxide. Proceedings of the National Academy of Sciences of the United States of America, 99, 1730–1735.CrossRefGoogle ScholarPubMed
Blum, A. (2005). Drought resistance, water-use efficiency, and yield potential: are they compatible, dissonant, or mutually exclusive?Australian Journal of Agricultural Research, 56, 1159–1168.CrossRefGoogle Scholar
Blum, A. (2009). Effective use of water (EUW) and not water-use efficiency (WUE) is the target of crop yield improvement under drought stress. Field Crops Research, 112, 119–123.CrossRefGoogle Scholar
Boardman, N.K. (1977). Comparative photosynthesis of sun and shade plants. Annual Review of Plant Physiology, 28, 355–377.CrossRefGoogle Scholar
Boccalandro, H.E., Rugnone, M.L., Moreno, J.E., et al. (2009). Phytochrome B enhances photosynthesis at the expense of water-use efficiency in Arabidopsis. Plant Physiology, 150, 1083–1092.CrossRefGoogle ScholarPubMed
Boccara, M., Boue, C., Garmier, M., et al. (2001). Infrared thermography revealed a role for mitochondria in pre-symptomatic cooling during harpin-induced hypersensitive response. Plant Journal, 28, 663–670.CrossRefGoogle ScholarPubMed
Bode, S., Quentmeier, C.C., Liao, P.-N., et al. (2009). On the regulation of photosynthesis by excitonic interactions between carotenoids and chlorophylls. Proceedings of the National Academy of Sciences USA, 106, 12311–12316.CrossRefGoogle ScholarPubMed
Bogeat-Triboulot, M.B, Brosché, M., Renaut, J., et al. (2007). Gradual soil water depletion results in reversible changes of gene expression, protein profiles, ecophysiology, and growth performance in Populus euphratica, a poplar growing in arid regions. Plant Physiology, 143, 876–892.CrossRefGoogle ScholarPubMed
Bolhàr-Nordenkampf, H.R. and Draxler, G. (1993). Functional leaf anatomy. In: Photosynthesis and Production in a Changing Environment (eds Hall, D.O., Scurlock, J.M.O., Bolhàr-Nordenkampf, H.R., Leegood, R.C. and Long, S. P.) Chapman and Hall, London, pp. 91–111.Google Scholar
Bolton, J.R. and Hall, D.O. (1991). The maximum efficiency of photosynthesis. Photochemistry and Photobiology, 53, 545–548.CrossRefGoogle Scholar
Bona, L., Carver, B.F., Wright, R.J., et al. (1994). Aluminum tolerance of segregating wheat populations in acidic soil and nutrient solutions. Communications in Soil Science and Plant Analysis, 25, 327–339.CrossRefGoogle Scholar
Bonan, G.B. (2008). Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science, 320, 1444–1449.CrossRefGoogle ScholarPubMed
Bonan, G.B., Davis, K.J., Baldocchi, D., et al. (1997). Comparison of the NCAR LSM1 land surface model with BOREAS aspen and jack pine tower fluxes. Journal of Geophysical Research-Atmospheres, 102, 29065–29075.CrossRefGoogle Scholar
Bonaventure, G., Gfeller, A., Proebsting, W.M., et al. (2007). A gain of function allele of TPC1 activates oxylipin biogenesis after leaf wounding in Arabidopsis. The Plant Journal, 49, 889–898.CrossRefGoogle ScholarPubMed
Bond, B.J. (2000). Age-related changes in photosynthesis of woody plants. Trends in Plant Science, 5, 349–353.CrossRefGoogle ScholarPubMed
Bond, B.J. and Ryan, M.G. (2000). Comment on ‘Hydraulic limitation of tree height: a critique’ by Becker, Meinzer and Wullschleger. Functional Ecology, 14, 137–140.CrossRefGoogle Scholar
Bond, B.J., Farnsworth, B.T., Coulombe, R.A., et al. (1999). Foliage physiology and biochemistry in response to light gradients in conifers with varying shade tolerance. Oecologia, 120, 183–192.CrossRefGoogle ScholarPubMed
Bond, W.J. (2005). Large parts of the world are brown or black: A different view on the ‘Green World’ hypothesis. Journal of Vegetation Science, 16, 261–266.Google Scholar
Bond, W.J. and Keeley, J.E. (2005). Fire as global ‘herbivore’: the ecology and evolution of flammable ecosystems. Trends in Ecology and Evolution, 20, 387–394.CrossRefGoogle ScholarPubMed
Bond, W.J. and Midgley, G.F. (2000). A proposed CO2-controlled mechanism of woody plant invasion in grasslands and savannas. Global Change Biology, 6, 865–870.CrossRefGoogle Scholar
Bond, W.J., Midgley, G.F. and Woodward, F.I. (2003). The importance of low atmospheric CO2 and fire in promoting the spread of grasslands and savannas. Global Change Biology, 9, 973–982.CrossRefGoogle Scholar
Bond, W.J., Woodward, F.I. and Midgley, G.F. (2005). The global distribution of ecosystems in a world without fire. New Phytologist, 165, 525–538.CrossRefGoogle Scholar
Bonfig, K.B., Schreiber, U., Gabler, A., et al. (2006). Infection with virulent and avirulent P. syringae strains differentially effects photosynthesis and sink metabolism in Arabidopsis leaves. Planta, 225, 1–12.CrossRefGoogle ScholarPubMed
Bongi, G. and Loreto, F. (1989). Gas-exchange properties of salt-stressed olive (Olea europaea L.) leaves. Plant Physiology, 90, 1408–1416.CrossRefGoogle ScholarPubMed
Bonhomme, L., Barbaroux, C. and Monclus, R. (2008) Genetic variation in productivity, leaf traits and carbon isotope discrimination in hybrid poplars cultivated on contrasting sites. Annals of Forest Science, 65, 1–9.Google Scholar
Bonhomme, R. (2000). Beware of comparing RUE values calculated from PAR vs solar radiation or absorbed vs intercepted radiation. Field Crops Research, 68, 247–252.CrossRefGoogle Scholar
Boote, K.J. and Pickering, N.B. (1994). Modeling photosynthesis of row crop canopies. HortScience, 29, 1423–1434.Google Scholar
Borchert, R. (1991). Growth periodicity and dormancy. In: Physiology of Trees (ed. Raghavendra, A.S.), John Wiley and Sons Inc., New York, USA, pp. 221–245.Google Scholar
Borel, C. and Simonneau, T. (2002). Is the ABA concentration in the sap collected by pressurizing leaves relevant for analysing drought effects on stomata? Evidence from ABA-fed leaves of transgenic plants with modified capacities to synthesize ABA. Journal of Experimental Botany, 53, 287–296.CrossRefGoogle ScholarPubMed
Borel, C., Frey, A., Marion-Poll, A., et al. (2001). Does engineering abscisic acid biosynthesis in Nicotiana plumbaginifolia modify stomatal response to drought?Plant Cell and Environment, 24, 477–489.CrossRefGoogle Scholar
Borisjuk, L., Walenta, S., Rolletschek, H., et al. (2002). Spatial analysis of plant metabolism: sucrose imaging within Vicia faba cotyledons reveals specific developmental patterns. The Plant Journal, 29, 521–530.CrossRefGoogle ScholarPubMed
Borland, A.M. and Griffiths, H. (1997). A comparative study on the regulation of C3 and C4 carboxylation processes in the constitutive crassulacean acid metabolism (CAM) plant Kalanchoë daigremontiana and the C3-CAM intermediate Clusia minor. Planta, 201, 368–378.CrossRefGoogle Scholar
Borland, A.M., Griffiths, H., Broadmeadow, M.S.J., et al. (1993). Short-term changes in carbon-isotope discrimination in the C3-CAM intermediate Clusia minor L. growing in Trinidad. Oecologia, 95, 444–453.CrossRefGoogle Scholar
Borland, A.M., Griffiths, H., Maxwell, C., et al. (1992). On the ecophysiology of the Clusiaceae in Trinidad: expression of CAM in Clusia minor L. during the transition from wet to dry season and characterization of three endemic species. New Phytologist, 122, 349–357.CrossRefGoogle Scholar
Borrás, L., Slafer, G.A. and Otegui, M.E. (2004). Seed dry weight response to source-sink manipulation in wheat, maize and soybean: a quantitative reappraisal. Field Crops Research, 86, 131–146.CrossRefGoogle Scholar
Bort, J., Brown, R.H. and Araus, J.L. (1996). Refixation of respiratory CO2 in the ears of C3-cereals. Journal of Experimental Botany, 47, 1567–1575.CrossRefGoogle Scholar
Bortier, K., Ceulemans, R. and De Temmerman, L. (2000). Effects of ozone exposure on growth and photosynthesis of beach seedlings (Fagus sylvatica). The New Phytologist, 146, 271–280.CrossRefGoogle Scholar
Bosian, G. (1960). Züm kuvettenklimaproblem: beweisführung für die nichtexistenz 2- gipfeliger assimilationskurven bei verwendung von klimatisierten. küvetten. Flora, 149, 167–188.Google Scholar
Boston, R.S., Viitanen, P.V. and Vierling, E. (1996). Molecular chaperones and protein folding in plants. Plant Molecular Biology, 32, 191–222.CrossRefGoogle ScholarPubMed
Bota, J., Flexas, J. and Medrano, H. (2001). Genetic variability of photosynthesis and water use in Balearic grapevine cultivars. Annals of Applied Biology, 138, 353–365.CrossRefGoogle Scholar
Bota, J., Medrano, H. and Flexas, J. (2004). Is photosynthesis limited by decreased Rubisco activity and RuBP content under progressive water stress?New Phytologist, 162, 671–681.CrossRefGoogle Scholar
Bottin, H. and Mathis, P. (1985). Interaction of plastocyanin with the photosystem I reaction center: a kinetic study by flash absorption spectroscopy. Biochemistry, 24, 6453–6460.CrossRefGoogle Scholar
Bottrill, D.E., Possingham, J.V. and Kriedemann, P.E. (1970). The effect of nutrient deficiencies on photosynthesis and respiration in spinach. Plant and Soil, 32, 424–438.CrossRefGoogle Scholar
Boucher, O., Jones, A. and Betts, R.A. (2009). Climate response to the physiological impact of carbon dioxide on plants in the Met Office Unified Model HadCM3. Climate Dynamics, 32, 237–249.CrossRefGoogle Scholar
Boucher, Y., Douady, C.J., Papke, R.T., et al. (2003). Lateral gene transfer and the origins of prokaryotic groups. Annual Review of Genetics, 37, 283–328.CrossRefGoogle ScholarPubMed
Bounoua, L., DeFries, R., Collatz, G.J., et al. (2002). Effects of land cover conversion on surface climate. Climatic Change, 52, 29–64.CrossRefGoogle Scholar
Bouriaud, O., Soudani, K. and Bréda, N. (2003). Leaf-area index from litter collection: impact of specific leaf area variability within a beach stand. Canadian Journal of Remote Sensing, 29, 371–380.CrossRefGoogle Scholar
Boussac, A., Maison-Peteri, B., Vernotte, C., et al. (1985). The charge accumulation in NaCl - washed and in Ca2+ - reactivated Photosystem-II particles. Biochimica et Biophysica Acta, 808, 225–230.CrossRefGoogle Scholar
Bowden, R.L. and Rouse, D.I. (1991). Effects of Verticillium dahliae on gas exchange of potato. Phytopathology, 81, 293–301.CrossRefGoogle Scholar
Bowden, R.L., Rouse, D.I. and Sharkey, T.D. (1990). Mechanism of photosynthesis decrease by Verticillium dahliae in potato. Plant Physiology, 94, 1048–1055.CrossRefGoogle ScholarPubMed
Bowes, J.M. and Crofts, A.R. (1980). Binary oscillations in the rate of reoxidation of the primary acceptor of photosystem II. Biochimica et Biophysica Acta, 590, 373–384.CrossRefGoogle ScholarPubMed
Bowie, M.R., Wand, S.J.E. and Esler, K.J. (2000). Seasonal gas exchange responses under three different temperature treatments in a leaf-succulent and a drought-deciduous shrub from the Succulent Karoo. South African Journal of Botany, 66, 118–123.CrossRefGoogle Scholar
Bowling, D.R., Sargent, S.D., Tanner, B.D., et al. (2003). Tunable diode laser absorption spectroscopy for stable isotope studies of ecosystem–atmosphere CO2 exchange. Agricultural and Forest Meteorology, 118, 1–19.CrossRefGoogle Scholar
Bowling, D.R., Tans, P.P. and Monson, R.K. (2001). Partitioning net ecosystem carbon exchange with isotopic fluxes of CO2. Global Change Biology, 7, 127–145.CrossRefGoogle Scholar
Bowman, W.D., Hubick, K.T., von Caemmerer, S., et al. (1989). Short-term changes in leaf carbon isotope discrimination in salt- and water-stressed C4 grasses. Plant Physiology, 90, 162–166.CrossRefGoogle Scholar
Bown, H.E., Watt, M.S., Clinton, P.W., et al. (2009). The influence of N and P supply and genotype on carbon flux and partitioning in potted Pinus radiata plants. Tree Physiology, 29, 1143–1151.CrossRefGoogle ScholarPubMed
Bowyer, W.J., Ning, L., Daley, L.S., et al. (1998). In vivo fluorescence imaging for detection of damage to leaves by fungal phytotoxins. Spectroscopy, 13, 36–44.Google Scholar
Box, E.O. (1981). Macroclimate and Plant Forms: An Introduction to Predictive Modelling in Phytogeography, Dr. W. Junk Publishers, The Hague – Boston – London.CrossRef
Boxall, S.F., Foster, J.M., Bohnert, H.J., et al. (2005). Conservation and divergence of circadian clock operation in a stress-inducible crassulacean acid metabolism species reveals clock compensation against stress. Plant Physiology, 137, 969–982.CrossRefGoogle Scholar
Boyce, R. and Lucero, S. (2002). Role of roots in winter water relations of Engelmann spruce saplings. Tree Physiology, 19, 893–898.CrossRefGoogle Scholar
Boyer, J.S., Wong, S.C. and Farquhar, G.D. (1997). CO2 and water vapour exchange across leaf cuticle (epidermis) at various water potentials. Plant Physiology, 114, 185–191.CrossRefGoogle Scholar
Bradford, K.J. (1983). Effects of flooding on leaf gas exchange of tomato plants. Plant Physiology, 73, 475–479.CrossRefGoogle ScholarPubMed
Bradford, M.M. (1976). A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Analytical Biochemistry, 72, 248–254.CrossRefGoogle ScholarPubMed
Brandon, P.C. (1967). Temperature features of enzymes affecting crassulacean acid metabolism. Plant Physiology, 42, 977–984.CrossRefGoogle ScholarPubMed
Brandt, S., Kehr, J., Walz, C., et al. (1999). A rapid method for detection of plant gene transcripts from single epidermal, mesophyll and companion cells of intact leaves. Plant Journal, 20, 245–250.CrossRefGoogle ScholarPubMed
Bravo, L.A. and Griffith, M. (2005). Characterization of antifreeze activity in Antarctic plants. Journal of Experimental Botany, 56, 1189–1196.CrossRefGoogle ScholarPubMed
Bravo, L.A., Saavedra-Mella, F.A., Vera, F., et al. (2007). Effect of cold acclimation on the photosynthetic performance of two ecotypes of Colobanthus quitensis (Kunth) Bartl. Journal of Experimental Botany, 58, 3581–3590.CrossRefGoogle ScholarPubMed
Bray, E.A. (2002). Abscisic acid regulation of gene expression during water-deficit stress in the era of the Arabidopsis genome. Plant Cell and the Environment, 25, 153–161.CrossRefGoogle ScholarPubMed
Bréhélin, C., Kessler, F. and van Wijk, K.J. (2007). Plastoglobules: versatile lipoprotein particles in plastids. Trends in Plant Sciences, 12, 260–166.CrossRefGoogle ScholarPubMed
Bremer, D.J., Auen, L.M., Ham, J.M., et al. (2001). Evapotranspiration in a prairie ecosystem: effects of grazing by cattle. Agronomy Journal, 93, 338–348.CrossRefGoogle Scholar
Brendel, O., Le Thiec, D., Scotti-Saintagne, C., et al. (2008). Quantitative trait loci controlling water use efficiency and related traits in Quercus robur L. Tree Genetics and Genomes, 4, 263–278.CrossRefGoogle Scholar
Breyton, C., Nandha, B, Johnson, G.N., Joliot, P. and Finazzi, G. (2006). Redox modulation of cyclic electron flow around photosystem I in C3 plants. Biochemistry, 45, 13465–13475.CrossRef
Briantais, J.M., Dacosta, J., Goulas, Y., et al. (1996). Heat stress induces in leaves an increase of the minimum level of chlorophyll fluorescence, F-0: a time-resolved analysis. Photosynthesis Research, 48, 189–196.CrossRefGoogle Scholar
Briantais, J.M., Vernotte, C., Picaud, M., et al. (1979). Quantitative study of the slow decline of chlorophyll alpha-fluorescence in isolated-chloroplasts. Biochim Biophys Acta, 548, 128–138.CrossRefGoogle Scholar
Briggs, L.J. and Shantz, H.J. (1914). Relative water requirements of plants. Journal Agriculture Research, 3, 1–63.Google Scholar
Briggs, L.J. and Shantz, HJL. (1913). The water requeriments of plants: a review of the literature. USDA Plant Industries Bulletin 285.
Briggs, W.R. (2005). Physiology of plant responses to artificial lighting. In: Ecological Consequences of Artificial Night Lighting (eds Rich, C. and Longcore, T.), Island Press, Washington, DC, USA, pp. 389–412.Google Scholar
Briggs, W.R. and Olney, M.A. (2001). Photoreceptors in plant photomorphogenesis to date: five phytochromes, two cryptochromes, one phototropin, and one superchrome. Plant Physiology, 125, 85–88.CrossRefGoogle ScholarPubMed
Bright, J., Desikan, R., Hancock, J.T., et al. (2006). ABA-induced NO generation and stomatal closure in Arabidopsis are dependent on H2O2 synthesis. Plant Journal, 45, 113–122.CrossRefGoogle ScholarPubMed
Brilli, F., Barta, C., Fortunati, A., et al. (2007). Response of isoprene emission and carbon metabolism to drought in white poplar (Populus alba) saplings. New Phytologist, 175, 244–254.CrossRefGoogle ScholarPubMed
Brodersen, C.R. and Vogelmann, T.C. (2007). Do epidermal lens cells facilitate the absorptance of diffuse light?American Journal of Botany, 94, 1061–1066.CrossRefGoogle ScholarPubMed
Brodersen, C.R. and Vogelmann, T.C. (2010). Do changes in light direction affect absorption profiles in leaves?Functional Plant Biology, 37, 403–412.CrossRefGoogle Scholar
Brodersen, C.R., Vogelmann, T.C., Williams, W.E., et al. (2008). A new paradigm in leaf-level photosynthesis: direct and diffuse lights are not equal. Plant Cell Environ, 31, 159–164.Google Scholar
Brodribb, T. and Hill, R.S. (1993). A physiological comparison of leaves and phyllodes in Acacia melanoxylon. Australian Journal of Botany, 41, 293–305.CrossRefGoogle Scholar
Brodribb, T. and Hill, R.S. (1999). The importance of xylem constraints in the distribution of coniferous plants. New Phytologist, 143, 365–372.CrossRefGoogle Scholar
Brodribb, T.J. and Cochard, H. (2009). Hydraulic failure defines the recovery and point of death in water-stressed conifers. Plant Physiology, 149, 575–584.CrossRefGoogle ScholarPubMed
Brodribb, T.J. and Holbrook, N.M. (2003). Stomatal closure during leaf dehydration, correlation with other leaf physiological traits. Plant Physiology, 132, 2166–2173.CrossRefGoogle ScholarPubMed
Brodribb, T.J. and Jordan, G.J. (2008). Internal coordination between hydraulics and stomatal control in leaves. Plant, Cell and Environment, 31, 1557–1564.CrossRefGoogle ScholarPubMed
Brodribb, T.J. and McAdam, S.A.M. (2011). Passive origins of stomatal control in vascular plants. Science, 331, 582–585.CrossRefGoogle ScholarPubMed
Brodribb, T.J., Field, T.S. and Jordan, G.J. (2007). Leaf maximum photosynthetic rate and venation are linked by hydraulics. Plant Physiology, 144, 1890–1898.CrossRefGoogle ScholarPubMed
Brodribb, T.J., McAdam, S.A.M., Jordan, G.J., et al. (2009). Evolution of stomatal responsiveness to CO2 and optimization of water-use efficiency among land plants. The New Phytologist, 183, 839–847.CrossRefGoogle Scholar
Brooks, A. (1986). Effects of phosphorus nutrition on ribulose 1,5-bisphosphate carboxylase activation, photosynthetic quantum yield and amounts of some Calvin cycle metabolites in spinach leaves. Australian Journal of Plant Physiology, 13, 221–237.CrossRefGoogle Scholar
Brooks, A. and Farquhar, G.D. (1985). Effect of temperature on the CO2/O2 specificity of ribulose-1,5-bisphosphate carboxylase/oxygenase and the rate of respiration in the light. Planta, 165, 397–406.CrossRefGoogle ScholarPubMed
Brooks, J.R., Flanagan, L.B., Buchmann, N., et al. (1997). Carbon isotope composition of Boreal plants: functional grouping of life forms. Oecologia, 110, 301–311.CrossRefGoogle ScholarPubMed
Brooks, J.R., Hinckley, T.M. and Sprugel, D.G. (1994). Acclimation responses of mature Abies amabilis sun foliage to shading. Oecologia, 100, 316–324.CrossRefGoogle ScholarPubMed
Brooks, J.R., Sprugel, D.G. and Hinckley, T.M. (1996). The effects of light acclimation during and after foliage expansion on photosynthesis of Abies amabilis foliage within the canopy. Oecologia, 107, 21–32.CrossRefGoogle Scholar
Brouwer, R. and de Wit, C.T. (1969). A simulation model of plant growth with special attention to root growth and its consequences. In: Root Growth (ed. Whittington, W.J.), Plenum, New York, USA, pp. 224–244.Google Scholar
Brown, K.R. and Courtin, P.J. (2003). Effects of phosphorus fertilization and liming on growth, mineral nutrition, and gas exchange of Alnus rubra seedlings grown in soils from mature alluvial Alnus stands. Canadian Journal of Forest Research-Revue Canadienne De Recherche Forestiere, 33, 2089–2096.CrossRefGoogle Scholar
Brown, R.A., Rosenberg, N.J., Hays, C.J., et al. (2000). Potential production and environmental effects of switchgrass and traditional crops under current and greenhouse-altered climate in the central United States: a simulation study. Agriculture, Ecosystems and Environment, 78, 31–47.CrossRefGoogle Scholar
Brown, R.H. (1978). A difference in N use efficiency in C3 and C4 plants and its implications in adaptation and evolution. Crop Science, 18, 93–98.CrossRefGoogle Scholar
Brown, R.H. and Simmons, R.E. (1979). Photosynthesis of grass species differing in CO2 fixation pathways. I. Water-use efficiency. Crop Science, 19, 375–379.CrossRefGoogle Scholar
Browse, J. and Xin, Z. (2001). Temperature sensing and cold acclimation. Current Opinion in Plant Biology, 4, 241–246.CrossRefGoogle ScholarPubMed
Brugnoli, E. and Farquhar, G.D. (2000). Photosynthetic fractionation of carbon isotopes. In: Photosynthesis Physiology and Metabolism, Advances in Photosynthesis (eds Leegood, R.C., Sharkey, T.D. and von Caemmerer, S.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 399–434.Google Scholar
Brugnoli, E., Hubick, K.T., von Caemmerer, S., et al. (1988). Correlation between the carbon isotope discrimination in leaf starch and sugars of C3 plants and the ratio of intercellular and atmospheric partial pressures of carbon dioxide. Plant Physiology, 88, 1418–1424.CrossRefGoogle Scholar
Brugnoli, E., Lauteri, M. and Guido, M.C. (1994). Carbon isotpe discrimination and photosynthesis: response and adaptation to environmental stress. In: Plant Sciences 1994 Second General Colloquium on Plant Sciences (eds de Kouchkovsky, Y. and Larher, F.). SFPV Université de Rennes, Rennes, France, pp. 269–272.Google Scholar
Brugnoli, E., Scartazza, A., Lauteri, M., et al. (1998). Carbon isotope discrimination in structural and non-structural carbohydrates in relation to productivity and adaptation to unfavourable conditions. In: Stable Isotopes: Integration of Biological, Ecological and Geochemical Processes (ed. Griffiths, H.), BIOS Scientific Publishers, Oxford, UK, pp. 133–144.Google Scholar
Bruinsma, J. (1961). A comment on the spectrophotometric determination of chlorophyll. Biochemistry and Biophysics Acta, 52, 576–578.CrossRefGoogle ScholarPubMed
Brunes, L., Öquist, G. and Eliasson, L. (1980). On the reason for the different photosynthetic rates of seedlings of Pinus sylvestris and Betula verrucosa. Plant Physiology, 66, 940–944.CrossRefGoogle Scholar
Bruni, N.C., Young, J.P. and Dengler, N.C. (1996). Leaf developmental plasticity of Ranunculus flabellaris in response to terrestrial and submerged environments. Canadian Journal of Botany, 74, 823–837.CrossRefGoogle Scholar
Bruzzese, B.M., Bowler, R., Massicotte, H.B. et al. (2010). Photosynthetic light response in three carnivorous plant species: Drosera rotundifolia, D. Capensis and Sarracenia leucophylla. Photosynthetica, 48, 103–109.CrossRefGoogle Scholar
Bryant, J.P., Chapin, F.S.I., Reichardt, P.B., et al. (1987). Response of winter chemical defense in Alaska paper birch and green alder to manipulation of carbon/nutrient balance. Oecologia, 72, 510–514.CrossRefGoogle ScholarPubMed
Buchanan, B.B (1984). The ferredoxin/thioredoxin system: a key element in the regulatory function of light in photosynthesis. BioScience, 34, 378–383.CrossRefGoogle ScholarPubMed
Buchanan, B.B. and Balmer, Y. (2005). Redox regulation: a broadening horizon. Annual Review Plant Biology, 56, 187–220.CrossRefGoogle ScholarPubMed
Buchanan-Bollig, I.C. and Kluge, M. (1981). Crassulacean acid metabolism (CAM) in Kalanchoë daigremontiana: temperature response of phosphoenolpyruvate (PEP)- carboxylase in relation to allosteric effectors. Planta, 152, 181–188.CrossRefGoogle ScholarPubMed
Buchanan-Bollig, I.C., Kluge, M. and Müller, D. (1984). Kinetic changes with temperature of phosphoenolpyruvate carboxylase from a CAM plant. Plant, Cell and Environment 7, 63–70.CrossRefGoogle Scholar
Buchmann, N., Brooks, J.R., Rapp, K.D., et al. (1996). Carbon isotope composition of C4 grasses is influenced by light and water supply. Plant, Cell and Environment, 19, 392–402.CrossRefGoogle Scholar
Buchner, O., Holzinger, A. and Lütz, C. (2007). Effects of temperature and light on the formation of chloroplast protrusions in leaf mesophyll cells of high alpine plants. Plant, Cell and Environment, 30, 1347–1356.CrossRefGoogle ScholarPubMed
Buckland, S.M., Grime, J.P., Hodgson, J.G., et al. (1997). A comparison of plant responses to the extreme drought of 1995 in northern England. Journal of Ecology 85, 875–882.CrossRefGoogle Scholar
Buckley, T.N. (2005). The control of stomata by water balance. New Phytologist, 168, 275–292.CrossRefGoogle ScholarPubMed
Buckley, T.N. and Farquhar, G.D. (2004). A new analytical model for whole leaf potential electron transport rate. Plant Cell and Environment, 27, 1487–1502.CrossRefGoogle Scholar
Buckley, T.N., Farquhar, G.D. and Mott, K.A. (1997). Qualitative effects of patchy stomatal conductance distribution features on gas-exchange calculations. Plant, Cell and Environment, 20, 867–880.CrossRefGoogle Scholar
Buckley, T.N., Mott, K.A. and Farquhar, G.D. (2003). A hydromechanical and biochemical model of stomatal conductance. Plant, Cell and Environment, 26, 1767–1786.CrossRefGoogle Scholar
Budde, R.J.A. and Randall, D.D. (1990). Pea leaf mitochondrial PDH complex is inactivated in vivo in a light-dependent manner. Proceedings of the National Academy of Sciences USA, 87, 673–676.CrossRefGoogle Scholar
Bugbee, B. (1992). Steady state canopy gas exchange: system design and operation. HortScience, 27, 770–776.Google ScholarPubMed
Bugbee, B.G. and Salisbury, F.B. (1988). Exploring the limits of crop productivity. I. Photosynthetic efficiency of wheat in high irradiance environments. Plant Physiolopgy, 88, 869–878.CrossRefGoogle ScholarPubMed
Buick, R. (2008). When did oxygenic photosynthesis evolve?Philosophical Transactions of the Royal Society B, 363, 2731–2743.CrossRefGoogle ScholarPubMed
Bukhov, N. and Carpentier, R. (2004). Alternative photosystem-I driven electron transport routes: mechanisms and functions. Photosynthesis Research, 82, 17–33.CrossRefGoogle ScholarPubMed
Bukhov, N.G., Wiese, C., Neimanis, S., et al. (1999). Heat sensitivity of chloroplasts and leaves: leakage of protons from thylakoids and reversible activation of cyclic electron transport. Photosynthesis Research, 59, 81–93.CrossRefGoogle Scholar
Bulman, P., Mather, D.E. and Smith, D.L. (1993). Genetic improvement of spring barley cultivars grown in eastern Canada from 1910 to 1988. Euphytica, 71, 35–48.CrossRefGoogle Scholar
Bunce, J.A. (2000). Responses of stomatal conductance to light, humidity and temperature in winter wheat and barley grown at three concentrations of carbon dioxide in the field. Global Change Biology, 6, 371–382.CrossRefGoogle Scholar
Bunce, J.A. (2001). Seasonal patterns of photosynthetic response and acclimation to elevated carbon dioxide in field-grown strawberry. Photosynthesis Research, 68, 237–245.CrossRefGoogle ScholarPubMed
Bunce, J.A. (2008). Acclimation of photosynthesis to temperature in Arabidopsis thaliana and Brassica oleracea. Photosynthetica, 46, 517–524.CrossRefGoogle Scholar
Bunce, J.A. (2009). Use of the response of photosynthesis to oxygen to estimate mesophyll conductance to carbon dioxide in water-stressed soybean leaves. Plant, Cell and Environment, 32, 875–881.CrossRefGoogle ScholarPubMed
Bunce, J.A. (2010). Variable responses of mesophyll conductance to substomatal carbon dioxide concentration in common bean and soybean. Photosynthetica, 48, 507–512.CrossRefGoogle Scholar
Bungard, R.A. (2004). Photosynthetic evolution in parasitic plants: insight from the chloroplast genome. BioEssays, 26, 235–247.CrossRefGoogle ScholarPubMed
Bungard, R.A., Ruban, A.V., Hibberd, J.M., et al. (1999). Unusual carotenoid composition and a new type of xanthophyll cycle in plants. Proceedings of the National Academy of Sciences USA, 96, 1135–1139.CrossRefGoogle Scholar
Bungard, R.A., Zipperlen, S.A., Press, M.C., et al. (2002). The influence of nutrients on growth and photosynthesis of seedlings of two rainforest dipterocarp species. Functional Plant Biology, 29, 501–515.CrossRefGoogle Scholar
Burghardt, M. and Riederer, M. (2003). Ecophysiological relevance of cuticular transpiration of deciduous and evergreen plants in relation to stomatal closure and leaf water potential. Journal of Experimental Botany, 54, 1941–1949.CrossRefGoogle ScholarPubMed
Burkart, S., Manderscheid, R. and Weigel, H.J. (2000). Interacting effects of photosynthetic photon-flux density and temperature on canopy CO2 exchange rate of spring wheat under different CO2-concentrations. Journal of Plant Physiology, 157, 31–39.CrossRefGoogle Scholar
Burkart, S., Manderscheid, R. and Weigel, H.J. (2004). Interactive effects of elevated atmospheric CO2 concentrations and plant available soil water content on canopy evapotranspiration and conductance of spring wheat. European Journal of Agronomy, 21, 401–417.CrossRefGoogle Scholar
Burkart, S., Manderscheid, R. and Weigel, H.J. (2007). Design and performance of a portable gas exchange chamber system for CO2- and H2O-flux measurements in crop canopies. Environmental Experimental Botany, 61, 25–34.CrossRefGoogle Scholar
Busch, F., Huner, N.P.A. and Ensminger, I. (2007). Increased air temperature during simulated autumn conditions does not increase photosynthetic carbon gain but affects the dissipation of excess energy in seedlings of the evergreen conifer jack pine. Plant Physiology, 143, 1242–1251.CrossRefGoogle Scholar
Busch, F., Hüner, N.P.A. and Ensminger, I. (2008). Increased air temperature during simulated autumn conditions impairs photosynthetic electron transport between photosystem II and photosystem I. Plant Physiology, 147, 402–414.CrossRefGoogle ScholarPubMed
Buschmann, C. (1999). Thermal dissipation during photosynthetic induction and subsequent dark recovery as measured by photoacoustic signals. Photosynthetica, 36, 149–161.CrossRefGoogle Scholar
Buschmann, C. and Lichtenthaler, H.K. (1998). Principles and characteristics of multi-colour fluorescence imaging of plants. Journal of Plant Physiology, 152, 297–314.CrossRefGoogle Scholar
Buschmann, C., Langsdorf, G. and Lichtenthaler, H.K. (2000). Imaging of the blue, green, and red fluorescence emission of plants: an overview. Photosynthetica, 38, 483–491.CrossRefGoogle Scholar
Businger, J.A. (1986). Evaluation of the accuracy with wich dry deposition can be measured with current micrometeorological techniques. Journal of Climate and Applied Meteorology, 25, 1100–1124.2.0.CO;2>CrossRefGoogle Scholar
Bussotti, F. and Ferretti, M. (1998). Air pollution, forest condition and forest decline in Southern Europe: an overview. Environmental Pollution, 101, 49–65.CrossRefGoogle ScholarPubMed
Bussotti, F., Desotgiu, R., Cascio, C., et al. (2007). Photosynthesis responses to ozone in young trees of three species with different sensitivities, in a 2-year open-top chamber experiment (Curno, Italy). Physiologia Plantarum, 130, 122–135.CrossRefGoogle Scholar
Butz, N.D. and Sharkey, T.D. (1989). Activity ratios of ribulose-1,5-bisphosphate carboxylase accurately reflect carbamylation ratios. Plant Physiology, 89, 735–739.CrossRefGoogle ScholarPubMed
Bykova, N.V., Keerberg, O., Pärnik, T., et al. (2005). Interaction between photorespiration and respiration in transgenic potato plants with antisense reduction in glycine decarboxylase. Planta, 222, 130–140.CrossRefGoogle ScholarPubMed
Byrd, G.T., Sage, R.F. and Brown, R.H. (1992). A comparison of dark respiration between C3 and C4 plants. Plant Physiology, 100, 191–198.CrossRefGoogle Scholar
Byrdwell, W.C. and Neff, W.E. (2002). Dual parallel electrospray ionization and atmospheric pressure chemical ionization mass spectrometry (MS), MS/MS and MS/MS/MS for the analysis of triacylglycerols and triacylglycerol oxidation products. Rapid Communications in Mass Spectrometry, 16, 300–319.CrossRefGoogle Scholar
Cabrera, H.M., Rada, F. and Cavieres, L. (1998). Effects of temperature on photosynthesis of two morphologically contrasting plant species along an altidudinal gradient in the tropical high Andes. Oecologia, 114, 145–152.CrossRefGoogle ScholarPubMed
Cai, H., Biswas, D.K., Shang, A.Q., et al. (2007). Photosynthetic response to water stress and changes in metabolites in Jasminum sambac. Photosynthetica, 45, 503–509.CrossRefGoogle Scholar
Cai, T., Flanagan, L.B., Jassal, R.S., et al. (2008). Modelling environmental controls on ecosystem photosynthesis and the carbon isotope composition of ecosystem-respired CO2 in a coastal Douglas fir forest. Plant, Cell and Environment, 31, 435–453.CrossRefGoogle Scholar
Cai, Y.F., Zhang, S.B., Hu, H., et al. (2010). Photosynthetic performance and acclimation of Incarvillea delavayi to water stress. Biologia Plantarum, 54, 89–96.CrossRefGoogle Scholar
Cai, Z.Q., Chen, Y.J., Guo, Y.H., et al. (2005). Responses of two field-grown coffee species to drought and tobacco. Photosynthetica, 43, 187–193.CrossRefGoogle Scholar
Caldwell, M.M., White, R.S., Moore, R.T., et al. (1977). Carbon balance, productivity, and water use of cold-winter desert shrub communities dominated by C3 and C4 species. Oecologia, 29, 275–300.CrossRefGoogle ScholarPubMed
Calfapietra, C., Mugnozza, G.S., Karnosky, D.F., et al. (2008). Isoprene emission rates under elevated CO2 and O3 in two field-grown aspen clones differing in their sensitivity to O3. New Phytologist, 179, 55–61.CrossRefGoogle ScholarPubMed
Calfapietra, C., Tulva, I., Eensalu, E., et al. (2005). Canopy profiles of photosynthetic parameters under elevated CO2 and N fertilization in a poplar plantation. Environmental Pollution, 137, 525–535.CrossRefGoogle Scholar
Calvin, M. (1962). The path of carbon in photosynthesis. Science, 135, 879–889.CrossRefGoogle ScholarPubMed
Calvin, M. and , A. A. (1948). The path of carbon in photosynthesis. Science, 107, 476–480.CrossRefGoogle ScholarPubMed
Calvin, M., Heidelberger, C., Reid, J.C., et al. (1949). Isotopic carbon techniques. In: Measurement and Chemical Manipulation, John Wiley and Sons, Inc, New York, USA.Google Scholar
Camenen, L., Goulas, Y., Guyot, G., et al. (1996) Estimation of the chlorophyll fluorescence lifetime of plant canopies: validation of a deconvolution method based on the use of a 3-D canopy mockup. Remote Sensing of the Environment, 57, 79–87.Google Scholar
Cameron, D.D., Geniez, J.M., Seel, W.E., et al. (2007). Suppression of host photosynthesis by the parasitic plant Rhinanthus minor. Annals of Botany, 101, 573–578.CrossRefGoogle Scholar
Campbell, G.S. (1986). Extinction coefficients for radiation in plant canopies calculated using an ellipsoidal inclination angle distribution. Agricultural and Forest Meteorology, 36, 317–321.CrossRefGoogle Scholar
Campbell, G.S. and Norman, J.M. (1998). An Introduction to Environmental Biophysics. Springer-Verlag, New York, USA.CrossRefGoogle Scholar
Campbell, W.J., Allen, L.H. Jr. and Bowes, G. (1990). Response of soybean canopy photosynthesis to CO2 concentration, light and temperature. Journal of Experimenal Botany, 41, 427–433.CrossRefGoogle Scholar
Canaani, O. and Havaux, M. (1990). Evidence for a biological role in photosynthesis for cytochrome b-559, a component of photosystem II reaction center. Proceedings of the National Academy of Sciences USA, 87, 9295–9299.CrossRefGoogle Scholar
Canadell, J.G. and Raupach, M.R. (2008). Managing forests for climate change mitigation. Science, 320, 1456–1457.CrossRefGoogle ScholarPubMed
Canadell, J.P., Le Quéré, C., Raupach, M.R., et al. (2007). Contributions to accelerating atmospheric CO2 growth from economic activity, carbon intensity, and efficiency of natural sinks. Proceedings of the National Academy of Sciences of the USA, 104, 18866–18870.CrossRefGoogle ScholarPubMed
Canham, C.D., Finzi, A.C., Pacala, S.W., et al. (1994). Causes and consequences of resource heterogeneity in forests: interspecific variation in light transmission by canopy trees. Canadian Journal of Forest Research, 24, 337–349.CrossRefGoogle Scholar
Cannani, O. and Malkin, S. (1984). Distribution of light excitation in an intact leaf between the two photosystems of photosynthesis: changes in absorption cross-section following state 1 and state 2 transitions. Biochimica et Biophysica Acta, 766, 513–524.CrossRefGoogle Scholar
Cannell, M.G.R. and Thornley, J.H.M. (1998). Temperature and CO2 responses of leaf and canopy photosynthesis: a clarification using the non-rectangular hyperbola model of photosynthesis. Annals of Botany, 82, 883–892.CrossRefGoogle Scholar
Canvin, D.T., Berry, J.A., Badger, M.R., et al. (1980). Oxygen exchange in leaves in the light. Plant Physiology, 66, 302–307.CrossRefGoogle Scholar
Cao, J., Hesketh, J.D., Zur, B., et al. (1988). Leaf area development in maize and soybean plants. Biotronics, 17, 9–15.Google Scholar
Cao, K.F. and Ohkubo, T. (1998). Allometry, root/shoot ratio and root architecture in understory saplings of deciduous dicotyledonous trees in central Japan. Ecological Research, 13, 217–227.CrossRefGoogle Scholar
Cao, L., Bala, G., Caldeira, K., et al. (2010). Importance of carbon dioxide physiological forcing to future climate change. Proceedings of the National Academy of Sciences USA, 107, 9513–9518.CrossRefGoogle ScholarPubMed
Cape, J.N. (2003). Effects of airborne volatile organic compounds on plants. Environmental Pollution, 122, 145–157.CrossRefGoogle ScholarPubMed
Capell, B. and Dorffling, K. (1993). Genotype-specific differences in chilling tolerance of maize in relation to chilling-induced changes in water status and abscisic-acid accumulation. Physiologia Plantarum, 88, 638–646.CrossRefGoogle Scholar
Caporn, S.J.M. (1989). The effects of oxides of nitrogen and carbon dioxide enrichment on photosynthesis and growth of lettuce (Lactuca sativa L.). New Phytologist, 111, 473–481.CrossRefGoogle Scholar
Caporn, S.J.M., Risager, M. and Lee, J.A. (1994). Effect of atmospheric nitrogen deposition on frost hardiness in Calluna vulgaris. New Phytologist, 128, 461–468.CrossRefGoogle Scholar
Carlberg, I., Hansson, M., Kieselbach, T., et al. (2003). A novel plant protein undergoing light-induced phosphorylation and release from the photosynthetic thylakoid membranes. Proceedings of the National Academy of Sciences USA, 100, 757–762.CrossRefGoogle ScholarPubMed
Carlson, R.W., Bazzaz, F.A. and Rolfe, G.L. (1975). The effect of heavy metals on plants. II. Net photosynthesis and transpiration of whole corn and sunflower plants treated with Pb, Cd, Ni, and Tl. Environmental Research, 10, 113–120.CrossRefGoogle Scholar
Carmo-Silva, A.E., da Silva, A.B., Keys, A.J., et al. (2008b). The activities of PEP carboxylase and the C4 acid decarboxylases are little changed by drought stress in three C4 grasses of different subtypes. Photosynthesis Research, 97, 223–233.CrossRefGoogle ScholarPubMed
Carmo-Silva, A.E., Powers, S.J., Keys, A.J., et al. (2008a). Photorespiration in C4 grasses remains slow under drought conditions. Plant, Cell and Environment, 31, 925–940.CrossRefGoogle ScholarPubMed
Carnicer, J., Coll, M., Ninyerola, M. et al. (2011). Widespread crown condition decline, food web disruption, and amplified tree mortality with increased climate change-type drought. Proceedings of the National Academy of Sciences USA, 108, 1474–1478.CrossRefGoogle ScholarPubMed
Carol, P. and Kuntz, M. (2001). A plastid terminal oxidase comes to light: implications for carotenoid biosynthesis and chlororespiration. Trends in Plant Science, 6, 31–36.CrossRefGoogle ScholarPubMed
Carpenter, K.J. (2005). Stomatal architecture and evolution in basal angiosperms. American Journal of Botany, 92, 1595–1615.CrossRefGoogle ScholarPubMed
Carpenter, K.J. (2006). Specialized structures in the leaf of basal angiosperms: morphology, distribution, and homology. American Journal of Botany, 93, 665–681.CrossRefGoogle ScholarPubMed
Carpenter, R.J., Jordan, G.J., Leigh, A., et al. (2007). Giant cuticle pores in Eidothea zoexylocarya (Proteaceae) leaves. American Journal of Botany, 94, 1282–1288.CrossRefGoogle ScholarPubMed
Carrara, S., Pardosi, A., Soldatini, G.F., et al. (2001). Photosynthetic activity of ripening tomato fruit. Photosynthetica, 39, 75–78.CrossRefGoogle Scholar
Carrari, F., Nunes-Nesi, A., Gibon, Y., et al. (2003). Reduced expression of aconitase results in an enhanced rate of photosynthesis and ed leaves of wild species tomato. Plant Physiology, 133, 1322–1335.CrossRefGoogle Scholar
Carroll, T.W. (1970). Relation of barley stripe mosaic virus to plastids. Virology, 42, 1015–1022.CrossRefGoogle ScholarPubMed
Carroll, T.W. and Kosuge, T. (1969). Changes in structure of chloroplasts accompanying necrosis of tobacco leaves systemically infected with tobacco mosaic virus. Phytopathology, 59, 953–962.Google Scholar
Carswell, F.E., Whitehead, D., Rogers, G.N.D., et al. (2005). Plasticity in photosynthetic response to nutrient supply of seedlings from a mixed conifer-angiosperm forest. Austral Ecology, 30, 426–434.CrossRefGoogle Scholar
Cartelat, A., Cerovic, Z., Goulas, Y., et al. (2005). Optically assessed polyphenolics and chlorophyll content of leaves as indicators of nitrogen deficiency in wheat. Field Crops Research, 91, 35–49.CrossRefGoogle Scholar
Carter, G.A., Jones, J.H., Mitchell, R.J., et al. (1996). Detection of solar-excited chlorophyll a fluorescence and leaf photosynthetic capacity using a Fraunhofer line radiometer. Remote Sensing of the Environment, 55, 89–92.CrossRefGoogle Scholar
Carter, G.A., Theisen, A.F. and Mitchell, R.J. (1990). Chlorophyll fluorescence measured using the Fraunhofer line-depth principle and relationship to photosynthetic rate in the field. Plant Cell Environment, 13, 79–83.CrossRefGoogle Scholar
Carter, P.J., Wilkins, M.B., Nimmo, H.G., et al. (1995). Effects of temperature on the activity of phosphoenolpyruvate carboxylase and on the control of CO2 fixation in Bryophyllum fedtschenkoi. Planta, 196, 375–380.Google Scholar
Casella, E. and Sinoquet, H. (2003). A method for describing the canopy architecture of coppice poplar with allometric relationships. Tree Physiology, 23, 1153–1170.CrossRefGoogle ScholarPubMed
Caspar, T., Huber, S.C. and Somerville, C. (1985). Alterations in growth, photosynthesis, and respiration in a starchless mutant of Arabidopsis thaliana (L.) deficient in chloroplast phosphoglucomutase activity. Plant Physiology, 79, 11–17.CrossRefGoogle Scholar
Cassman, K.G., Dobermann, A., Walters, D.T., et al. (2003). Meeting cereal demand while protecting natural resources and improving environmental quality. Annual Review of Environment and Resources, 28, 315–358.CrossRefGoogle Scholar
Castrillo, M. (1995). Ribulose-1,5-bisphosphate carboxylase activity in altitudinal populations of Espeletia schultzii Wedd. Oecologia, 101, 193–196.CrossRefGoogle Scholar
Castro, A.J., Carapito, C., Zorn, N., et al. (2005). Proteomic analysis of grapevine (Vitis vinifera L.) tissues subjected to herbicide stress. Journal of Experimental Botany, 56, 2783–2795.CrossRefGoogle ScholarPubMed
Caswell, H. and Reed, F.C. (1975). Indigestibility of C4 bundle sheath cells by the grasshopper Melanoplus confusus. Annals of the Entomological Society of America, 68, 686–688.CrossRefGoogle Scholar
Caswell, H. and Reed, F.C. (1976). Plant-herbivore interactions: the indigestibilty of C4 bundle sheath cells by grasshoppers. Oecologia, 26, 151–156.CrossRefGoogle ScholarPubMed
Caswell, H., Reed, F., Stephenson, S.N., et al. (1973). Photosynthetic pathways and selective herbivory: a hypothesis. American Naturalist, 107, 465–480.CrossRefGoogle Scholar
Causin, H.B., Tremmel, D.C., Rufty, T.W., et al. (2004). Growth, nitrogen uptake, and metabolism in two semiarid shrubs grown at ambient and elevated atmospheric CO2 concentrations: effects of nitrogen supply and source. American Journal of Botany, 91, 565–572.CrossRefGoogle ScholarPubMed
Cavender-Bares, J., Apostol, S., Moya, I., et al. (1999). Chilling-induced photoinhibition in two oak species: are evergreen leaves better protected than deciduous leaves?Photosynthetica, 36, 587–596.CrossRefGoogle Scholar
Cechin, I. and Press, M.C. (1993). Influence of nitrogen on growth and photosynthesis of a C3 cereal, Oryza sativa, infected with the root hemiparasite Striga hermonthica. Journal of Experimental Botany, 45, 925–930.CrossRefGoogle Scholar
Cen, Y.P. and Sage, R.F. (2005). The regulation of Rubisco activity in response to variation in temperature and atmospheric CO2 partial pressure in sweet potato. Plant Physiology, 139, 979–990.CrossRefGoogle ScholarPubMed
Cen, Y.P., Turpin, D.H. and Layzell, D.B. (2001). Whole-plant gas exchange and reductive biosynthesis in white lupin. Plant Physiology, 126, 1555–1565.CrossRefGoogle ScholarPubMed
Centritto, M., Lauteri, M., Monteverdi, M.C., et al. (2009). Leaf gas exchange, carbon isotope discrimination, and grain yield in contrasting rice genotypes subjected to water deficits during the reproductive stress. Journal of Experimental Botany, 60, 2325–2339.CrossRefGoogle Scholar
Centritto, M., Lee, H.S.J. and Jarvis, P.G. (1999a). Increased growth in elevated [CO2]: an early, short-term response?Global Change Biology, 5, 623–633.CrossRefGoogle Scholar
Centritto, M., Lee, H.S.J. and Jarvis, P.G. (1999b). Interactive effects of elevated CO2 and drought on cherry (Prunus avium) seedlings: I. Growth, wholeplant water use efficiency and water loss. New Phytologist, 141, 129–140.CrossRefGoogle Scholar
Centritto, M., Loreto, F. and Chartzoulakis, K. (2003). The use of low [CO2] to estimate diffusional and non-diffusional limitations of photosynthetic capacity of salt-stressed olive saplings. Plant, Cell and Environment, 26, 585–594.CrossRefGoogle Scholar
Centritto, M., Lucas, M.E. and Jarvis, P.G. (2002). Gas exchange, biomass, whole-plant water-use efficiency and water uptake of peach (Prunus persica) seedlings in response to elevated carbon dioxide concentration and water availability. Tree physiology, 22, 699–706.CrossRefGoogle ScholarPubMed
Centritto, M., Magnani, F., Lee, H.S.J., et al. (1999c). Interactive effects of elevated [CO2] and drought on cherry (Prunus avium) seedlings. II. Photosynthetic capacity and water relations. New Phytologist, 141, 141–153.CrossRefGoogle Scholar
Centritto, M., Nascetti, P., Petrilli, L., et al. (2004). Profiles of isoprene emission and photosynthetic parameters in hybrid poplars exposed to free-air CO2 enrichment. Plant, Cell and Environment, 27, 403–412.CrossRefGoogle Scholar
Cerasoli, S., McGuire, M.A., Faria, J., et al. (2009). CO2 efflux, CO2 concentration and photosynthetic refixation in stems of Eucalyptus globulus (Labill.). Journal of Experimental Botany, 60, 99–105.CrossRefGoogle Scholar
Cerling, T.E. (1999). Paleorecords of C4 plants and ecosystems. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, CA, USA, pp. 445–469.Google Scholar
Cerling, T.E., Ehleringer, J.R. and Harris, J.M. (1998). Carbon dioxide starvation, the development of C4 ecosystems, and mammalian evolution. Philosophical Transactions of the Royal Society of London, Series B., 353, 159–170.CrossRefGoogle ScholarPubMed
Cerling, T.E., Harris, J.M.MacFadden, B.J., et al. (1997). Global vegetation change through the Miocene/Pliocene boundary. Nature, 389, 153–158.CrossRefGoogle Scholar
Cerling, T.E., Wang, Y. and Quade, J. (1993). Expansion of C4 ecosystems as an indicator of global ecological change in the late Miocene. Nature, 361, 344–345.CrossRefGoogle Scholar
Cernusak, L.A. and Hutley, L.B. (2011). Stable isotopes reveal the contribution of corticular photosynthesis to growth in branches of Eucalyptus miniata. Plant Physiology, 155, 515–523.CrossRefGoogle ScholarPubMed
Cernusak, L.A. and Marshall, J.D. (2000). Photosynthetic refixation in branches of western white pine. Functional Ecology, 14, 300–311.Google Scholar
Cernusak, L.A., Arthur, D.J., Pate, J.S., et al. (2003). Water relations link carbon and oxygen isotope discrimination to phloem sap sugars concentration in Eucalyptus globulus. Plant Physiology, 134, 1544–1554.CrossRefGoogle Scholar
Cernusak, L.A., Farquhar, G.D. and Pate, J.S. (2005). Environmental and physiological controls over the oxygen and carbon isotope composition of the Tasmanian blue gum, Eucalyptus globulus. Tree Physiology, 25, 129–146.CrossRefGoogle ScholarPubMed
Cernusak, L.A., Pate, J.S. and Farquhar, G.D. (2002). Diurnal variation in the stable isotope composition of water and dry matterin fruiting Lupinus angustifolius under field conditions. Plant, Cell and the Environment, 25, 893–907.CrossRefGoogle Scholar
Cernusak, L.A., Winter, K., Aranda, J., et al. (2008). Conifers, angiosperm trees, and lianas: growth, whole-plant water and nitrogen use efficiency, and stable isotope composition (δ13C and δ18O) of seedlings grown in a tropical environment. Plant Physiology, 148, 642–659.CrossRefGoogle Scholar
Cerovic, Z.G., Bergher, M., Goulas, Y., et al. (1993). Simultaneous measurement of changes in red and blue fluorescence in illuminated isolated chloroplasts and leaf pieces: the contribution of NADPH to the blue fluorescence signal. Photosynthesis Research, 36, 193–204.CrossRefGoogle ScholarPubMed
Cerovic, Z.G., Goulas, Y., Gorbunov, M., et al. (1996). Fluorosensing of water stress in plants. Diurnal changes of the mean lifetime and yield of chlorophyll fluorescence, measured simultaneously and at distance with a t-LIDAR and a modified PAM-fluorimeter, in maize, sugar beet and Kalanchoë J. Remote Sensing of Environment, 58, 311–321.CrossRefGoogle Scholar
Cerovic, Z.G., Langrand, E., Latouche, G., et al. (1998). Spectral characterization of NAD(P)H fluorescence in intact isolated chloroplasts and leaves: effect of chlorophyll concentration on reabsorption of blue-green fluorescence. Photosynthesis Research, 56, 291–301.CrossRefGoogle Scholar
Cerovic, Z.G., Ounis, A., Cartelat, A., et al. (2002). The use of chlorophyll fluorescence excitation spectra for the non-destructive in situ assessment of UV-absorbing compounds in leaves. Plant Cell Environment, 25, 1663–1676.CrossRefGoogle Scholar
Cerovic, Z.G., Samson, G., Morales, F., et al. (1999). Ultraviolet-induced fluorescence for plant monitoring: present state and prospects. Agronomie, 19, 543–578.CrossRefGoogle Scholar
Cescatti, A. (1998). Effects of needle clumping in shoots and crowns on the radiative regime of a Norway spruce canopy. Annales des Sciences Forestieres, 55, 89–102.CrossRefGoogle Scholar
Cescatti, A. and Niinemets, Ü. (2004). Sunlight capture. Leaf to landscape. In: Photosynthetic Adaptation: Chloroplast to Landscape (eds Smith, W. K., Vogelmann, T. C. and Chritchley, C.), Ecological Studies, 178, 42–85, Springer Verlag, Berlin.CrossRefGoogle Scholar
Cescatti, A. and Zorer, R. (2003). Structural acclimation and radiation regime of silver fir (Abies alba Mill.) shoots along a light gradient. Plant, Cell and Environment, 26, 429–442.CrossRefGoogle Scholar
Ceulemans, R. and Mousseau, M. (1994). Effects of elevated atmospheric CO2 on woody plants. Tansley Review No. 71. New Phytologist, 127, 425–446.CrossRefGoogle Scholar
Ceulemans, R., Janssens, I.A. and Jach, M.E. (1999). Effects of CO2 enrichment on trees and forests: lessons to be learned in view of future ecosystem studies. Annals of Botany, 84, 577–590.CrossRefGoogle Scholar
Chabot, B.F. and Hicks, D.J. (1982). The ecology of leaf lifespans. Annual Review of Ecology and Systematics, 13, 229–259.CrossRefGoogle Scholar
Chaerle, L. and Van Der Straeten, D. (2000). Imaging techniques and the early detection of plant stress. Trends in Plant Science, 5, 495–501.CrossRefGoogle ScholarPubMed
Chaerle, L. and Van Der Straeten, D. (2001). Seeing is believing: imaging techniques to monitor plant health. Biochimica et Biophysica Acta, 1519, 153–166.CrossRefGoogle ScholarPubMed
Chaerle, L., Hagenbeek, D., De Bruyne, E., et al. (2004). Thermal and chlorophyll-fluorescence imaging distinguish plant-pathogen interactions at an early stage. Plant Cell Physiology, 45, 887–896.CrossRefGoogle ScholarPubMed
Chaerle, L., Hagenbeek, D., De Bruyne, E., et al. (2007b). Chlorophyll fluorescence imaging for disease-resistance screening of sugar beet. Plant Cell Tissue and Organ Culture, 91, 97–106.CrossRefGoogle Scholar
Chaerle, L., Leinonen, I., Jones, H.G., et al. (2007a). Monitoring and screening plant populations with combined thermal and chlorophyll fluorescence imaging. Journal of Experimental Botany, 58, 773–784.CrossRefGoogle ScholarPubMed
Chaerle, L., Lenk, S., Buschmann, C., et al. (2007c). Multicolour fluorescence imaging for early detection of the hypersensitive reaction to tobacco mosaic virus. Journal of Plant Physiology, 164, 253–262.CrossRefGoogle Scholar
Chaerle, L., Pineda, M., Romero-Aranda, R., et al. (2006). Robotized thermal and chlorophyll fluorescence imaging of pepper mild mottle virus infection in Nicotiana benthamiana. Plant Cell Physiology, 47, 1323–1336.CrossRefGoogle ScholarPubMed
Chaerle, L., Van Caeneghem, W., Messens, E., et al.(1999). Presymptomatic visualization of plant-virus interactions by thermography. Nature Biotechnology, 17, 813–816.CrossRefGoogle ScholarPubMed
Chai, T.T., Simmonds, D., Day, D.A., et al. (2010). Photosynthetic performance and fertility are repressed in GmAOX2b antisense soybean. Plant Physiology, 152, 1638–1649.CrossRefGoogle ScholarPubMed
Chalker-Scott, L. (1999). Environmental significance of anthocyanins in plant stress responses. Photochemistry and Photobiology, 70, 1–9.CrossRefGoogle Scholar
Chameides, W., Lindsay, R., Richardson, J., et al. (1988). The role of biogenic hydrocarbons in urban photochemical smog: Atlanta as a case study. Science, 241, 1473–1475.CrossRefGoogle ScholarPubMed
Chameides, W.L., Fehsenfeld, F., Rodgers, M.O., et al. (1992). Ozone precursor relationships in the ambient atmosphere. Journal of Geophysical Research, 97, 6037–6055.CrossRefGoogle Scholar
Chandler, J.W. and Dale, J.E. (1993). Photosynthesis and nutrient supply in needles of Sitka spruce [Picea sitchensis (Bong.) Carr.]. The New Phytologist, 125, 101–111.CrossRefGoogle Scholar
Chaney, R.L. (1993). Zinc phytotoxicity. In: Zinc in Soil and Plants (ed. Robson, A.D.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 131–150.Google Scholar
Changnon, S.A., Kunkel, K.E. and Winstanley, D. (2002). Climate factors that caused the unique tall grass prairie in the central United States. Physical Geography, 23, 259–280.CrossRefGoogle Scholar
Chaoui, A., Mazhoudi, S., Ghorbal, M.H., et al. (1997). Cadmium and zinc induction of lipid peroxidation and effects on antioxidant enzyme activities in bean (Phaseolus vulgaris L.). Plant Science, 127, 139–147.CrossRefGoogle Scholar
Chapin, F.S. III, Bloom, A.J., Field, C.B., et al. (1987). Plant responses to multiple environmental factors. BioScience, 37, 49–57.CrossRefGoogle Scholar
Chapin, F.S., Matson, P.A. and Mooney, H.A. (2002). Principles of Terrestrial Ecosystem Ecology. Springer-Verlag, New York, USA.Google Scholar
Chappelle, E.W., Wood, F.M., McMurtrey, J.E., et al. (1984). Laser-induced fluorescence of green plants. 1. A technique for the remote detection of plant stress and species differentiation. Applied Optics, 23, 134–138.CrossRefGoogle ScholarPubMed
Charlesworth, B., Morgan, M.T. and Charlesworth, D. (1993). The effects of deleterious mutations on neutral molecular variation. Genetics, 134, 1289–1303.Google Scholar
Charlesworth, D. and Wright, S. (2001). Breeding systems and genome evolution. Current Opinions in Genetics and Development, 11, 685–690.CrossRefGoogle ScholarPubMed
Chaumont, M., Morot-Gaudry, J.F. and Foyer, C.H. (1995). Effects of photoinhibitory treatment on CO2 assimilation, the quantum yield of CO2 assimilation, D-1 protein, ascorbate, glutathione and xanthophyll contents and the electron transport rate in vine leaves. Plant Cell Environment, 18, 1358–1366.CrossRefGoogle Scholar
Chave, J., Coomes, D., Jansen, S., et al. (2009) Towards a worldwide wood economics spectrum. Ecology Letters, 12, 351–366.CrossRefGoogle ScholarPubMed
Chaves, M.M. (1991). Effects of water deficits on carbon assimilation. Journal of Experimental Botany, 42, 1–16.CrossRefGoogle Scholar
Chaves, M.M. and Oliveira, M.M. (2004). Mechanisms underlying plant resilience to water deficits: prospects for water-saving agriculture. Journal of Experimental Botany, 55, 2365–2384.CrossRefGoogle ScholarPubMed
Chaves, M.M., Flexas, J. and Pinheiro, C. (2009). Photosynthesis under drought and salt stress: regulation mechanisms from whole plant to cell. Annals of Botany, 103, 551–560.CrossRefGoogle ScholarPubMed
Chaves, M.M., Maroco, J.P. and Pereira, J.S. (2003). Understanding plant responses to drought: from genes to the whole plant. Functional Plant Biology, 30, 239–264.CrossRefGoogle Scholar
Chaves, M.M., Osório, J. and Pereira, J.S. (2004). Water use efficiency and photosynthesis. In: Water Use Efficiency in Plant Biology (ed. Bacon, M.A.) Blackwell Publishing, Oxford, UK, pp. 42–74.Google Scholar
Chaves, M.M., Pereira, J.S., Maroco, J., et al. (2002). How plants cope with water stress in the field. Photosynthesis and growth. Annals of Botany, 89, 1–10.CrossRefGoogle Scholar
Chaw, S.M., Parkinson, C.L., Cheng, Y., et al. (2000). Seed plant phylogeny inferred from all three plant genomes: monophyly of extant gymnosperms and origin of Gnetales from conifers. Proceedings of the National Academy of Sciences USA, 97, 4086–4091.CrossRefGoogle ScholarPubMed
Chazdon, R.L. (1985). Leaf display, canopy structure and ligth interception of two understory palm species. American Journal of Botany, 72, 1493–1502.CrossRefGoogle Scholar
Chazdon, R.L. (1988). Sunflecks and their importance to forest understory plants. Advances in Ecological Research, 18, 1–63.CrossRefGoogle Scholar
Chazdon, R.L. (1992). Photosynthetic plasticity of two rain forest shrubs across natural gap transects. Oecologia, 92, 586–595.CrossRefGoogle ScholarPubMed
Chazdon, R.L. and Fetcher, N. (1984). Photosynthetic light environments in a lowland tropical rainforest in Costa Rica. Journal of Ecology, 72, 553–564.CrossRefGoogle Scholar
Chazdon, R.L. and Field, C.B. (1987). Determinants of photosynthetic capacity in six rainforest Piper species. Oecologia, 73, 222–230.CrossRefGoogle ScholarPubMed
Chazdon, R.L. and Pearcy, R.W. (1986a). Photosynthetic Responses to Light Variation in Rain-Forest Species. I. Induction under constant and fluctuating light conditions. Oecologia, 69, 517–523.CrossRefGoogle ScholarPubMed
Chazdon, R.L. and Pearcy, R.W. (1986b). Photosynthetic responses to light variation in rain forest species. II. Carbon gain and photosynthetic efficiency during lightflecks. Oecologia, 69, 524–531.CrossRefGoogle ScholarPubMed
Chazdon, R.L., Williams, K. and Field, C.B. (1988). Interactions between crown structure and light environment in five rain forest Piper species. American Journal of Botany, 75, 1459–1471.CrossRefGoogle Scholar
Cheeseman, J.M. (1991). PATCHY: simulating and visualizing the effects of stomatal patchiness on photosynthetic CO2 exchange studies. Plant, Cell and Environment, 14, 593–599.CrossRefGoogle Scholar
Cheeseman, J.M. (2006). Hydrogen peroxide concentrations in leaves under natural conditions. Journal of Experimental Botany, 57, 2435–2444.CrossRefGoogle ScholarPubMed
Cheeseman, J.M. (2009). Seasonal patterns of leaf H2O2 content: reflections of leaf phenology, or environmental stress?Functional Plant Biology, 36, 721–731.CrossRefGoogle Scholar
Cheeseman, J.M. and Lovelock, C.E. (2004). Photosynthetic characteristics of dwarf and fringe Rhizophora mangle L. in a Belizean mangrove. Plant Cell and Environment, 27, 769–780.CrossRefGoogle Scholar
Cheeseman, J.M., Clough, B.F., Carter, D.R., et al. (1991). The analysis of photosynthetic performance in leaves under field conditions: a case study using Bruguiera mangroves. Photosynthesis Research, 29, 11–22.Google ScholarPubMed
Cheeseman, J.M., Herendeen, L.B., Cheeseman, A.T., et al. (1997). Photosynthesis and photoprotection in mangroves under field conditions. Plant, Cell and Environment, 20, 579–588.CrossRefGoogle Scholar
Chelle, M. (2005). Phylloclimate or the climate perceived by individual plant organs: What is it? How to model it? What for?New Phytologist, 166, 781–790.CrossRefGoogle ScholarPubMed
Chen, C.P., Zhu, X.G. and Long, S.P. (2008). The effect of leaf-level spatial variability in photosynthetic capacity on biochemical parameter estimates using the Farquhar model: A theoretical analysis. Plant Physiology, 148, 1139–1147.CrossRefGoogle ScholarPubMed
Chen, L., Fuchigami, L.H. and Breen, P.J. (2001). The relationship between photosystem II efficiency and quantum yield for CO2 assimilation is not affected by nitrogen content in apple leaves. Journal of Experimental Botany, 52, 1865–1872.CrossRefGoogle Scholar
Chen, M., Wang, R., Yang, L., et al. (2003). Development of east Asian summer monsoon environments in the late Miocene: radiolarian evidence from Site 1143 of ODP Leg 184. Marine Geology, 201, 169–177.CrossRefGoogle Scholar
Chen, R.D. and Gadal, P. (1990). Do mitochondria provide 2-oxoglutarate needed for glutamate synthesis in higher plant chloroplasts?Plant Physiology and Biochemistry, 28, 141–145.Google Scholar
Chen, W.R., Yang, X., He, Z.L., et al. (2008). Differential changes in photosynthetic capacity, 77 K chlorophyll fluorescence and chloroplast ultrastructure between Zn-efficient and Zn-inefficient rice genotypes (Oryza sativa) under low zinc stress. Physiologia Plantarum, 132, 89–101.Google ScholarPubMed
Chen, Y.Z., Murchie, E.H., Hubbart, S., et al. (2003). Effects of season-dependent irradiance levels and nitrogen-deficiency on photosynthesis and photoinhibition in field-grown rice (Oryza sativa). Physiologia Plantarum, 117, 343–351.CrossRefGoogle Scholar
Cheng, W., Sims, D.A., Luo, Y., et al. (2000). Photosynthesis, respiration, and net primary production of sunflower stands in ambient and elevated atmospheric CO2 concentrations: an invariant NPP:GPP ratio?Global Change Biology, 6, 931–941.CrossRefGoogle Scholar
Chida, H., Nakazawa, A., Akazak, H., et al. (2007). Expression of algal cytochrome c6 in Arabidopsis enhances photosynthesis and growth. Plant Cell Physiology, 48, 948–957.CrossRefGoogle Scholar
Chidumayo, E.N. (2004). Development of Brachystegia-Julbernardia woodland after clear-felling in central Zambia: Evidence for high resilience. Applied Vegetation Science, 7, 237–242.Google Scholar
Cho, U.H. and Park, J.O. (2000). Mercury induced oxidative stress in tomato seedlings. Plant Science, 156, 1–9.CrossRefGoogle ScholarPubMed
Chopra, J., Kaur, N. and Gupta, A.K. (2002). A comparative developmental pattern of enzymes of carbon metabolism and pentose phosphate pathway in mung bean and lentil nodules. Acta Physiologia Plantarum, 24, 67–72.CrossRefGoogle Scholar
Chou, H.M., Bundock, N., Rolfe, A.S., et al. (2000). Infection of Arabidopsis thaliana leaves with Albugo candida (white blister rust) causes a reprogramming of host metabolism. Molecular Plant Pathology, 2, 99–113.CrossRefGoogle Scholar
Chow, W.S. and Hope, A.B. (2004). Electron fluxes through photosystem I in cucumber leaf discs probed by far-red light. Photosynthesis Research, 81, 77–89.CrossRefGoogle ScholarPubMed
Christen, D., Schonmann, S., Jermini, M., et al. (2007). Characterization and early detection of grapevine (Vitis vinifera) stress responses to esca disease by in situ chlorophyll fluorescence and comparison with drought stress. Environmental and Experimental Botany, 60, 504–514.CrossRefGoogle Scholar
Christensen, J.H., Hewitson, B., Busuioc, A., et al. (2007). Regional climate projections. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (eds) Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M and Miller, H.L., Cambridge University Press, Cambridge, UK and New York, USA.Google Scholar
Christin, P.A.Besnard, G., Samaritani, E., et al. (2008) Oligocene CO2 decline promoted C4 photosynthesis in grasses. Current Biology, 18, 37–43.CrossRefGoogle ScholarPubMed
Christin, P.A., Salamin, N., Savolainen, V., et al. (2007). C4 photosynthesis evolved in grasses via parallel adaptive genetic changes. Current Biology, 17, 1241–1247.CrossRefGoogle ScholarPubMed
Christmann, A., Hoffmann, T., Teplova, I., et al. (2005). Generation of active pools of abscisic acid revealed by in vivo imaging of water-stressed Arabidopsis. Plant Physiology, 137, 209–219.CrossRefGoogle ScholarPubMed
Christmann, A., Weiler, E., Steudle, E., et al. (2007). A hydraulic signal in root-to-shoot signalling of water shortage. Plant Journal, 52, 167–174.CrossRefGoogle ScholarPubMed
Ciais, P. and Meijer, H.A.J. (1998). The 18O/16O isotope ratio of atmospheric CO2 and its role in global carbon cycle research. In: Stable Isotopes. Integration of Biological, Ecological and Geochemical Processes (ed. Griffiths, H.), BIOS Scientific Publishers, Oxford, UK, pp. 409–431.Google Scholar
Ciais, P., Denning, A.S., Tans, P.P., et al. (1997a). A three-dimensional synthesis study of δ18O in atmospheric CO2: 1. Surface fluxes. Journal of Geophysical Research Atmospheres, 102, 5857–5872.CrossRefGoogle Scholar
Ciais, P., Friedlingstein, P., Friend, A., et al. (2001). Integrating global models of terrestrial primary productivity. In: Terrestrial Global Productivity (eds Mooney, H.A., Saugier, B. and Roy, J.), Academic Press, San Diego, CA, USA, pp. 449–478.Google Scholar
Ciais, P., Reichstein, M., Viovy, N., et al. (2005). Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature, 437, 529–533.CrossRefGoogle ScholarPubMed
Ciais, P., Tans, P.P., Denning, A.S., et al. (1997b) A three-dimensional synthesis study of δ18O in atmospheric CO2: 2. Simulation with the TM2 transport model. Journal of Geophysical Research Atmospheres, 102, 5873–5883.Google Scholar
Ciais, P., Tans, P.P., Trolier, M., et al. (1995a). A large northern-hemisphere terrestrial CO2 sink indicated by the 13C/12C ratio of atmospheric CO2. Science, 269, 1098–1102.CrossRefGoogle ScholarPubMed
Ciais, P., Tans, P.P., White, J.W.C., et al. (1995b). Partitioning of ocean and land uptake of CO2 as inferred by the δ13C measurements from the NOAA climate monitoring and diagnostic laboratory global air sampling network. Journal of Geophysical Research Atmospheres, 100, 5051–5070.CrossRefGoogle Scholar
Cifre, J., Bota, J., Escalona, J.M., et al. (2005). Physiological tools for irrigation scheduling in grapevines (Vitis vinifera L.): an open gate to improve water-use efficiency?Agriculture Ecosystems and Environment, 106, 159–170.CrossRefGoogle Scholar
Cipollini, D. (2005). Interactive effects of lateral shading and jasmonic acid on morphology, physiology, seed production, and defense traits in Arabidopsis thaliana. International Journal of Plant Science, 166, 955–959.CrossRefGoogle Scholar
Cipollini, M.L. and Levey, D.J. (1991). Why some fruits are green when they are ripe: carbon balance in fleshy fruits. Oecologia, 88, 371–377.CrossRefGoogle ScholarPubMed
Cirilo, A.G. and Andrade, F.H. (1994). Sowing date and maize productivity: I. Crop growth and dry matter partitioning. Crop Science, 34, 1039–1043.CrossRefGoogle Scholar
Cleland, R.E. (1998). Voltammetric measurement of the plastoquinone redox state in isolated thylakoids. Photosynthesis Research, 58, 183–92.CrossRefGoogle Scholar
Clement, J.M.A.M., Venema, J.H. and Hasselt, P.R. (1995). Short-term exposure to atmospheric ammonia does not affect low-temperature hardening of winter wheat. New Phytologist, 131, 345–351.CrossRefGoogle Scholar
Clement, R.J., Burba, G.G., Grelle, A., et al. (2009). Improved trace gas flux estimation through IRGA sampling optimization. Agricultural and Forest Meteorology, 149, 623–638.CrossRefGoogle Scholar
Close, D.C., Beadle, C.L. and Hovenden, M.J. (2003). Interactive effects of nitrogen and irradiance on sustained xanthophyll cycle engagement in Eucalyptus nitens leaves during winter. Oecologia, 134, 32–36.Google ScholarPubMed
Clough, B.F. (1992). Primary productivity and growth of mangrove forests. In: Tropical Mangrove Ecosystems (eds Robertson, A.I. and Alongi, D.M.), American Geophysical Union, Washington DC, USA, pp. 225–249.CrossRef
Clough, B.F. and Sim, R.G. (1989). Changes in gas exchange characteristics and water use efficiency of mangroves in response to salinity and vapour pressure deficit. Oecologia, 79, 38–44.CrossRefGoogle ScholarPubMed
Cochard, H., Coll, L., Le Roux, X., et al. (2002). Unraveling the effects of plant hydraulics on stomatal closure during water stress in walnut. Plant Physiology, 128, 282–290.CrossRefGoogle Scholar
Cochard, H., Venisse, J.S., Barigah, T.S., et al. (2007). Putative role of aquaporins in variable hydraulic conductance of leaves in response to light. Plant Physiology, 143, 122–133.CrossRefGoogle ScholarPubMed
Cochrane, M.A., Alencar, A., Schulze, M. D., et al. (1999). Positive feedbacks in the fire dynamic of closed canopy tropical forests. Science, 284, 1832–1835.CrossRefGoogle ScholarPubMed
Cohen, J. and Loebenstein, G. (1975). An electron microscope study of starch lesions in cucumber cotyledons infected with tobacco mosaic virus. Phytopathology, 65, 32–39.CrossRefGoogle Scholar
Cohen, W.B. and Goward, S.N. (2004). Landsat’s role in ecological applications of remote sensing. BioScience, 54, 535–545.CrossRefGoogle Scholar
Cole, D.R. and Monger, H.C. (1994). Infuence of atmospheric CO2 on the decline of C4 plants during the last deglaciation. Nature, 368, 533–536.CrossRefGoogle Scholar
Coleman, J.S., McConnaughay, K.D.M. and Bazzaz, F.A. (1993). Elevated CO2 and plant nitrogen-use: is reduced tissue nitrogen concentration size-dependent?Oecologia, 93, 195–200.CrossRefGoogle 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
Coley, P.D., Bryant, J.P. and Chapin, F.S.I. (1985). Resource availability and antiherbivore defense. Science, 230, 895–899.CrossRefGoogle ScholarPubMed
Collatz, G.J., Ball, J.T., Grivet, C., et al. (1991). Physiological and environmental regulation of stomatal conductance, photosynthesis and transpiration: a model that includes a laminar boundary layer. Agricultural and Forest Meteorology, 54, 107–136.CrossRefGoogle Scholar
Collatz, G.J., Berry, J.A. and Clark, J.S. (1998). Effects of climate and atmospheric CO2 partial pressure on the global distribution of C4 grasses: present, past, and future. Oecologia, 114, 441–454.CrossRefGoogle ScholarPubMed
Collatz, G.J., Berry, J.A., Farquhar, G.D., et al. (1990). The relationship between the Rubisco reaction-mechansim and models of photosynthesis. Plant, Cell and Environment, 13, 219–225.CrossRefGoogle Scholar
Collatz, G.J., Ribas-Carbo, M. and Berry, J.A. (1992). Coupled photosynthesis-stomatal conductance model for leaves of C4 plants. Austalian Journal of Plant Physiology, 19, 519–538.CrossRefGoogle Scholar
Collier, D.E. and Thibodeau, B.A. (1995). Changes in respiration and chemical content during autumnal senescence of Populus tremuloides and Quercus rubra leaves. Tree Physiology, 15, 759–764.CrossRefGoogle ScholarPubMed
Colmer, T.D. and Pedersen, O. (2008). Underwater photosynthesis and respiration in leaves of submerged wetland plants: gas films improve CO2 and O2 exchange. New Phytologist, 177, 918–926.CrossRefGoogle ScholarPubMed
Comai, L., Young, K., Till, B.J., et al. (2004). Efficient discovery of DNA polymorphisms in natural populations by ecotilling. Plant Journal, 37, 778–786.CrossRefGoogle ScholarPubMed
Comstock, J.P. (2002). Hydraulic and chemical signalling in the control of stomatal conductance and transpiration. Journal of Experimental Botany, 53, 195–200.CrossRefGoogle ScholarPubMed
Comstock, J.P. and Ehleringer, J.R. (1988). Contrasting photosynthetic behaviour in leaves and twigs of Hymenoclea salsola, a green twigged wann desert shrub. American Journal of Botany, 75, 1369–1370.CrossRefGoogle Scholar
Conde, L.F. and Kramer, P.J. (1975). The effect of vapour pressure deficit on diffusion resistance of Opuntia compressa. Canadian Journal of Botany, 53, 2923–2926.CrossRefGoogle Scholar
Condon, A.G., Richards, R.A., Rebetzke, G.J., et al. (2002). Improving intrinsic water-use efficiency and crop yield. Crop Science, 42, 122–131.CrossRefGoogle ScholarPubMed
Condon, A.G., Richards, R.A., Rebetzke, G.J., et al. (2004). Breeding for high water-use efficiency. Journal of Experimental Botany, 55, 2447–2460.CrossRefGoogle ScholarPubMed
Conroy, J.P., Milham, P.J., Reed, M.L., et al. (1990). Increases in phosphorus requirements for CO2-enriched pine species. Plant Physiology, 92, 977–982.CrossRefGoogle Scholar
Conti, G.C., Vegetti, G., Bassi, M., et al. (1972). Some ultrastructural and cytochemical observations on Chinese cabbage leaves infected with cauliflower mosaic virus. Virology, 47, 694–700.CrossRefGoogle ScholarPubMed
Conway, T.J., Tans, P., Waterman, LS., et al. (1994). Evidence for interannual variability of the carbon cycle from the National Oceanic and Atmospheric Administration/Climate Monitoring and Diagnostics Laboratory Global Air Sampling Network. Journal of Geophysical Research, 99, 831–855.CrossRefGoogle Scholar
Coops, N.C., Hilker, T., Hall, F.G. et al. (2010). Estimation of light-use efficiency of terrestrial ecosystems from space: a status report. BioScience, 60, 788–797.CrossRefGoogle Scholar
Coplen, T.B. (1995). Discontinuance of SMOW and PDB. Nature, 375, 285.CrossRefGoogle Scholar
Copolovici, L.O., Filella, I., Llusià, J., et al. (2005). The capacity for thermal protection of photosynthetic electron transport varies for different monoterpenes in Quercus ilex. Plant Physiology, 139, 485–496.CrossRefGoogle ScholarPubMed
Corcuera, L., Camarero, J.J. and Gil-Pelegrín, E. (2002). Functional groups in Quercus species derived from the analysis of pressure–volume curves. Trees, 16, 465–472.CrossRefGoogle Scholar
Cordell, S., Goldstein, G., Mueller-Dombois, D., et al. (1998). Physiological and morphological variation in Metrosideros polymorpha, a dominant Hawaiian tree species, along an altitudinal gradient: the role of phenotypic plasticity. Oecologia, 113, 188–196.CrossRefGoogle ScholarPubMed
Corelli-Grappadelli, L. and Magnanini, E. (1993). A whole-tree system for gas-exchange studies. HortScience, 28, 41–45.Google Scholar
Corey, K.A. and Wheeler, R.M. (1992). Gas exchange in NASA’s biomass production chamber. BioScience, 42, 503–509.CrossRefGoogle ScholarPubMed
Corley, R.H.V. and Lee, C.H. (1992). The physiological basis for genetic improvement of oil palm in Malaysia. Euphytica, 60, 179–84.Google Scholar
Cornic, G. (1973). Etude de l’inhibition de la respiration par la lumière chez la moutarde blanche Sinapis alba L. Physiologie Végetale, 11, 663–679.Google Scholar
Cornic, G. (2000). Drought stress inhibits photosynthesis by decreasing stomatal aperture – not by affecting ATP synthesis. Trends in Plant Science, 5, 187–188.CrossRefGoogle Scholar
Cornic, G. and Briantais, J.M. (1991). Partitioning of photosynthetic electron flow between CO2 and O2 reduction in a C3 leaf (Phaseolus vulgaris L.) at different CO2 concentrations and during drought. Planta, 183, 178–184.CrossRefGoogle Scholar
Cornic, G., Bukhov, N.G., Wiese, C., et al. (2000). Flexible coupling between light-dependent electron and vectorial proton transport in illuminated C3 plants. Role of photosystem I-dependent proton pumping. Planta, 210, 468–477.CrossRefGoogle ScholarPubMed
Cornic, G., Le Gouallec, J.L., Briantais, J.M., et al. (1989). Effect of dehydration and high light on photosynthesis of two C3 plants. Phaseolus vulgaris L. and Elastostema repens (hour.) Hall f.). Planta, 177, 84–90.CrossRefGoogle Scholar
Corp, L., McMurtrey, J., Middleton, E., et al. (2003). Fluorescence sensing systems: in vivo detection of biophysical variations in field corn due to nitrogen supply. Remote Sensing of Environment, 86, 470–479.CrossRefGoogle Scholar
Corp, L.A., McMurtrey, J.E., Chappelle, E.W., et al. (1997). UV band fluorescence (in vivo) and its implications to the remote assessment of nitrogen supply in vegetation. Remote Sensing of the Environment, 61, 110–117.CrossRefGoogle Scholar
Cosgrove, D.J. (2000). Expansive growth of plant cell walls. Plant Physiology and Biochemistry, 38, 109–124.CrossRefGoogle ScholarPubMed
Cosgrove, D.J. (2005). Growth of the plant cell wall. Nature Reviews Molecular Cell Biology, 6, 850–861.CrossRefGoogle ScholarPubMed
Costa, J.M., Ortuño, M.F. and Chaves, M.M. (2007). Deficit irrigation as a strategy to save water: physiology and potential application to horticulture. Journal of Integrative Plant Biology, 49, 1421–1434.CrossRefGoogle Scholar
Cousins, A.B., Badger, M.R. and von Caemmerer, S. (2006). Carbonic anhydrase and its influence on carbon isotope discrimination during C4 photosynthesis. Insights from antisense RNA in Flaveria bidentis. Plant Physiology, 141, 232–242.CrossRefGoogle ScholarPubMed
Cowan, I.R. (1986). Economics of carbon fixation in higher plants. In: On the Economy of Plant Form and Function (ed. Givnish, T.J.), Cambridge University Press, Cambridge, UK, pp. 133–170.Google Scholar
Cowan, I.R. and Faquhar, G.D. (1977). Stomatal function in relation to leaf metabolism and environment. Symposia of the Society for Experimental Biology, 31, 471–505.Google Scholar
Cowling, S.A., Jones, C.D. and Cox, P.M. (2007). Consequences of the evolution of C4 photosynthesis for surface energy and water exchange. Journal of Geophysical Research – Biogeosciences 112: Article number G01020.CrossRefGoogle Scholar
Cox, P.M., Betts, R.A., Jones, C.D., et al. (2000). Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate model. Nature, 408, 184–187.CrossRefGoogle Scholar
Crafts-Brandner, S.J. and Salvucci, M.E. (2000). Rubisco activase constrains the photosynthetic potential of leaves at high temperature and CO2. Proceedings of the National Academy of Sciences of the United States of America, 97, 13430–13435.CrossRefGoogle ScholarPubMed
Craig, H. (1953). The geochemistry of the stable carbon isotopes. Geochimica et Cosmochimica Acta, 3, 53–92.CrossRefGoogle Scholar
Craig, H. (1954). Carbon 13 in plants and the relationship between carbon 13 and carbon 14 variation in nature. The Journal of Geology, 62, 115–149.CrossRefGoogle Scholar
Craig, H. (1961). Isotopic variations in meteoric waters. Science, 133, 1702–1703.CrossRefGoogle ScholarPubMed
Craig, H. and Gordon, L.I. (1965). Deuterium and oxygen-18 variations in the ocean and the marine atmosphere. In: Proceedings of a Conference on Stable Isotopes in Oceanographic Studies and Paleo-temperatures (ed. Tongiorgi, E.), Lischi and Figli, Pisa, Italy, pp. 9–130.Google Scholar
Craine, J.M. and Reich, P.B. (2005). Leaf-level light compensation points in shade-tolerant woody seedlings. New Phytologist, 166, 710–713.CrossRefGoogle ScholarPubMed
Cramer, M.D., Hawkins, H.J. and Verboom, G.A. (2009). The importance of nutritional regulation of plant water flux. Oecologia, 161, 15–24.CrossRefGoogle ScholarPubMed
Cramer, M.D., Kleizen, C. and Morrow, C. (2007). Does the prostrate-leaved geophyte Brunsvigia orientalis utilize soil-derived CO2 for photosynthesis?Annals of Botany, 99, 835–844.CrossRefGoogle ScholarPubMed
Crane, J.H. and Davies, F.S. (1988). Flooding duration and seasonal effects on growth and development of young rabbiteye blueberry plants. Journal of the American Society for Horticultural Science, 113, 180–184.Google Scholar
Crane, P.R. (1996). The fossil history of the Gnetales. International Journal of Plant Sciences, 157 (suppl), S50–S57.CrossRefGoogle Scholar
Crane, P.R., Herendeen, P. and Friis, E.M. (2004). Fossils and plant phylogeny. American Journal of Botany, 91, 1683–1699.CrossRefGoogle ScholarPubMed
Crayn, D.M., Terry, R.G., Smith, J.A.C., et al. (2000). Molecular systematic investigations in Pitcarnioideae (Bromeliaceae) as a basis for understanding the evolution of crassaulacean acid metabolism (CAM). In: Monocots: Systematics and Evolution (eds Wilson, K.L. and Morrison, D.A.), CSIRO, Melbourne, Australia, pp. 569–579.Google Scholar
Crayn, D.M., Winter, K. and Smith, J.A.C. (2004). Multiple origins of crassulacean acid metabolism and the epiphytic habit in the neotropical family Bromeliaceae. Proceedings of the National Academy of Sciences USA, 101, 3703–3708.CrossRefGoogle ScholarPubMed
Crofts, A.R., Deamer, D.W. and Packer, L. (1967). Mechanisms of light-induced structural change in chloroplasts ii. The role of ion movements in volume changes. Biochimica et Biophysica Acta, 131, 97–118.CrossRefGoogle Scholar
Crofts, A.R., Hong, S., Ugulava, N., et al. (1999). Pathways for proton release during ubihydroquinone oxidation by the bc1 complex. Proceedings of the National Academy of Sciences USA, 96, 10021–10026.CrossRefGoogle Scholar
Croker, J.L., Witte, W.T. and Augé, R.M. (1998). Stomatal sensitivity of six temperate, deciduous tree species to non-hydraulic root-to-shoot signalling of partial soil drying. Journal of Experimental Botany, 49, 761–774.CrossRefGoogle Scholar
Cromer, R.N., Kriedemann, P.E., Sands, P.J., et al. (1993). Leaf growth and photosynthetic response to nitrogen and phosphorus in seedling trees of Gmelina arborea. Australian Journal of Plant Physiology, 20, 83–98.CrossRefGoogle Scholar
Crosatti, C., Soncini, C., Stanca, A.M., et al. (1995). The accumulation of a cold-regulated chloroplastic protein is light-dependent. Planta, 196, 458–465.CrossRefGoogle ScholarPubMed
Crous, K. and Ellsworth, D. (2004). Canopy position affects photosynthetic adjustments to long-term elevated CO2 concentration (FACE) in aging needles in a mature Pinus taeda forest. Tree Physiology, 24, 961–970.CrossRefGoogle Scholar
Cseh, E., Fodor, F., Varga, A., et al. (2000). Effect of lead treatment on the distribution of essential elements in cucumber. Journal of Plant Nutrition, 23, 1095–1105.CrossRefGoogle Scholar
Csintalan, Z., Tuba, Z. and Lichtenthaler, HK. (1998). Changes in laser-induced chlorophyll fluorescence ratio F690/F735 in the poikilochlorophyllous desiccation tolerant plant Xerophyta scabrida during desiccation. Journal of Plant Physiology, 152, 540–544.CrossRefGoogle Scholar
Cui, M., Vogelmann, T.C. and Smith, W.K. (1991). Chlorophyll and light gradients in sun and shade leaves of Spinacia oleracea. Plant Cell Environment, 14, 493–500.CrossRefGoogle Scholar
Cui, X.H., Hao, F.S., Chen, H., et al. (2008). Expression of the Vicia faba VfPIP1 gene in Arabidopsis thaliana plants improves their drought resistance. Journal of Plant Research, 121, 207–214.CrossRefGoogle ScholarPubMed
Cullen, L.E., Adams, M.A., Anderson, M.J., et al. (2008). Analyses of δ13C and δ18O in tree rings of Callitris columellaris provide evidence of a change in stomatal control of photosynthesis in response to regional changes in climate. Tree Physiology, 28, 1525–1533.CrossRefGoogle Scholar
Cunnif, J., Osborne, C.P., Ripley, B.S., et al. (2008). Response of wild C4 crop progenitors to subambient CO2 highlights a possible role in the origin of agriculture. Global Change Biology, 14, 576–587.CrossRefGoogle Scholar
Cunningham, F.X. Jr. (2002). Regulation of carotenoid synthesis and accumulation in plants. Pure and Applied Chemistry, 74, 1409–1417.CrossRefGoogle Scholar
Curtis, P.S. (1996). A meta-analysis of leaf gas exchange and nitrogen in trees grown under elevated carbon dioxide. Plant Cell And Environment, 19, 127–137.CrossRefGoogle Scholar
Cushman, J.C. (2001). Crassulacean acid metabolism. A plastic photosynthetic adaptation to arid environments. Plant Physiology, 127, 1439–1448.CrossRefGoogle ScholarPubMed
Cushman, J.C. and Bohnert, H.J. (2002). Induction of crassulacean acid metabolism by salinity: molecular aspects. In: Salinity: Environmet-Plants-Molecules (eds Läuchli, A. and Lüttge, U.), Dordrecht, Boston, London. Kluwer Academic Publishers, 361–393.Google Scholar
Cushman, J.C. and Borland, A.M. (2002). Induction of crassulacean acid metabolism by water limitatation. Plant, Cell and Environment, 25, 295–310.CrossRefGoogle Scholar
Cutler, S.R., Rodriguez, P.L., Finkelstein, R.R., et al. (2010). Abscisic acid: emergence of a core signaling network. Annual Review of Plant Biology, 61, 651–679.CrossRefGoogle ScholarPubMed
D’Antonio, C.M. (2000). Fire, plant invasions, and global changes. In: Invasive Species in a Changing World (ed. Mooney, H.A. and Hobbs, R.J.), Island Press, Washington, D.C., USA, pp. 65–93.Google Scholar
Dai, Y.J., Dickinson, R.E. and Wang, Y.P. (2004). A two-big-leaf model for canopy temperature, photosynthesis, and stomatal conductance. Journal of Climate, 17, 2281–2299.2.0.CO;2>CrossRefGoogle Scholar
Dal Bosco, C., Busconi, M., Covoni, C., et al. (2003). Cor gene expression in barley mutants affected in chloroplast development and photosynthetic electron transport. Plant Physiology, 131, 793–802.CrossRefGoogle ScholarPubMed
Dale, J.E. (1982). The growth of leaves. International Botanical Congress, 1981. Edward Arnold Ltd., London.Google Scholar
Dale, V.H., Joyce, L.A., McNulty, S., et al. (2001). Climate change and forest disturbances. Bioscience, 51, 723–734.CrossRefGoogle Scholar
Daley, P.F. (1995). Chlorophyll fluorescence analysis and imaging in plant stress and disease. Canadian Journal of Plant Pathology, 17, 167–173.CrossRefGoogle Scholar
Daley, P.F., Raschke, K., Ball, J.T., et al. (1989). Topography of photosynthetic activity of leaves obtained from video images of chlorophyll fluorescence. Plant Physiology, 90, 1233–1238.CrossRefGoogle ScholarPubMed
Dalling, J.W. and Harms, K.E. (1999). Damage tolerance and cotyledonary resource use in the tropical tree Gustavia superba. Oikos, 85, 257–264.CrossRefGoogle Scholar
Damesin, C. and Rambal, S. (1995). Field-study of leaf photosynthetic performance by a Mediterranean deciduous oak tree (Quercus pubescens) during a severe summer drought. The New Phytologist, 131, 159–167.Google Scholar
Damesin, C., Ceschia, E., Le Goff, N., et al. (2002). Stem and branch repiration of beech: from tree measurements to estimations at the stand level. New Phytologist, 153, 159–172.CrossRefGoogle Scholar
Damesin, C., Rambal, S. and Joffre, R. (1998). Co-occurrence of trees with differing leaf habit: a functional approach on Mediterranean oaks. Acta Oecologica, 19, 195–204.CrossRefGoogle Scholar
Damm, A., Elbers, J., Releer, A., et al. (2010). Remote sensing of sun induced fluorescence to improve modelling of diurnal courses of gross primary production. Global Change Biology, 16, 171–186.CrossRefGoogle Scholar
Damour, G., Vandame, M. and Urban, L. (2009). Long-term drought results in a reversible decline in photosynthetic capacity in mango leaves, not just a decrease in stomatal conductance. Tree Physiology, 29, 675–684.CrossRefGoogle Scholar
Danner, B.T. and Knapp, A.K. (2001). Growth dynamics of oak seedlings (Quercus macrocarpa Michx. and Quercus muhlenbergii Engelm.) from gallery forests: implications for forest expansion into grasslands. Trees – Structure and Function, 15, 271–277.CrossRefGoogle Scholar
Danon, A., Coll, N.S. and Apel, K. (2006). Cryptochrome-1-dependent execution of programmed cell death induced by singlet oxygen in Arabidopsis thaliana. Proceedings of the National Academy of Sciences (USA), 103, 17036–17041.CrossRefGoogle ScholarPubMed
Dansgaard, W. (1954). The 18O abundance in fresh water. Geochimica et Cosmochimica Acta, 6, 241–260.CrossRefGoogle Scholar
Dansgaard, W. (1964). Stable isotopes in precipitation. Tellus, 16, 436–467.CrossRefGoogle Scholar
D’Antonio, C.M., Hughes, R.F. and Vitousek, P.M. (2001). Factors influencing dynamics of two invasive C4 grasses in seasonally dry Hawaiian woodlands. Ecology, 82, 89–104.Google Scholar
Dardick, C. (2007). Comparative expression profiling of Nicotiana benthamiana leaves systemically infected with three fruit tree viruses. Molecular Plant-Microbe Interactions, 20, 1004–1017.CrossRefGoogle ScholarPubMed
Darley, E.F. (1966). Studies of the effect of cement-kiln dust on vegetation. Journal of the Air Pollution Control Association, 16, 145–150.CrossRefGoogle ScholarPubMed
Darrall, N.M. (1989). The effect of air pollutants on physiological processes in plants. Plant, Cell and Environnment, 12, 1–30.CrossRefGoogle Scholar
Darwin, C. (1881a). The movements of leaves. Nature, 23, 603–604.CrossRefGoogle Scholar
Darwin, C. (1881b). Leaves injured at night by free radiation. Nature, 24, 459.CrossRefGoogle Scholar
Darwin, C. (1882). The action of carbonate of ammonia on chlorophyll-bodies. Journal of the Linnean Society of London (Botany), 19, 262–284.CrossRefGoogle Scholar
Darwin, F. (1916). On the relation between transpiration and stomatal aperture. Philosophical Transactions of the Royal Society of London Series B, 207, 413–437.CrossRefGoogle Scholar
Darwin, F. and Pertz, D.F.M. (1911) On a new method of estimating the aperture of stomata. Proceedings of the Royal Society of London Series B, 84, 136–154.CrossRefGoogle Scholar
Dat, J., Vandenabeele, S., Vranova, E., et al. (2000). Dual action of the active oxygen species during plant stress responses. Cellular and Molecular Life Sciences: CMLS, 57, 779–795.CrossRefGoogle ScholarPubMed
Dau, H. (1994). Molecular mechanisms and quantitative models of variable photosystem II fluorescence. Photochemistry Photobiology, 60, 1–23.CrossRefGoogle Scholar
Daughtry, C.S.T., Gallo, K.P., Goward, S.N., et al. (1992). Spectral estimates of absorbed radiation and phytomass production in corn and soybean canopies. Remote Sensing of the Environment, 39, 141–52.CrossRefGoogle Scholar
Daumard, F. (2010) “Contribution à la télédetection passive de la fluorescence chlorophyllienne des végétaux”. PhD thesis. L.M.D, Ecole Polytechnique, Palaiseau - France.
Daumard, F., Champagne, S., Fournier, A., et al. (2010). A field platform for long-term measurements of canopy fluorescence. IEEE Trans. Geosci. Remote Sens., 48 (9), 3358–3368.CrossRef
Daumard, F., Goulas, Y., Champagne, S., et al. (2012) Canopy level chlorophyll fluorescence at 760 nm better tracks in-field sorghum growht. IEEE Trans. Geosci. Remote Sens (in press)
Daumard, F., Goulas, Y., Ounis, A., et al. (2007). Atmospheric correction of airborne passive measurements of fluorescence. 10th International Symposium on Physical Measurements and Signatures in Remote Sensing, March 12–14, 2007, Davos, Switzerland (ISPMSRS07).
Davey, P.A., Hunt, S., Hymus, G.J., et al. (2004). Respiratory oxygen uptake is not decreased by an instantaneous elevation of [CO2], but is increased with long-term growth in the field at elevated [CO2](1). Plant Physiology, 134, 520–527.CrossRefGoogle Scholar
Davey, P.A., Olcer, H., Zakhleniuk, O., et al. (2006). Can fast-growing plantation trees escape biochemical down-regulation of photosynthesis when grown throughout their complete production cycle in the open air under elevated carbon dioxide?Plant, Cell and Environment, 29, 1235–1244.CrossRefGoogle ScholarPubMed
Davi, H., Dufrêne, E., Granier, A., et al. (2005). Modelling carbon and water cycles in a beech forest. Part II.: Validation of the main processes from organ to stand scale. Ecological Modeling, 185, 387–405.CrossRefGoogle Scholar
Davies, B. and Sharp, B. (Ed.) (2000). Water deficits and plant growth. Journal of Experimental Botany, (51). 350 WD Special Issue.Google Scholar
Davies, B., Tuberosa, R., Blum, A., et al. (2007). Integrated approaches to sustain and improve plant production under drought stress. Journal of Experimental Botany, 58, (2). Special Issue.Google Scholar
Davies, B.H. (1976). Carotenoids. In: Chemistry and Biochemistry of Plant Pigments (ed. Goodwin, T.W.) Vol 2. Academic Press, London, UK, pp. 38–165.Google Scholar
Davies, S.J. (1998). Photosynthesis of nine pioneer Macaranga species from Borneo in relation to life history. Ecology, 79, 2292–2308.CrossRefGoogle Scholar
Davies, W.J. and Zhang, J. (1991). Root signals and the regulation of growth and development of plants in drying soil. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 55–76.CrossRefGoogle Scholar
Davies, W.J., Metcalfe, J., Lodge, T.A., et al. (1986). Plant growth substances and regulation of growth under drought. Australian Journal of Plant Physiology, 13, 105–125.Google Scholar
Davis, S.C., Anderson-Teixeira, K.J. and DeLucia, E.H. (2009). Life-cycle analysis and the ecology of biofuels. Trends in Plant Science, 14, 140–146.CrossRefGoogle ScholarPubMed
Davison, P.A., Hunter, C.N. and Horton, P. (2002). Overexpression of β-carotene hydroxylase enhances stress tolerance in Arabidopsis. Nature, 418, 203–206.CrossRefGoogle ScholarPubMed
Dawson, J.O. and Gordon, J.C. (1979). Nitrogen fixation in relation to photosynthesis in Alnus glutinosa. Botanical Gazette, 140, S70–S75.CrossRefGoogle Scholar
Dawson, T.E. (1993). Water sources of plants as determined from xylem-water isotopic composition: perspectives on plant competition, distribution, and water relations. In: Stable Isotopes and Plant Carbon-Water Relations (eds. Ehleringer, J.R., Hall, A.E. and Farquhar, G.D.) Academic Press, San Diego, California, USA, pp. 465–496.Google Scholar
Dawson, T.E., Mambelli, S., Plamboeck, A.H., et al. (2002). Stable isotopes in plant ecology. Annual Review of Ecology and Systematics, 33, 507–559.CrossRefGoogle Scholar
Day, D.A., Krab, K., Lambers, H., et al. (1996). The cyanide-resistant oxidase: to inhibit or not to inhibit, that is the question. Plant Physiology, 110, 1–2.CrossRefGoogle ScholarPubMed
Day, D.A., Neuberger, M. and Douce, R. (1985). Interactions between glycine decarboxylase, the tricarboxylic acid cycle and the respiratory chain in pea leaf mitochondria. Australian Journal of Plant Physiology, 12, 119–130.CrossRefGoogle Scholar
Day, T.A., Ruhland, C.T., Grobe, C.W., et al. (1999). Growth and reproduction of Antarctic vascular plants in response to warming and UV radiation reductions in the field. Oecologia, 119, 24–35.CrossRefGoogle ScholarPubMed
De Chalain, T.M.B. and Berjak, B. (1979). Cell death as a functional event in the development of the leaf intercellular spaces in Avicennia marina (Forsskål) Vierh. New Phytologist, 83, 147–155.CrossRefGoogle Scholar
De Jong, G. (2005). Evolution of phenotypic plasticity: patterns of plasticity and the emergence of ecotypes. New Phytologist, 166, 101–118.CrossRefGoogle ScholarPubMed
De Niro, M.J. and Epstein, S. (1979). Relationship between the oxygen isotope ratio of terrestrial plant cellulose, carbon dioxide and water. Science, 204, 51–53.CrossRefGoogle Scholar
De Pury, D.G.G. and Farquhar, G.D. (1997). Simple scaling of photosynthesis from leaves to canopies without the errors of big-leaf models. Plant Cell Environment, 20, 537–557.CrossRefGoogle Scholar
De Ronde, J.A., Cress, W.A., Krüger, G.H.J., et al. (2004). Photosynthetic response of transgenic soybean plants, containing an Arabidopsis P5CR gene, during heat and drought stress. Journal of Plant Physiology, 161, 1211–1224.CrossRefGoogle ScholarPubMed
De Souza, A.A., Takita, M.A., Coletta-Filho, H.D., et al. (2007). Analysis of expressed sequence tags from Citrus sinensis L. Osbeck infected with Xylella fastidiosa. Genetics and Molecular Biology, 30, 957–964.CrossRefGoogle Scholar
De Souza, J., Silka, P.A. and Davis, S.D. (1986). Comparative physiology of burned and unburned Rhus laurina after chaparral wildfire. Oecologia, 71, 63–68.CrossRefGoogle Scholar
Deamer, D.W., Crofts, A.R. and Packer, L. (1967). Mechanisms of light-induced structural changes in chloroplasts. I. Light-scattering increments and ultrastructural changes mediated by proton transport. Biochimica et Biophysica Acta, 131, 81–96.CrossRefGoogle Scholar
Decker, H. (1969). Phytonematologie. In: Biologie und Bekämpfung Pflanzenparasitärer Nematoden. Deutscher Landwirtschaftsverlag, Berlin, Germany, pp. 505.Google Scholar
DeFries, R.S., Townshend, J.R.G. and Hansen, M.C. (1999). Continuous fields of vegetation characteristics at the global scale at 1km resolution. Journal of Geophysical Research, 104, 16911–16925.CrossRefGoogle Scholar
Deighton, N., Magill, W.J., Bremner, D.H., et al. (1997). Malondialdehyde and 4-hydroxy 2-nonenal in plant tissue cultures: LC-MS determination of 2,4-dinitrophenylhydrazone derivatives. Free Radicals Research, 27, 255–265.CrossRefGoogle Scholar
De Jong, G. (2005). Evolution of phenotypic plasticity: patterns of plasticity and the emergence of ecotypes. New Phytologist, 166, 101–118.CrossRefGoogle ScholarPubMed
DeJong, T.M. (1986). Fruit effects on photosynthesis in Prunus perscia. Physiologia Plantarum, 66, 149–53.CrossRefGoogle Scholar
DeKok, L.J., Stuiver, C.E.E. and Stulen, I. (1998). Impact of H2S on plants. In: Responses of Plant Metabolism to Air Pollutants and Global Change (eds deKok, L.J. and Stulen, I.), Backhuys Publishers, Leiden, Netherlands, pp. 51–63.Google Scholar
Del Arco, J.M., Escudero, A. and Vega Garrido, M. (1991). Effects of site characteristics on nitrogen retranslocation from senescing leaves. Ecology, 72, 701–708.CrossRefGoogle Scholar
Del Pierre, N., Soudani, K., Francois, C., et al. (2009). Exceptional carbon uptake in European forests during the warm spring of 2007: a data–model analysis. Global Change Biology, 15, 1455–1474.CrossRefGoogle Scholar
Del Río, D., Stewart, A.J. and Pellegrini, N. (2005). A review of recent studies on malondialdehyde as toxic molecule and biological marker of oxidative stress. Nutrition Metabolism and Cardiovascular Diseases, 15, 316–328.CrossRefGoogle ScholarPubMed
Del Río, L.A., Pastori, G.M., Palma, J.M., et al. (1998). The activated oxygen role of peroxisomes in senescence. Plant Physiology, 116, 1195–1200.CrossRefGoogle ScholarPubMed
Delagrange, S., Messier, C., Lechowicz, M.J., et al. (2004). Physiological, morphological and allocational plasticity in understory deciduous trees: importance of plant size and light availability. Tree Physiology, 24, 775–784.CrossRefGoogle ScholarPubMed
Delaney, K.J. and Higley, L.G. (2006). An insect countermeasure impacts plant physiology: midrib vein cutting, defoliation and leaf photosynthesis. Plant, Cell and Environment, 29, 1245–1258.CrossRefGoogle ScholarPubMed
Delfine, S., Alvino, A., Villani, M.C., et al. (1999). Restrictions to carbon dioxide conductance and photosynthesis in spinach leaves recovering from salt stress. Plant Physiology, 119, 1101–1106.CrossRefGoogle ScholarPubMed
Delgado, E., Medrano, H., Keys, A.J., et al. (1995). Species variation in Rubisco specifity factor. Journal of Experimental Botany, 46, 1775–1777.CrossRefGoogle Scholar
Delgado, E., Parry, M.A.J., Vadell, J., et al. (1993). Photosynthesis, Ribulose-1,5-bisphosphate carboxylase and leaf characteristics of Nicotiana tabacum L. genotypes selected by survival at low CO2 concentrations. Journal of Experimental Botany, 44, 1–7.CrossRefGoogle Scholar
Delosme, R. (1967). Étude de l’induction de fluorescence des algues et des chloroplastes au début d’une illumination intense. Biochimica et Biophysica Acta, 143, 108–128.CrossRefGoogle Scholar
De Lucia, E.H., Coleman, J.S., Dawson, T.E., et al. (2001). Plant physiological ecology: linking the organism to scales above and below. The New Phytologist, 149, 9–16.Google Scholar
Delwiche, C.F. and Palmer, J.D. (1996). Rampant horizontal transfer and duplication of rubisco genes in eubacteria and plastids. Molecular Biology and Evolution, 13, 873–882.CrossRefGoogle ScholarPubMed
Demarty, M., Morvan, C. and Thellier, M. (1984). Calcium and the cell wall. Plant Cell Environment, 7, 441–448.CrossRefGoogle Scholar
Demicco, R.V., Lowenstein, T. K. and Hardie, L.A. (2003). Atmospheric pCO2 since 60 Ma from records of seawater pH, calcium, and primary carbonate mineralogy. Geology, 31, 793–796.CrossRefGoogle Scholar
Demmig-Adams, B. (1998). Survey of thermal energy dissipation and pigment composition in sun and shade leaves. Plant and Cell Physiology, 39, 474–482.CrossRefGoogle Scholar
Demmig-Adams, B. and Adams, III W.W. (2002). Antioxidants in photosynthesis and human nutrition. Science, 298, 2149–2153.CrossRefGoogle ScholarPubMed
Demmig-Adams, B. and Adams, III W.W. (2003). Photoinhibition. In: Encyclopedia of Applied Plant Science (eds Thomas, B., Murphy, D. and Murray, B.), Academic Press, New York, USA, pp. 707–714.Google Scholar
Demmig-Adams, B. and Adams, W.W. (1996b). The role of xanthophyll cycle carotenoids in the protection of photosynthesis. Trends Plant Science, 1, 21–26.CrossRefGoogle Scholar
Demmig-Adams, B. and Adams, W.W. III (1996a). Chlorophyll and carotenoid composition in leaves of Euonymus kiautschovicus acclimated to different degrees of light stress in the field. Australian Journal of Plant Physiology, 23, 649–659.CrossRefGoogle Scholar
Demmig-Adams, B. and Adams, W.W. III (2006). Photoprotection in an ecological context: the remarkable complexity of thermal energy dissipation. New Phytologist, 172, 11–21.CrossRefGoogle Scholar
Demmig-Adams, B., Adams, III W.W., Barker, D.H., et al. (1996a). Using chlorophyll fluorescence to assess the fraction of absorbed light allocated to thermal dissipation of excess excitation. Physiologia Plantarum, 98, 253–264.CrossRefGoogle Scholar
Demmig-Adams, B., Adams, W.W. III, Winter, K., et al. (1989). Photochemical efficiency of photosystem II, photon yield of oxygen evolution, photosynthetic capacity, and carotenoid composition during the midday depression of net CO2 uptake in Arbutus unedo frowing in Portugal. Planta, 177, 377–387.CrossRefGoogle Scholar
Demmig-Adams, B., Gilmore, A.M. and Adams, III W.W. (1996b). In vivo functions of carotenoids in higher plants. FASEB Journal, 10, 403–412.CrossRefGoogle Scholar
Den Hertog, J., Stulen, I. and Lambers, H. (1993). Assimilation, respiration and allocation of carbon in Plantago major as affected by atmospheric CO2 levels. A case study. Vegetation, 104, 369–378.CrossRefGoogle Scholar
Deng, X. and Melis, A. (1986). Phosphorylation of the light-harvesting complex II in higher plant chloroplasts: effect on photosystem II and photosystem I absorption cross section. Photobiochemistry and Photobiophysics, 13, 41–52.Google Scholar
Dengler, N. and Wilson, T. (1999). Leaf structure and development in C4 plants. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, CA, USA, pp. 133–72.Google Scholar
Dengler, N.G. (1980). Comparative histological basis of sun and shade leaf dimorphism in Helianthus annuus. Canadian Journal of Botany, 58, 717–730.CrossRefGoogle Scholar
Dengler, N.G. and Kang, J. (2001). Vascular patterning and leaf shape. Current Opinion in Plant Biology, 4, 50–56.CrossRefGoogle ScholarPubMed
Dermody, O., Long, S.P. and DeLucia, E.H. (2006). How does elevated CO2 or ozone affect the leaf-area index of soybean when applied independently?New Phytologist, 169, 145–155.CrossRefGoogle ScholarPubMed
Desai, A.R., Richardson, A.D., Moffat, A., et al. (2008). Cross-site evaluation of eddy covariance GPP and RE decomposition techniques. Agricultural and Forest Meteorology, 148, 821–838.CrossRefGoogle Scholar
Desikan, R., Cheung, M.K., Bright, J., et al. (2004). ABA, hydrogen peroxide and nitric oxide signalling in stomatal guard cells. Journal of Experimental Botany, 55, 205–212.CrossRefGoogle ScholarPubMed
Desikan, R., Mackerness, S.A.H., Hancock, J.T., et al. (2001). Regulation of the Arabidopsis transcriptome by oxidative stress. Plant Physiology, 127, 159–172.CrossRefGoogle ScholarPubMed
Dewar, R.C. (1995). Interpretation of an empirical model for stomatal conductance in terms of guard cell function. Plant, Cell and Environment, 18, 365–372.CrossRefGoogle Scholar
Dewar, R.C. (2002). The Ball-Berry Leuning and Tardieu-Davies stomatal models: synthesis and extension within a spatially aggregated picture of guard cell function. Plant, Cell and Environment, 25, 1383–1398.CrossRefGoogle Scholar
Di Cagno, R., Guidi, L., De Gara, L., et al. (2001). Combined cadmium and ozone treatments affect photosynthesis and ascorbate-dependent defences in sunflower. New Phytologist, 151, 627–636.CrossRefGoogle Scholar
Di Marco, G., Manes, F., Tricoli, D., et al. (1990). Fluorescence parameters measured concurrently with net photosynthesis to investigate chloroplastic CO2 concentration in leaves of Quercus ilex L. Journal of Plant Physiology, 136, 538–543.CrossRefGoogle Scholar
Dias, M.C. and Brüggemann, W. (2010). Limitations of photosynthesis in Phaseolus vulgaris under drought stress: gas exchange, chlorophyll fluorescence and Calvin cycle enzymes. Photosynthetica, 48, 96–102.CrossRefGoogle Scholar
Díaz, M. and Granadillo, E. (2005). The significance of episodic rains for reproductive phenology and productivity of trees in semiarid regions of northwestern Venezuela. Trees, 19, 336–348.CrossRefGoogle Scholar
Díaz-Espejo, A., Nicolás, E. and Fernández, J.E. (2007). Seasonal evolution of diffusional limitations and photosynthetic capacity in olive under drought. Plant, Cell and Environment, 30, 922–933.CrossRefGoogle ScholarPubMed
Díaz-Espejo, A., Walcroft, A.S., Fernández, J.E., et al. (2006). Modelling photosynthesis in olive leaves under drought conditions. Tree Physiology, 26, 1445–1456.CrossRefGoogle Scholar
Dickmann, D.I. (1971). Photosynthesis and respiration by developing leaves of cottonwood (Populus deltoides Bartr.). Botanical Gazette, 132, 253–259.CrossRefGoogle Scholar
Diefendorf, A.F., Mueller, K.E., Wing, S.L., et al. (2010). Global patterns in leaf 13C discrimination and implications for studies of past and future climate. Proceedings of the National Academy of Sciences USA, 107, 5738–5743.CrossRefGoogle ScholarPubMed
Diemer, M. and Körner, C. (1996). Lifetime leaf carbon balances of herbaceous perennial plants from low and high altitudes in the central Alps. Functional Ecology, 10, 33–43.CrossRefGoogle Scholar
Dietz, K.J. (1985). A possible rate limiting function of chloroplast hexosemonophosphate isomerase in starch synthesis of leaves. Biochimica et Biophysica Acta, 839, 240–248.CrossRefGoogle Scholar
Dietz, K.J. and Keller, F. (1997). Transient storage of photosynthates in leaves. In: Handbook of Photosynthesis (ed. Pessarakli, M.). Marcel Dekker, New York, USA, pp. 717–737.Google Scholar
Dietz, K.J., Jacob, S., Oelze, M.L., et al. (2006). The function of peroxiredoxins in plant organelle redox metabolism. Journal of Experimental Botany, 57, 1697–1709.CrossRefGoogle ScholarPubMed
Diffenbaugh, N.S., Pal, J.S., Trapp, R.J., et al. (2005). Fine-scale processes regulate the response of extreme events to global climate change. Proceedings of the National Academy of Sciences, 102, 15774–15778.CrossRefGoogle ScholarPubMed
Dinakar, C., Raghavendra, A.S. and Padmasree, K. (2010). Importance of AOX pathway in optimizing photosynthesis under high light stress: role of pyruvate and malate in activating AOX. Physiologia Plantarum, 139, 13–26.CrossRefGoogle ScholarPubMed
Ding, S., Lu, Q., Zhang, Y., et al. (2008). Enhanced sensitivity to oxidative stress in transgenic tobacco plants with decreased glutathione reductase activity leads to a decrease in ascorbate pool and ascorbate redox state. Plant Molecular Biology, 69, 577–592.CrossRefGoogle ScholarPubMed
Disante, K.B., Fuentes, D. and Cortina, J. (2011). Response to drought of Zn-stressed Quercus suber L. seedlings. Environmental and Experimental Botany, 70, 96–103.CrossRefGoogle Scholar
Dismukes, G.C., Klimov, V.V., Baranov, S.V., et al. (2001). The origin of atmospheric oxygen on Earth: the innovation of oxygenic photosynthesis. Proceedings of the National Academy of Sciences, 98, 2170–2175.CrossRefGoogle ScholarPubMed
Dittmar, C., Zech, W. and Elling, W. (2003). Growth variations of common beech (Fagus sylvatica L.) under different climatic and environmental conditions in Europe – a dendroecological study. Forest Ecology and Management, 173, 63–78.CrossRefGoogle Scholar
Dixon, R.K., Solomon, A.M., Brown, S., et al. (1994). Carbon pools and flux of global forest ecosystems. Science, 263, 185–190.CrossRefGoogle ScholarPubMed
Dobrowski, S.Z., Pushnik, J.C., Zarco-Tejada, P.J., et al. (2005). Simple reflectance indices track heat and water stress-induced changes in steady state chlorophyll fluorescence at the canopy scale. Remote Sensing of Environment, 97, 403–414.CrossRefGoogle Scholar
Doi, M. and Shimazaki, K.I. (2008). The stomata of the fern Adiantum capillus-veneris do not respond to CO2 in the dark and open by photosynthesis in guard cells. Plant Physiology, 147, 922–930.CrossRefGoogle Scholar
Dolan, L. (2009). Body building on land – morphological evolution of land plants. Current Opinion in Plant Biology, 12, 4–8.CrossRefGoogle ScholarPubMed
Dole, M. (1935). The relative atomic weight of oxygen in water and air. Journal of the American Chemical Society, 57, 2731–1935.CrossRefGoogle Scholar
Dole, M., Lane, G.A., Rudd, D.P. et al. (1954). Isotopic composition of atmospheric oxygen and nitrogen. Geochimica et Cosmochimica Acta, 6, 65–78.CrossRefGoogle Scholar
Döll, P. and Siebert, S. (2002). Global modelling of irrigation water requirements. Water Resource Research, 38, 1037.CrossRefGoogle Scholar
Dolman, A.J., Moors, E.J. and Elbers, J.A. (2002). The carbon uptake of a mid latitude pine forest growing on sandy soil. Agricultural and Forest Meteorology, 111, 157–170.CrossRefGoogle Scholar
Domingo, R., Pérez-Pastor, A. and Ruiz-Sánchez, M.C. (2002). Physiological responses of apricot plants grafted on two different rootstocks to flooding conditions. Journal of Plant Physiology, 159, 725–732.CrossRefGoogle Scholar
Donahue, R.A., Poulson, M.E. and Edwards, G.E. (1997). A method for measuring whole plant photosynthesis in Arabidopsis thaliana. Photosynthesis Research, 52, 263–269.CrossRefGoogle Scholar
Donald, P.W. and Strobel, G.A. (1970). Adenosine diphosphate-glucose pyrophosphorylase control of starch accumulation in rust-infected wheat leaves. Plant Physiology, 46, 126–135.Google Scholar
Donatelli, M., Hammer, G.L. and Vanderlip, R.L. (1992). Genotype and water limitation effects on phenology, growth, and transpiration efficiency in grain sorghum. Crop Science, 32, 781–786.CrossRefGoogle Scholar
Dongmann, G. (1974). Contribution of land photosynthesis to stationery enrichment of 18O in atmosphere. Radiation and Environmental Biophysics, 11, 219–225.CrossRefGoogle Scholar
Donigian, J.A.S., Barnwell, J.T.O., Jackson, R.B., et al. (1994). Assessment of alternative management practices and policies affecting soil carbon in agroecosystems of the Central United States, U.S.-EPA, Athens, GA.
Donovan, L.A. and Ehleringer, J.R. (1991). Ecophysiological differences among juvenile and reproductive plants of several woody species. Oecologia, 86, 594–597.CrossRefGoogle ScholarPubMed
Donovan, L.A. and Ehleringer, J.R. (1992). Contrasting water-use patterns among size and life-history classes of a semi-arid shrub. Functional Ecology, 6, 482–488.CrossRefGoogle Scholar
Donovan, L.A., Dudley, S.A., Rosenthal, D.M., et al. (2007). Phenotypic selection on leaf water use efficiency and related ecophysiological traits for natural populations of desert sunflowers. Oecologia, 152, 13–25.CrossRefGoogle ScholarPubMed
Donovan, L.A., Ludwig, F., Rosenthal, D.M., et al. (2009). Phenotypic selection on leaf ecophysiological traits in Helianthus. The New Phytologist, 183, 868–879.CrossRefGoogle ScholarPubMed
Donovan, L.A., Maherali, H., Caruso, C.M. et al. (2011). The evolution of the worldwide leaf economics spectrum. Trends in Plant Science, 26, 88–95.Google ScholarPubMed
Dorchin, N., Cramer, M.D. and Hoffmann, J.H. (2006). Photosynthesis and sink activity of wasp-induced galls in Acacia pycnantha. Ecology, 87, 1781–1791.CrossRefGoogle ScholarPubMed
Dore, S., Hymus, G.J., Johnson, D.P., et al. (2003). Cross-validation of open-top chamber and eddy covariance measurements of ecosystem CO2 exchange in a Florida scrub-oak ecosystem. Global Change Biology, 9, 84–95.CrossRefGoogle Scholar
Dorey, S., Baillieul, F., Pierrel, M.A., et al. (1997). Spatial and temporal induction of cell death, defense genes, and accumulation of salicylic acid in tobacco leaves reacting hypersensitively to a fungal glycoprotein elicitor. Molecular Plant-Microbe Interactions, 10, 646–655.CrossRefGoogle Scholar
Dorne, A.J., Cadel, G. and Douce, R. (1986). Polar lipid composition of leaves from nine typical alpine species. Phytochemistry, 25, 65–68.CrossRefGoogle Scholar
Douglas, N.A. and Manos, P.A. (2007). Molecular phylogeny of Nyctaginaceae: taxonomy, biogeography, and characters associated with a radiation of xerophytic genera in North America. American Journal of Botany, 94, 856–872.CrossRefGoogle ScholarPubMed
Doulis, A.G., Hausladen, A., Mondy, B., et al. (1993). Antioxidant response and winter hardiness in red spruce (Picea rubens Sarg.). New Phytologist, 23, 365–374.CrossRefGoogle Scholar
Downes, R.W. (1969). Differences in transpiration rates between tropical and temperate grasses under controlled conditions. Planta, 88, 261–273.CrossRefGoogle ScholarPubMed
Downton, W.J.S. (1971). Adaptive and evolutionary aspects of C4 photosynthesis. In: Photosynthesis and Photorespiration (eds Hatch, M.D., Osmond, C.B. and Slatyer, R.O.), Wiley Interscience, New York, USA, pp. 3–17.Google Scholar
Downton, W.J.S., Berry, J.A. and Seemann, J.R. (1984). Tolerance of photosynthesis to high temperature in desert plants. Plant Physiology, 74, 786–90.CrossRefGoogle ScholarPubMed
Downton, W.J.S., Loveys, B.R. and Grant, W.J.R. (1988a). Stomatal closure fully accounts for the inhibition of photosynthesis by abscisic acid. The New Phytologist, 108, 263–266.CrossRefGoogle Scholar
Downton, W.J.S., Loveys, B.R. and Grant, W.J.R. (1988b). Non-uniform stomatal closure induced by water stress causes putative non-stomatal inhibition of photosynthesis. The New Phytologist, 110, 503–509.CrossRefGoogle Scholar
Dowsey, A.W., English, J., Pennington, K., et al. (2006). Examination of 2-DE in the human proteome organisation brain proteome project pilot studies with the new RAIN gel matching technique. Proteomics, 6, 5030–5047.CrossRefGoogle ScholarPubMed
Doyle, J.A. (1996). Seed plant phylogeny and the relationships of the Gnetales. International Journal of Plant Sciences, 157 (suppl), S3–S39.CrossRefGoogle Scholar
Drake, B. (1992). A field study of the effects of elevated CO2 on ecosystem processes in a Chesapeake Bay wetland. Australian Journal of Botany, 40, 579–595.CrossRefGoogle Scholar
Drake, B., Leadley, P., Arp, W., et al. (1989). An open top chamber for field studies of elevated atmospheric CO2 concentration on saltmarsh vegetation. Functional Ecology, 3, 363–71.CrossRefGoogle Scholar
Drake, B.G. and Leadley, P.W. (1991). Canopy photosynthesis of crops and native plant communities exposed to long-term elevated CO2. Plant Cell Environment, 14, 853–860.CrossRefGoogle Scholar
Drake, B.G., Gonzàlez-Meler, M.A. and Long, S.P. (1997). More efficient plants: a consequence of rising atmospheric CO2?Annual Review of Plant Physiology and Molecular Biology, 48, 607–637.CrossRefGoogle ScholarPubMed
Drake, J.E., Raetz, L.M., Davis, S.C., et al. (2010). Hydraulic limitation not declining nitrogen availability causes the age-related photosynthetic decline in loblolly pine (Pinus taeda L.)Plant, Cell and Environment, 33, 1756–1766.CrossRefGoogle Scholar
Draper, H.H. and Hadley, M. (1990). Malondialdehyde determination as index of lipid peroxidation. Methods in Enzymology, 186, 421–431.CrossRefGoogle ScholarPubMed
Drennan, P.N. and Nobel, P.S. (2000). Response of CAM species to increasing atmospheric CO2 concentrations. Plant, Cell and Environment 23, 767–781.CrossRefGoogle Scholar
Dreyer, E., Le Roux, X., Montpied, P., et al. (2001). Temperature response of leaf photosynthetic capacity in seedlings from seven temperate tree species. Tree Physiology, 21, 223–232.CrossRefGoogle ScholarPubMed
Driever, S.M. and Baker, N.R. (2011). The water–water cycle in leaves is not a major alternative electron sink for dissipation of excess excitation energy when CO2 assimilation is restricted. Plant, Cell and Environment, 34, 837–846.CrossRefGoogle ScholarPubMed
Driscoll, S.P., Prins, A., Olmos, E., et al. (2006). Specification of adaxial and abaxial stomata, epidermal structure and photosynthesis to CO2 enrichment in maize leaves. Journal of Experimental Botany, 57, 381–390.CrossRefGoogle ScholarPubMed
Drolet, G.G., Huemmrich, K.F., Hall, F.G., et al. (2005). A MODIS-derived photochemical reflectance index to detect inter-annual variations in the photosynthetic light-use efficiency of a boreal deciduous forest. Remote Sensing of Environment, 98, 212–224.CrossRefGoogle Scholar
Drolet, G.G., Middleton, E.M., Huemmrich, K.F., et al. (2008). Regional mapping of gross light-use efficiency using MODIS spectral indices. Remote Sensing of Environment, 112, 3064–3078.CrossRefGoogle Scholar
Drusch, M., Moreno, J., Goulas, Y., et al. (2008). Candidate Earth Explorer Core Mission. FLEX - Fluorescence Explorer. Report for Assessment. Noordwijk, The Netherlands, European Space Agency. ESA SP-1313/4.
Dry, P.R., Loveys, B.R., McCarthy, M.G., et al. (2001). Strategic irrigation management in Australian vineyards. Journal International de la Science de la Vigne et du Vin, 35, 129–139.Google Scholar
Du, Y.C., Kawamitsu, Y., Nose, A., et al. (1996). Effects of water stress on carbon exchange rate and activities of of photosynthetic enzymes in leaves of sugarcane (Saccharum sp.). Australian Journal of Plant Physiology, 23, 719–726.CrossRefGoogle Scholar
Duan, B., Li, Y., Zhang, X., et al. (2009). Water deficit affects mesophyll limitation of leaves more strongly in sun than in shade in two contrasting Picea asperata populations. Tree Physiology, 29, 1551–1561.CrossRefGoogle Scholar
Duarte, H.M. and Lüttge, U. (2007). Circadian rhythmicity. In: Clusia: A Woody Neotropical Genus of Remarkable Plasticity and Diversity (ed. Lüttge, U.), Berlin – Heidelberg – New York. Springer, 245–256.Google Scholar
Dubois, J.J.B., Fiscus, E.L., Booker, F.L., et al. (2007). Optimizing the statistical estimation of the parameters of the Farquhar–von Caemmerer–Berry model of photosynthesis. New Phytologist, 176, 402–414.CrossRefGoogle ScholarPubMed
Ducrey, M. (1994). Influence of shade on photosynthetic gas exchange of 7 tropical rain-forest species from guadeloupe (French-West-Indies). Annales Des Sciences Forestieres, 51, 77–94.CrossRefGoogle Scholar
Ducruet, J.M. (2003). Chlorophyll thermoluminescence of leaf discs: simple instruments and progress in signal interpretation open the way to new ecophysiological indicators. Journal of Experimental Botany, 54, 2419–2430.CrossRefGoogle ScholarPubMed
Ducruet, J.M., Gaillardon, P. and Vienot, J. (1984). Use of chlorophyll fluorescence induction kinetics to study translocation and detoxication of DCMU-type herbicides in plant leaves. Zeitschrift für Naturforschung, 39c, 354–358.Google Scholar
Ducruet, J.M., Peeva, V. and Havaux, M. (2007). Chlorophyll thermofluorescence and thermoluminescence as complementary tools for the study of temperature stress in plants. Photosynthesis Research, 93, 169–171.CrossRefGoogle Scholar
Ducruet, J.M., Roman, M., Havaux, M., et al. (2005). Cyclic electron flow around PSI monitored by afterglow luminescence in leaves of maize inbred lines (Zea mays L.): correlation with chilling tolerance. Planta, 221, 567–579.CrossRefGoogle ScholarPubMed
Dudley, S.A. (1996a). The response to differing selection of plant physiological traits: evidence for local adaptation. Evolution, 50, 103–110.CrossRefGoogle ScholarPubMed
Dudley, S.A. (1996b). Differing selection on plant physiological traits in response to environmental water availability: a test of adaptive hypotheses. Evolution, 50, 92–102.CrossRefGoogle ScholarPubMed
Duffin, K.I. (2008). The representation of rainfall and fire intensity in fossil pollen and charcoal records from a South African savanna. Review of Palaeobotany and Palynology, 151, 59–71.CrossRefGoogle Scholar
Dufresne, J.L., Fairhead, L., Le Treut, H., et al. (2002). On the magnitude of positive feedback between future climate change and the carbon cycle. Geophysical Research Letters, 29, 1405.CrossRefGoogle Scholar
Dugas, W.A. (1993). Micrometeorological and chamber measurements of CO2 flux from bare soil. Agric Forest Meteorology, 67, 115–28.CrossRefGoogle Scholar
Dunagan, S.C., Gilmore, M.S. and Varekamp, J.C. (2007). Effects of mercury on visible/near-infrared reflectance spectra of mustard spinach plants (Brassica rapa P.). Environmental Pollution, 148, 301–311.CrossRefGoogle Scholar
Dunbar-Co, S., Sporck, M.J. and Sack, L. (2009). Leaf trait diversification and design in seven rare taxa of the Hawaiian Plantago radiation. International Journal of Plant Sciences, 170, 61–75.CrossRefGoogle Scholar
Duniway, J.M. and Slatyer, R.O. (1971). Gas exchange studies on the transpiration and photosynthesis of tomato leaves affected by Fusarium oxysporum f. sp. lycopersici. Phytopathology, 61, 1377–1381.CrossRefGoogle Scholar
Duranceau, M., Ghashghaie, J., Badeck, F., et al. (1999). δ13C of CO2 respired in the dark in relation to δ13C of leaf carbohydrates in Phaseolus vulgaris L. under progressive drought. Plant, Cell and Environment, 22, 515–523.CrossRefGoogle Scholar
Durand, L.Z. and Goldstein, G. (2001). Photosynthesis, photoinhibition, and nitrogen use efficiency in native and invasive tree ferns in Hawaii. Oecologia, 126, 345–354.CrossRefGoogle ScholarPubMed
Dutilleul, C., Garmier, C., Noctor, G., et al. (2003). Leaf mitochondria modulate whole cell redox homeostasis, set antioxidant capacity, and determine stress resistance through altered signaling and diurnal regulation. Plant Cell, 15, 1212–1226.CrossRefGoogle ScholarPubMed
Dutilleul, C., Lelarge, C., Prioul, J.L., et al. (2005). Mitochondria-driven changes in leaf NAD status exert a crucial influence on the control of nitrate assimilation and the integration of carbon and nitrogen metabolism. Plant Physiology, 139, 64–78.CrossRefGoogle ScholarPubMed
Dutton, R., Jiao, J., Tsujita, M.J., et al. (1988). Whole plant CO2 exchange measurements for nondestructive estimation of growth. Plant Physiology, 86, 355–358.CrossRefGoogle Scholar
Duursma, R.A., Kolari, P., Perämäki, M., et al. (2009). Contributions of climate, leaf-area index and leaf physiology to variation in gross primary production of six coniferous forests across Europe: a model-based analysis. Tree Physiology, 29, 621–639.CrossRefGoogle ScholarPubMed
Eagleson, O.S. (1982). Ecological optimality in water limited natural soil-vegetation systems. 1. Theory and hypothesis. Water Resources Research, 18, 325–340.CrossRefGoogle Scholar
Earl, H.J. and Davis, R.F. (2003). Effect of drought stress on leaf and whole canopy radiation use efficiency and yield of maize. Agronomy Journal, 95, 688–696.CrossRefGoogle Scholar
Earl, H.J. and Ennhali, S. (2004). Estimating photosynthetic electron transport via chlorophyll fluorometry without photosystem II light saturation. Photosynthesis Research, 82, 177–186.CrossRefGoogle ScholarPubMed
Earl, H.J. and Tollenaar, M. (1999). Using chlorophyll fluorometry to compare photosynthetic performance of commercial maize (Zea mays L.) hybrids in the field. Field Crops Res., 61, 201–210.CrossRefGoogle Scholar
Eastmond, P. and Rawsthorne, S. (1998). Comparison of the metabolic properties of plastids isolated from developing leaves or embryos of Brassica napus. Journal of Experimental Botany, 49, 1105–1111.Google Scholar
Ebbert, V., Adams, W.W., Mattoo, A.K., et al. (2005). Up-regulation of a photosystem II core protein phosphatase inhibitor and sustained D1 phosphorylation in zeaxanthin-retaining, photoinhibited needles of overwintering Douglas fir. Plant Cell Environment, 28, 232–240.CrossRefGoogle Scholar
Eckardt, N.A., Snyder, G.W., Portis, A.R. Jr., et al. (1997). Growth and photosynthesis under high and low irradiance of Arabidopsis thaliana antisense mutants with reduced ribulose-1,5-bisphosphate carboxylase/oxygenase activase content. Plant Physiology, 113, 575–586.CrossRefGoogle ScholarPubMed
Edgerton, M.D. (2009). Increasing crop productivity to meet global needs for feed, food and fuel. Plant Physiology, 149, 7–13.CrossRefGoogle ScholarPubMed
Edmondson, D.L., Badger, M.R. and Andrews, T.J. (1990). Slow inactivation of ribulose bisphosphate carboxylase during catalysis is caused by accumulation of a slow, tight-binding inhibitor at the catalytic site. Plant Physiology, 93, 1390–1397.CrossRefGoogle Scholar
Edwards, D., Edwards, D.S. and Rayner, R. (1982). The cuticle of early vascular plants and its evolutionary significance. In: eds. Cutler, D.F., Alvin, K.L., Price, C.E., The Plant Cuticle, Linnean Society Symposium Series nº10. Academic Press, London pp. 341–361.
Edwards, D., Kerp, H. and Hass, H. (1998). Stomata in early land plants: an anatomical and ecophysiological approach. Journal of Experimental Botany, 49, 255–278.CrossRefGoogle Scholar
Edwards, E.J. and Still, C.J. (2008). Climate, phylogeny and the ecological distribution of C4 grasses. Ecology Letters, 11, 266–276.CrossRefGoogle ScholarPubMed
Edwards, E.J., Still, C.J. and Donoghue, M.J. (2007a). The relevance of phylogeny to studies of global change. Trends in Ecology and Evolution, 22, 243–249.CrossRefGoogle ScholarPubMed
Edwards, G.E. and Baker, N.R. (1993). Can CO2 assimilation in maize leaves be predicted accurately from chlorophyll fluorescence analysis?Photosynthesis Research, 37, 89–102.CrossRefGoogle ScholarPubMed
Edwards, G.E., Dai, Z., Cheng, S.H., et al. (1996). Factors affecting the induction of crassulacean acid metabolism in Mesembryanthemum crystallinum. In: Crassulacean Acid Metabolism: Biochemistry, Ecophysiology and Evolution (eds Winter, K. and Smith, J.A.C.), Springer-Verlag, Berlin, Germany, pp. 119–34.Google Scholar
Edwards, J., Salib, S., Thomson, F., et al. (2007b). The impact of Phaeomoniella chlamydospora infection on the grapevine’s physiological response to water stress. Part 1: Zinfandel. Phytopathologia Mediterranea, 46, 26–37.Google Scholar
Egea, G., González-Real, M.M., Baille, A. et al. (2011). Disentangling the contributions of ontogeny and water stress to photosynthetic limitations in almond trees. Plant, Cell and Environment, 34, 962–979.CrossRefGoogle ScholarPubMed
Eggers, T. and Jones, T.H. (2000). You are what you eat … or are you?Trends in Ecology and Evolution, 15, 265–266.CrossRefGoogle ScholarPubMed
Egli, D.B. (1998). Seed Biology and the Yield of Grain Crops. CAB International, Wallingford, UK and New York, USA.
Ehara, Y. and Misawa, T. (1975). Occurrence of abnormal chloroplasts in tobacco leaves infected systemically with the ordinary strain of cucumber mosaic virus. Phytopathologische Zeitschrift, 84, 233–252.CrossRefGoogle Scholar
Ehleringer, J. (1982). The influence of water stress and temperature of leaf pubescence in Encelia farinosa. American Journal of Botany, 69, 670–675.CrossRefGoogle Scholar
Ehleringer, J. (1983). Ecophysiology of Amaranthus palmeri, a Sonoran Desert summer annual. Oecologia, 57, 107–112.CrossRefGoogle ScholarPubMed
Ehleringer, J. (1977). Adaptive value of leaf hairs in Encelia farinosa. Carnegie Institution of Washington Yearbook, 77, 413–418.Google Scholar
Ehleringer, J. and Pearcy, R.W. (1983). Variation in quantum yield for CO2 uptake among C3 and C4 plants. Plant Physiology, 73, 555–559.CrossRefGoogle Scholar
Ehleringer, J.R. (1978). Implications of quantum yield differences on distributions of C3 and C4 grasses. Oecologia, 31, 255–267.CrossRefGoogle ScholarPubMed
Ehleringer, J.R. (1985). Annuals and perennials of warm deserts. In: Physiological Ecology of North American Plant Communities (eds Chabot, B.F. and Mooney, H.A.), Chapman and Hall, New York, USA, pp. 162–80.Google Scholar
Ehleringer, J.R. and Björkman, O. (1977). Quantum yields for CO2 uptake in C3 and C4 plants. Plant Physiology, 59, 86–90.CrossRefGoogle Scholar
Ehleringer, J.R. and Björkman, O. (1978a). A comparison of photosynthetic characteristics of Encelia species possessing glabrous and pubescent leaves. Plant Physiology, 62, 185–190.CrossRefGoogle ScholarPubMed
Ehleringer, J.R. and Björkman, O. (1978b). Pubescence and leaf spectral characteristics in a desert shrub, Encelia farinosa. Oecologia, 36, 151–162.CrossRefGoogle Scholar
Ehleringer, J.R. and Cooper, T.A. (1992). On the role of orientation in reducing photoinhibitory damage in photosynthetic-twig desert shrubs. Plant Cell and Environment, 15, 301–306.CrossRefGoogle Scholar
Ehleringer, J.R. and Field, C.B. (1993). Scaling Physiological processes: From Leaf to Globe. Academic Press, San Diego, CA, USA, 1–388.Google Scholar
Ehleringer, J.R. and Mooney, H.A. (1978). Leaf hairs: effects on physiological activity and adaptive value to a desert shrub. Oecologia, 37, 183–200.CrossRefGoogle ScholarPubMed
Ehleringer, J.R., Björkman, O. and Mooney, H.A. (1976). Leaf pubescence: effects on absorptance and photosynthesis in a desert shrub. Science, 192, 376–377.CrossRefGoogle Scholar
Ehleringer, J.R., Cerling, T. E. and Helliker, B. R. (1997). C4 photosynthesis, atmospheric CO2, and climate. Oecologia, 112, 285–299.CrossRefGoogle ScholarPubMed
Ehleringer, J.R., Comstock, J.P. and Cooper, T.A. (1987). Leaf-twig carbon isotope ratio differences in photosynthetic-twig desert shrubs. Oecologia, 71, 318–320.CrossRefGoogle ScholarPubMed
Ehleringer, J.R. and Cooper, T.A. (1988). Correlations between carbon isotopes and microhabitat in desert plants. Oecologia, 76, 562–566.CrossRefGoogle ScholarPubMed
Ehleringer, J.R., Mooney, H.A., Gulmon, S.L., et al. (1980). Orientation and its consequences for Copiapoa (Cactaceae) in the Atacama Desert. Oecologia, 46, 63–67.CrossRefGoogle ScholarPubMed
Ehleringer, J.R., Mooney, H.A., Gulmon, S.L., et al. (1981). Parallel evolution of leaf pubescence in Encelia in coastal deserts of North and South America. Oecologia, 49, 38–41.CrossRefGoogle ScholarPubMed
Ehleringer, J.R., Roden, J. and Dawson, T.E. (2000). Assessing ecosystem-level water relations through stable isotope ratio analyses. In: Methods in Ecosystem Science, 181–198.CrossRefGoogle Scholar
Ehleringer, J.R., Sage, R.F., Flanagan, L.B., et al. (1991). Climate change and the evolution of C4 photosynthesis. Trends in Ecology and Evolution, 6, 95–99.CrossRefGoogle Scholar
Eichelmann, H. and Laisk, A. (2000). Cooperation of photosystems II and I in leaves as analyzed by simultaneous measurements of chlorophyll fluorescence and transmittance at 800 nm. Plant and Cell Physiology, 41, 138–147.CrossRefGoogle ScholarPubMed
Eichelmann, H., Oja, B., Rasulov, B., et al. (2004). Development of leaf photosynthetic parameters in Betula pendula Roth. leaves: correlations with Photosystem I density. Plant Biology, 6, 307–318.CrossRefGoogle ScholarPubMed
Eichelmann, H., Oja, V., Rasulov, B., et al. (2005). Adjustment of leaf photosynthesis to shade in a natural canopy: reallocation of nitrogen. Plant, Cell and Environment, 28, 389–401.CrossRefGoogle Scholar
Eilenberg, H., Hanania, U., Stein, H., et al. (1998). Characterization of rbcS genes in the fern Pteris vittata and their photoregulation. Planta, 206, 204–214.CrossRefGoogle ScholarPubMed
Eisenhut, M.S., Kahlon, D., Hasse, R., et al. (2006). The plant-like C2 glycolate cycle and the bacterial-like glycerate pathway cooperate in phosphoglycolate metabolism in cyanobacteria. Plant Physiology, 142, 333–342.CrossRefGoogle ScholarPubMed
Eissenstat, D., Graham, J.H., Syvertsen, J.P., et al. (1993). Carbon economy of sour orange in relation to mycorrhizal colonization and phosphorus status. Annals of Botany, 71, 1–10.CrossRefGoogle Scholar
Elagoz, V. and Manning, W. (2005). Responses of sensitive and tolerant bush beans (Phaseolus vulgaris L.) to ozone in open-top chambers are influenced by phenotypic differences, morphological characteristics, and the chamber environment. Environmental Pollution, 136, 371–383.CrossRefGoogle ScholarPubMed
Ellenberg, H. (1981). Ursachen des vorkommens und fehlens von sukkulenten in den trockengebieten der erde. Flora, 171, 114–169.CrossRefGoogle Scholar
Eller, B.M. and Ferrari, S. (1997). Water use efficiency of two succulents with contrasting CO2 fixation pathways. Plant Cell Environment, 20, 93–100.CrossRefGoogle Scholar
Eller, B.M. and Ruess, B.R. (1986). Modulation of CAM and water balance of Senecio medley woodii by environmental factors and age of leaf. Journal of Plant Physiology, 125, 295–309.CrossRefGoogle Scholar
Eller, B.M., Ferrari, S. and Ruess, B.R. (1992). Spatial and diel variations of water relations in leaves of the CAM-plant Senecio medley woodii. Botanica Helvetica, 102, 193–200.Google Scholar
Ellis, J.R. and Leech, R.M. (1985). Cell size and chloroplast size in relation to chloroplast replication in light-grown wheat leaves. Planta, 165, 120–125.CrossRefGoogle ScholarPubMed
Ellis, R.P., Vogel, J.C. and Fuls, A. (1980). Photosynthetic pathways and the geographical distribution of grasses in southwest Africa/Namibia. South African Journal of Science, 76, 307–314.Google Scholar
Ellison, A.M. and Gotelli, N.J. (2009). Energetics and the evolution of carnivorous plants – Darwin’s ‘most wonderful plants in the world’. Journal of Experimental Botany, 60, 19–42.CrossRefGoogle ScholarPubMed
Ellsworth, D., Reich, P., Naumburg, E., et al. (2004). Photosynthesis, carboxylation and leaf nitrogen responses of 16 species to elevated pCO2 across four free-air CO2 enrichment experiments in forest, grassland and desert. Global Change Biology, 10, 2121–2138.CrossRefGoogle Scholar
Ellsworth, D.S. and Reich, P.B. (1993). Canopy structure and vertical patterns of photosynthesis and related leaf traits in a deciduous forest. Oecologia, 96, 169–178.CrossRefGoogle Scholar
Elmore, C.D. (1980). The paradox of no correlation between leaf photosynthetic rates and crop yields. In: Predicting Photosynthesis for Ecosystems Models (eds Hesketh, J.D. and Jones, J.W), CRC Press, Boca Raton, Florida, USA, pp. 155–67.Google Scholar
Else, M.A., Coupland, D., Dutton, L., et al. (2001). Decreased root hydraulic conductivity reduces leaf water potential, initiates stomatal closure and slows leaf expansion in flooded plants of castor oil (Ricinus communis) despite diminished delivery of ABA from roots to shoots in xylem sap. Physiologia Plantarum, 111, 46–54.CrossRefGoogle Scholar
Else, M.A., Davies, W.J., Whitford, P.N., et al. (1994). Concentrations of abscisic acid and other solutes in xylem sap from root systems of tomato and castor-oil plants are distorted by wounding and variable sap flow rates. Journal of Experimental Botany, 45, 317–324.CrossRefGoogle Scholar
Eltayeb, A.E., Kawano, N., Badawi, G.H., et al. (2007). Overexpression of monodehydroascorbate reductase in transgenic tobacco confers enhanced tolerance to ozone, salt and polyethylene glycol stresses. Planta, 225, 1255–1264.CrossRefGoogle ScholarPubMed
Elvira, S., Alonso, R., Castillo, F. J., et al. (1998). On the responses of pigments and antioxidants of Pinus halepensis seedlings to Mediterranean climatic factors and long-term ozone exposure. New Phytologist, 138, 419–432.CrossRefGoogle Scholar
Emberson, L.D., Ashmore, M.R., Cambridge, H.M., et al. (2000). Modelling stomatal ozone flux across Europe. Environmental Pollution, 109, 403–413.CrossRefGoogle ScholarPubMed
Enami, I., Kitamura, M., Tomo, T., et al. (1994). Is the primary cause of thermal inactivation of oxygen evolution in spinach PSII membranes release of the extrinsic 33 kDa protein or Mn?Biochimica et Biophysica Acta, 1186, 52–58.CrossRefGoogle Scholar
Endler, J.A. (1993). The colour of light in forests and its implications. Ecological Monographs, 61, 1–27.CrossRefGoogle Scholar
Engel, N., Schmidt, M., Lütz, C., et al. (2006). Molecular identification, heterologous expression and properties of light-insensitive plant catalases. Plant, Cell and Environment, 29, 593–607.CrossRefGoogle ScholarPubMed
Engelbrecht, B.M.J., Comita, L.S., Condit, R., et al. (2007). Drought sensitivity shapes species distribution patterns in tropical forests. Nature, 447, 80–U2.CrossRefGoogle ScholarPubMed
Englemann, S.C.W., Burscheidt, J., Gowik, U., et al. (2008). The gene for the P-subunit of glycine decarboxylase from the C4 species Flaveria trinervia: analysis of transcriptional control in transgenic Flaveria bidentis (C4) and Arabidopsis (C3). Plant Physiology, 146, 1773–1785.CrossRefGoogle Scholar
Ensminger, I., Busch, F. and Huner, N.P.A. (2006). Photostasis and cold acclimation: sensing low temperature through photosynthesis. Physiologia Planarum, 126, 28–44.CrossRefGoogle Scholar
Ensminger, I., Schmidt, L. and Lloyd, J. (2008). Soil temperature and intermittent frost modulate the rate of recovery of photosynthesis in Scots pine under simulated spring conditions. New Phytologist, 177, 428–442.CrossRefGoogle ScholarPubMed
Ensminger, I., Schmidt, L., Tittmann, S., et al. (2005). Will photosynthetic gain of boreal evergreen conifers increase in response to a potentially longer growing season? In: Photosynthesis: Fundamental Aspects to Global Perspectives (eds Carpentier, R., Bruce, D. and van der Est, A.) Vol. 2, Allen Press, Lawrence, KS, USA, pp 976–978.Google Scholar
Ensminger, I., Sveshnikov, D., Campbell, D.A., et al. (2004). Intermittent low temperatures constrain spring recovery of photosynthesis in boreal Scots pine forests. Global Change Biology, 10, 995–1008.CrossRefGoogle Scholar
Epron, D. and Dreyer, E. (1992). Effects of severe dehydration on leaf photosynthesis in Quercus petraea (Matt.) Liebl.: photosystem II efficiency, photochemical and non-photochemical quenching and electrolyte leakage. Tree Physiology, 10, 273–285.CrossRefGoogle Scholar
Epron, D., Dreyer, E. and Bréda, N. (1992). Photosynthesis of oak trees [Quercus petraea (Matt) Liebl.] during drought under field conditions: diurnal course of net CO2 assimilation and photochemical efficiency of photosystem II. Plant, Cell and Environment, 15, 809–820.CrossRefGoogle Scholar
Epron, D., Godard, G., Cornic, G., et al. (1995). Limitation of net CO2 assimilation rate by internal resistances to CO2 transfer in the leaves of two tree species (Fagus sylvatica and Castanea sativa Mill). Plant, Cell and Environment, 18, 43–51.CrossRefGoogle Scholar
Epstein, H.E., Lauenroth, W.K.,. Burke, I.C., et al. (1997). Productivity patterns of C3 and C4 functional types in the U.S. Great Plains. Ecology, 78, 722–731.Google Scholar
Epstein, S. and Mayeda, T. (1953). Variation of 18O content of waters from natural sources. Geochimica et Cosmochimica Acta, 4, 213–224.CrossRefGoogle Scholar
Ergova, E.A., Bukhov, N.G., Heber, U., et al. (2003). Effect of the pool size of stromal reductants on the alternative pathway of electron transfer to photosystem I in chloroplasts of intact leaves. Russian Journal of Plant Physiology, 50, 431–440.Google Scholar
Esau, K. (1968). Viruses in Plant Hosts. Form, Distribution and Pathologic Effect. Wisconsin Press, Madison, Milwaukee.Google Scholar
Escalona, J.M., Flexas, J., Bota, J., et al. (2003). From leaf photosynthesis to grape yield: influence of soil water availability. Vitis, 42, 57–64.Google Scholar
Eschrich, W., Fromm, J. and Essiamah, S. (1988). Mineral partitioning in the phloem during autumn senescence of beech leaves. Trees, 2, 73–83.CrossRefGoogle Scholar
Escoubas, J., Lomas, M., LaRoche, J., et al. (1995). Light intensity regulation of cab gene transcription is signaled by the redox state of the plastoquinone pool. Proceedings of the National Academy of Sciences USA, 92, 10237–10241.CrossRefGoogle ScholarPubMed
Esfeld, P., Siebke, K., Wacker, I., et al. (1995). Local defence-related shift in the carbon metabolism in chickpea leaves induced by a fungal pathogen. In: Photosynthesis: From Light to Biosphere (ed. Mathis, P.) Vol. 5, Kluwer Academic Publisher, Dordrecht, Netherlands, pp. 663–666.Google Scholar
Esler, K.J. and Rundel, P.W. (1999). Comparative patterns of phenology and growth form diversity in two winter rainfall deserts: the Succulent Karoo and the Mojave Desert ecosystems. Plant Ecology, 142, 97–104.CrossRefGoogle Scholar
Espinoza, C., Vega, A., Medina, C., et al. (2007). Gene expression associated with compatible viral diseases in grapevine cultivars. Functional and Integrative Genomics, 7, 95–110.CrossRefGoogle ScholarPubMed
Esteban, R., Fernández-Marín, B., Becerril, J.M., et al. (2008). Photoprotective implications of leaf variegation in E. dens-canis L. and P. officinalis L. Journal of Plant Physiology, 165, 1255–1263.CrossRefGoogle Scholar
Esteban, R., Olano, J.M., Castresana, J., et al. (2009). Distribution and evolutionary trends of photoprotective isoprenoids (xanthophylls and tocopherols) within the plant kingdom. Physiologia Plantarum, 135, 379–389.CrossRefGoogle ScholarPubMed
Ethier, G.J. and Livingston, N.J. (2004). On the need to incorporate sensitivity to CO2 transfer conductance into the Farquhar-von Caemmerer-Berry leaf photosynthesis model. Plant, Cell and Environment, 27, 137–153.CrossRefGoogle Scholar
Ethier, G.J., Livingston, N.J., Harrison, D.L., et al. (2006). Low stomatal and internal conductance to CO2 versus Rubisco deactivation as determinants of the photosynthetic decline of ageing evergreen leaves. Plant, Cell and Environment, 29, 2168–2184.CrossRefGoogle ScholarPubMed
Etiope, G. and Ciccioli, P. (2009). Earth’s degassing: a missing ethane and propane source. Science, 323, 478.CrossRefGoogle ScholarPubMed
Evain, S., Flexas, J. and Moya, I. (2004). A new instrument for passive remote sensing: 2. Measurement of leaf and canopy reflectance changes at 531 nm and their relationship with photosynthesis and chlorophyll fluorescence. Remote Sensing of Environment, 91, 175–185.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. (1986). A quantitative analysis of light distribution between the two photosystems, considering variation in both the relative amounts of the chlorophyll-protein complexes and the spectral quality of light. Photochemistry and Photobiophysics, 10, 135–147.Google Scholar
Evans, J.R. (1989). Photosynthesis and nitrogen relationships in leaves of C3 plants. Oecologia, 78, 9–19.CrossRefGoogle ScholarPubMed
Evans, J.R. (1999). Leaf anatomy enables more equal access to light and CO2 between chloroplasts. New Phytologist, 143, 93–104.CrossRefGoogle Scholar
Evans, J.R. (2009). Potential errors in electron transport rates calculated from chlorophyll fluorescence as revealed by a multilayer leaf model. Plant and Cell Physiology, 50, 698–706.CrossRefGoogle ScholarPubMed
Evans, J.R. and Poorter, H. (2001). Photosynthetic acclimation of plants to growth irradiance: the relative importance of specific leaf area and nitrogen partitioning in maximizing carbon gain. Plant, Cell and Environment, 24, 755–767.CrossRefGoogle Scholar
Evans, J.R. and Seemann, J.R. (1989). The allocation of protein nitrogen in the photosynthetic apparatus: costs, consequences, and control. In: Photosynthesis. Proceedings of the C.S. French Symposium on Photosynthesis held in Stanford, California, July 17–23, 1988 (ed. Briggs, W.R.), Alan R. Liss, New York, USA. pp. 183–205.Google Scholar
Evans, J.R. and Terashima, I. (1987). Effects of nitrogen nutrition on electron transport components and photosynthesis in spinach. Australian Journal of Plant Physiology, 14, 59–68.CrossRefGoogle Scholar
Evans, J.R. and Terashima, I. (1988). Photosynthetic characteristics of spinach leaves grown with different nitrogen treatments. Plant and Cell Physiology, 29, 157–165.Google Scholar
Evans, J.R. and Vellen, L. (1996). Wheat cultivars differ in transpiration efficiency and CO2 diffusion inside their leaves. In: Crop Research in Asia: Achievements and Perspectives. Proceedings of the 2nd Asian Crop Science Conference “Toward the Improvement of Food Production under Steep Population Increase and Global Environment Change” 21–23 August, 1995, the Fukui Prefectural University, Fukui, Japan (eds Ishii, R. and Horie, T.), Crop Science Society of Japan, Asian Crop Science Association (ACSA), Fukui, Japan, pp. 326–329.Google Scholar
Evans, J.R. and Vogelmann, T.C. (2003). Profiles of 14C fixation through spinach leaves in relation to light absorption and photosynthetic capacity. Plant, Cell and Environment, 26, 547–560.CrossRefGoogle Scholar
Evans, J.R. and Vogelmann, T.C. (2006). Photosynthesis within isobilateral Eucalyptus pauciflora leaves. The New Phytologist, 171, 171–182.CrossRefGoogle ScholarPubMed
Evans, J.R., Jakobsen, I. and Ögren, E. (1993). Photosynthetic light-response curves. 2. Gradients of light absorption and photosynthetic capacity. Planta, 189, 191–200.CrossRefGoogle Scholar
Evans, J.R., Kaldenhoff, R., Genty, B., et al. (2009). Resistances along the CO2 diffusion pathway inside leaves. Journal of Experimental Botany, 60, 2235–2248.CrossRefGoogle ScholarPubMed
Evans, J.R. and Loreto, F. (2000). Acquisition and diffusion of CO2 in higher plant leaves. In: Photosynthesis: Physiology and Metabolism (eds Leegood, R.C., Sharkey, T.D. and von Caemmerer, S.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 321–351.Google Scholar
Evans, J.R., Sharkey, T.D., Berry, J.A., et al. (1986). Carbon isotope discrimination measured concurrently with gas exchange to investigate CO2 diffusion in leaves of higher plants. Australian Journal of Plant Physiology, 13, 281–292.CrossRefGoogle Scholar
Evans, J.R., von Caemmerer, S., Setchell, B.A., et al. (1994). The relationship between CO2 transfer conductance and leaf anatomy in transgenic tobacco with a reduced content of Rubisco. Australian Journal of Plant Physiology, 21, 475–495.CrossRefGoogle Scholar
Evans, L.T. (1975). The physiological basis of crop yield. In: Crop Physiology (ed. Evans, L.T.), Cambridge University Press, Cambridge, UK, pp. 327–335.Google Scholar
Evans, L.T. (1993). Crop Evolution, Adaptation and Yield. Cambridge University Press, Cambridge, UK.Google Scholar
Eversman, S. and Sigal, L.L. (1984). Ultrastructural effects of Peroxyacetyl Nitrate (PAN) on two lichen species. The Bryologist, 87, 112–118.CrossRefGoogle Scholar
Ewe, S. and Sternberg, L.D.S.L. (2005). Growth and gas exchange responses of Brazilian pepper (Schinus terebinthifolius) and native South Florida species to salinity. Trees, 19, 119–128.CrossRefGoogle Scholar
Ewers, B.E., Oren, R., Phillips, N., et al. (2001). Mean canopy stomatal conductance responses to water and nutrient availabilities in Picea abies and Pinus taeda. Tree Physiology, 21, 841–850.CrossRefGoogle ScholarPubMed
Ewers, F.W. (1982). Secondary growth in needle leaves of Pinus longaeva (bristlecone pine) and other conifers: quantitative data. American Journal of Botany, 69, 1552–1559.CrossRefGoogle Scholar
Eyles, A., Pinkard, E.A., O’Grady, A.P., et al. (2009). Role of corticular photosynthesis following defoliation in Eucalyptus globules. Plant, Cell Environment, 32, 1004–1014.CrossRefGoogle Scholar
Eyster, C., Brown, T., Tanner, H., et al. (1958). Manganese requirement with respect to growth, Hill reaction and photosynthesis. Plant Physiology, 338, 235–241.CrossRefGoogle Scholar
Ezcurra, E., Montaña, C. and Arizaga, S. (1991). Architecture, light interception, and distribution of Larrea species in the Monte Desert, Argentina. Ecology, 72, 23–34.CrossRefGoogle Scholar
Fader, G.M. and Koller, H.R. (1985). Seed growth rate and carbohydrate pool sizes of the soybean fruit. Plant Physiology, 79, 663–666.CrossRefGoogle ScholarPubMed
Fahn, A. and Cutler, D.F. (1992). Xerophytes. Gebruder Borntraeger, Berlin and Sttutgart.Google Scholar
Falge, E., Baldocchi, D., Olson, R., et al. (2001). Gap filling strategies for defensible annual sums of net ecosystem exchange. Agricultural and Forest Meteorology, 107, 43–69.CrossRefGoogle Scholar
Falge, E., Ryel, R.J., Alsheimer, M., et al. (1997). Effects of stand structure and physiology on forest gas exchange: a simulation study for Norway spruce. Trees, 11, 436–448.CrossRefGoogle Scholar
Falge, E., Tenhunen, J.D., Ryel, R., et al. (2000). Modelling age- and density related gas exchange of Picea abies canopies in the Fichtelgebirge, Germany. Annals of Forest Science, 57, 229–243.CrossRefGoogle Scholar
Falk, S., Maxwell, D.P., Laudenbach, D.E., et al. (1996). Photosynthetic adjustment to temperature. In: Advances in Photosynthesis Vol.5: Photosynthesis and the Environment (ed. Baker, N.R.) Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 367–385.Google Scholar
Falkowski, P.G. and Chen, Y.B. (2003). Photoacclimation of light harvesting systems in eukaryotic algae. In: Advances in Photosynthesis and Respiration: Light Harvesting Antennas in Photosynthesis (eds Green, B.R. and Parson, W.W.) Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 423–447.Google Scholar
Falkowski, P.G. and Isozaki, Y. (2008). The story of O2. Science, 322, 540–542.CrossRefGoogle ScholarPubMed
Falster, D.S. and Westoby, M. (2003). Leaf size and angle vary widely across species: what consequences for light interception?New Phytologist, 158, 509–525.CrossRefGoogle Scholar
Fan, L.-M, Zhao, Z. and Assmann, S.M. (2004). Guard cells: a dynamic signaling model. Current Opinion in Plant Biology, 7, 537–546.CrossRefGoogle ScholarPubMed
Fan, S.H. and Grossnickle, S.C. (1998). Comparisons of gas exchange parameters and shoot water relations of interior spruce (Picea glauca (Moench)Voss × Picea engelmannii Parry ex Engelm.) clones under repeated soil drought. Canadian Journal of Forest Research-Revue Canadienne De Recherche Forestiere, 28, 820–830.CrossRefGoogle Scholar
Fan, S.M., Wofsy, S.C., Bakwin, P.S., et al. (1990). Atmosphere-biosphere exchange of CO2 and O3 in the central-Amazon-forest. Journal of Geophysical Research-Atmospheres, 95, 16851–16864.CrossRefGoogle Scholar
Fangmeier, A., Bender, J., Weigel, H.J., et al. (2002). Effects of pollutant mixtures. In: Air Pollution and Plant Life (eds Bell, J.N.B. and Treshow, M.) 2nd edn. John Wiley and Sons, Chichester, USA, pp. 251–272.Google Scholar
FAO (2005). Report of the 22nd session of the International Poplar Commission and 42nd session of its Executive Committee, Santiago, Chile, 28 November – 2 December 2004.
Farage, P.K. (1996). The effect of ozone fumigation over one season on photosynthetic processes of Quercus robur seedlings. New Phytologist, 134, 279–285.CrossRefGoogle Scholar
Farage, P.K. and Long, S.P. (1991). The occurrence of photoinhibition in an over-wintering crop of oilseed rape (Brassica-Napus L.) and its correlation with changes in crop growth. Planta, 185, 279–286.CrossRefGoogle Scholar
Farage, P.K., Blowers, D., Long, S.P., et al. (2006). Low growth temperatures modify the efficiency of light use by photosystem II for CO2 assimilation in leaves of two chilling-tolerant C4 species, Cyperus longus L. and Miscanthus × giganteus. Plant, Cell and Environment, 29, 720–728.CrossRefGoogle ScholarPubMed
Farage, P.K., McKee, I. and Long, S.P. (1998). Does a low nitrogen supply necessarily lead to acclimation of photosynthesis to elevated CO2?Plant Physiology, 118, 573–580.CrossRefGoogle ScholarPubMed
Fares, S., Barta, C., Ederli, L., et al. (2006). Impact of high ozone on isoprene emission and some anatomical and physiological parameters of developing Populus alba leaves directly or indirectly exposed to the pollutant. Physiologia Plantarum, 128, 456–465.CrossRefGoogle Scholar
Farmer, A. (2002). Effects of particulates. In: Air Pollution and Plant Life (eds Bell, J.N.B. and Treshow, M.), 2nd edn. John Wiley and Sons, Chichester, USA, pp. 187–200.Google Scholar
Farmer, A.M. (1993). The effects of dust on vegetation: a review. Environmental Pollution, 79, 63–75.CrossRefGoogle ScholarPubMed
Farmer, A.M. (1996). Carbon uptake by roots. In: Plant Roots: The Hidden Half (eds Waisel, Y., Eshel, A. and Kafkaki, U.), Marcel Dekker, New York, USA, pp. 679–687.Google Scholar
Farnsworth, E.J. and Ellison, A.M. (2008). Prey availability directly affects physiology, growth, nutrient allocation and scaling relationships among leaf traits in 10 carnivorous plant species. Journal of Ecology, 96, 213–221.Google Scholar
Farque, L., Sinoquet, H. and Colin, F. (2001). Canopy structure and light interception in Quercus petraea seedlings in relation to light regime and plant density. Tree Physiology, 21, 1257–1267.CrossRefGoogle ScholarPubMed
Farquhar, G.D. (1980). Carbon isotope discrimination by plants and the ratio of intercellular and atmospheric CO2 concentrations. In: Carbon Dioxide and Climate: Australian Research (ed. Pearman, G.I.), Australian Academy of Science, Canberra, Australia, pp. 105–110.Google Scholar
Farquhar, G.D. (1983). On the nature of carbon isotope discrimination in C4 plants. Australian Journal of Plant Physiology, 10, 205–226.CrossRefGoogle Scholar
Farquhar, G.D. (1989). Models of integrated photosynthesis of cells and leaves. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 323, 357–367.CrossRefGoogle Scholar
Farquhar, G.D. and Cernusak, L.A. (2005). On the isotopic composition of leaf water in the non-steady state. Functional Plant Biology, 32, 293–303.CrossRefGoogle Scholar
Farquhar, G.D. and Gan, K.S. (2003). On the progressive enrichment of the oxygen isotopic composition of water along a leaf. Plant, Cell and Environment, 26, 1579–1597.Google ScholarPubMed
Farquhar, G.D. and Lloyd, J. (1993). Carbon and oxygen isotope effects in the exchange of carbon dioxide between terrestrial plants and the atmosphere. In: Stable Isotopes and Plant Carbon-water Relations (eds Ehleringer, J.R., Hall, A.E. and Farquhar, G.D.), Academic Press, San Diego, CA, USA, pp. 47–70.Google Scholar
Farquhar, G.D. and Richards, R.A. (1984). Isotopic composition of plant carbon correlates with water-use efficiency of wheat genotypes. Australian Journal of Plant Physiology, 11, 359–552.CrossRefGoogle Scholar
Farquhar, G.D. and Sharkey, T.D. (1982). Stomatal conductance and photosynthesis. Annual Review of Plant Physiology, 33, 317–345.CrossRefGoogle Scholar
Farquhar, G.D. and von Caemmerer, S. (1982). Modeling of photosynthetic responses to environmental conditions. In: Encyclopedia of Plant Physiology (New Series) (eds Lange, O.L., Nobel, P.S., Osmond, C.B. and Ziegler, H.), Springer-Verlag, Berlin, Germany, pp. 549–587.Google Scholar
Farquhar, G.D. and Wong, S.C. (1984). An empirical model of stomatal conductance. Australian Journal of Plant Physiology, 11, 191–210.CrossRefGoogle Scholar
Farquhar, G.D., Cernusak, L.A. and Barnes, B. (2007). Heavy water fractionation during transpiration. Plant Physiology, 143, 11–18.CrossRefGoogle ScholarPubMed
Farquhar, G.D., Ehleringer, J.R. and Hubick, K.T. (1989). Carbon isotope discrimination and photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology, 40, 503–537.CrossRefGoogle Scholar
Farquhar, G.D., Firth, P.M., Wetselaar, R., et al. (1980c). On the gaseous exchange of ammonia between leaves and the environment: determination of the ammonia compensation point. Plant Physiology, 66, 710–714.CrossRefGoogle ScholarPubMed
Farquhar, G.D., Lloyd, J., Taylor, J.A., et al. (1993). Vegetation effects on the isotope composition of oxygen in the atmospheric CO2. Nature, 363, 439–443.CrossRefGoogle Scholar
Farquhar, G.D., O’Leary, M.H. and Berry, J.A. (1982).On the relationship between carbon isotope discrimination and the intercellular carbon dioxide concentration in leaves. Australian Journal of Plant Physiology, 9, 121–137.CrossRefGoogle Scholar
Farquhar, G.D., Schulze, E.D. and Küuppers, M. (1980b). Responses to humidity by stomata of Nicotiana glauca L. and Corylus avellana L. are consistent with the optimization of carbon-dioxide uptake with respect to water-loss. Australian Journal of Plant Physiology, 7, 315–327.CrossRefGoogle Scholar
Farquhar, G.D., von Caemmerer, S. and Berry, J.A. (1980a). A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta, 149, 78–90.CrossRefGoogle Scholar
Farquhar, G.D., von Caemmerer, S. and Berry, J.A. (2001). Models of photosynthesis. Plant Physiology, 125, 42–45.CrossRefGoogle ScholarPubMed
Farris, F. and Strain, B.R. (1978). The effects of water-stress on leaf H218O enrichment. Radiation and Environmental Biophysics, 15, 167–202.CrossRefGoogle Scholar
Fatemy, F., Trinder, P.K.E., Wingfield, J.N., et al. (1985). Effect of Globodera rostochiensis water stress and oxogeneris abscisic acid on stomatal function and water use of Cara and Pentland Dell potato plant. Review of Nematology, 8, 249–255.Google Scholar
Favali, M.A., Pellegrini, S., and Bassi, M. (1975). Ultrastructural alterations induced by rice tungro virus in rice leaves. Virology, 66, 502–507.CrossRefGoogle ScholarPubMed
Febrero, A., Fernández, A., Fernández, S., et al. (1998). Yield, carbon isotope discrimination, canopy reflectance and cuticular conductance of barley isolines of differing glaucosness. Journal of Experimental Botany, 49, 1575–1581.CrossRefGoogle Scholar
Federer, C.A. and Tanner, C.B. (1966). Spectral distribution of light in the forest. Ecology, 47, 555–560.CrossRefGoogle Scholar
Feierabend, J., Streb, P., Schmidt, M., et al. (1996). Expression of catalase and its relation to light stress and stress tolerance. In: Physical Stresses in Plants (eds Grillo, S. and Leone, A.) Springer, Berlin, Heidelberg, Germany, pp. 223–234.Google Scholar
Feild, T.S. and Arens, N.C. (2007). The ecophysiology of early angiosperms. Plant, Cell and Environment, 30, 291–309.CrossRefGoogle ScholarPubMed
Feild, T.S., Lee, D.W. and Holbrook, M.M. (2001). Why leaves turn red in autumn. The role of anthocyanins in senescing leaves of red-osier dogwood. Plant Physiology, 127, 566–574.CrossRefGoogle ScholarPubMed
Felzer, B., Cronin, T., Reilly, J.M., et al. (2007). Impacts of ozone on trees and crops. C. R. Geoscience, 339, 784–798.CrossRefGoogle Scholar
Feng, L., Han, H., Liu, G., et al. (2007a). Overexpression of sedoheptulose-1,7-bisphosphatase enhances photosynthesis and growth under salt stress in transgenic rice plants. Functional Plant Biology, 34, 822–834.CrossRefGoogle Scholar
Feng, L., Wang, K., Li, Y., et al. (2007b). Overexpression of SBPase enhances photosynthesis against high temperature stress in transgenic rice plants. Plant Cell Report, 26, 1635–1646.CrossRefGoogle ScholarPubMed
Feng, Y.L., Lei, Y.B., Wang, R.F., et al. (2009). Evolutionary tradeoffs for nitrogen allocation to photosynthesis versus cell walls in an invasive plant. Proceedings of the National Academy of Sciences USA, 106, 1853–1856.CrossRefGoogle Scholar
Feng, Z., Jin, M., Zhang, F., et al. (2003). Effects of ground-level ozone (O3) pollution on the yields of rice and winter wheat in the Yangtze River Delta. Journal of Environmental Sciences-China, 15, 360–362.Google Scholar
Fenton, J.M. and Crofts, A.R. (1990). Computer aided fluorescence imaging of photosynthetic systems: application of video imaging to the study of fluorescence induction in green plants and photosynthetic bacteria. Photosynthesis Research, 26, 59–66.Google Scholar
Fereres, E., Cruz-Romero, G., Hoffman, G.J., et al. (1979). Recovery of orange trees following severe water stress. Journal of Applied Ecology, 16, 833–842.CrossRefGoogle Scholar
Fernandes, G.W., deMattos, E.A., Franco, A.C., et al. (1998). Influence of the parasite Pilostyles ingai (Rafflesiaceae) on some physiological parameters of the host plant, Mimosa naguirei (Mimosaceae). Botanica Acta, 111, 51.CrossRefGoogle Scholar
Fernandez, A.P. and Strand, A. (2008). Retrograde signalling and plant stress: plastid signals initiate cellular stress responses. Current Opinion in Plant Biology, 11, 509–513.CrossRefGoogle Scholar
Ferrar, P.J. and Osmond, C.B. (1986). Nitrogen supply as a factor influencing photoinhibition and photosynthetic acclimation after transfer of shade-grown Solanum dulcamara to bright light. Planta, 168, 563–570.CrossRefGoogle ScholarPubMed
Ferrar, P.J., Slatyer, R.O. and Vranjic, J.A. (1989). Photosynthetic temperature acclimation in Eucalyptus species from diverse habitats, and a comparison with Nerium oleander. Australian Journal of Plant Physiology, 16, 199–217.CrossRefGoogle Scholar
Ferraro, F., Castagna, A., Soldatini, G.F., et al. (2003). Tomato (Licopersicon esculentum M.) T3238FER and T3238fer genotypes. Influence of different iron concentrations on thylakoid pigment and protein composition. Plant Science, 164, 783–792.CrossRefGoogle Scholar
Ferrio, J.P., Cuntz, M., Offermann, C., et al. (2009). Effect of water availability on leaf water isotopic enrichment in beech seedlings shows limitations of current fractionation models. Plant, Cell and Environment, 32, 1285–1296.CrossRefGoogle ScholarPubMed
Fetene, M., Nauke, P., Lüttge, U., et al. (1997). Photosyntesis and photoinhibition in a tropical alpine giant rosette plant, Lobelia rhynchopetalum. New Phytologist, 137, 453–461.CrossRefGoogle Scholar
Fey, V., Wagner, R., Bräutigam, K., et al. (2005a). Retrograde plastid redox signals in the expression of nuclear genes for chloroplast proteins Arabidopsis thaliana. The Journal of Biological Chemistry, 280, 5318–5328.CrossRefGoogle ScholarPubMed
Fey, V., Wagner, R., Bräutigam, K., et al. (2005b). Photosynthetic redox control of nuclear gene expression. Journal Experimental Botany, 56, 1491–1498.CrossRefGoogle ScholarPubMed
Field, C. (1983). Allocating leaf nitrogen for the maximization of carbon gain: leaf age as a control on the allocation program. Oecologia, 56, 341–347.CrossRefGoogle ScholarPubMed
Field, C. and Mooney, H.A. (1986). The photosynthesis – nitrogen relationship in wild plants. In: On the Economy of Plant Form and Function. Proceedings of the Sixth Maria Moors Cabot Symposium, “Evolutionary constraints on primary productivity: adaptive patterns of energy capture in plants”, Harvard Forest, August 1983 (ed. Givnish, T.J.). Cambridge University Press, Cambridge, UK, pp. 25–55.Google Scholar
Field, C., Merino, J. and Mooney, H.A. (1983). Compromises between water-use efficiency and nitrogen-use efficiency in 5 species of California evergreens. Oecologia, 60, 384–389.CrossRefGoogle Scholar
Filella, I. and Peñuelas, J. (1994). The red edge position and shape as indicators of plant chlorophyll content, biomass and hydric status. International Journal of Remote Sensing, 15, 1459–1470.CrossRefGoogle Scholar
Filella, I., Amaro, T., Araus, J.L., et al. (1996). Relationship between photosynthetic radiation-use-efficiency of barley canopies and the photochemical reflectance index (PRI). Physiologia Plantarum, 96, 211–216.CrossRefGoogle Scholar
Filippou, M., Fasseas, C. and Karabourniotis, G. (2007). Photosynthetic characteristics of olive tree (Olea europaea) bark. Tree Physiology, 27, 977–984.CrossRefGoogle ScholarPubMed
Finazzi, G. (2002). Redox-coupled proton pumping activity in cytochrome b6f, as evidenced by the pH dependence of electron transfer in whole cells of Chlamydomonas reinhardtii. Biochemistry, 41, 7475–7482.CrossRefGoogle Scholar
Finazzi, G., Jonson, G.N., Dall’Osto, L., et al. (2004). A zeaxanthin-independent nonphotochemical quenching mechanism localized in the photosystem II core complex. Proceedings of the National Academy of Sciences USA, 101, 12375–12380.CrossRefGoogle ScholarPubMed
Finazzi, G., Zito, F., Barbagallo, R.P., et al. (2001). Contrasted effects of inhibitors of cytochrome b6f complex on state transitions in Chlamydomonas reinhardtii: the role of Qo site occupancy in LHCII kinase activation. Journal of Biological Chemistry, 276, 9770–9774.CrossRefGoogle ScholarPubMed
Finn, M.W. and Tabita, F.R. (2004). Modified pathway to synthesize ribulose 1,5-bisphosphate in methanogenic archaea. The Journal of Bacteriology, 186, 6360–6366.CrossRefGoogle ScholarPubMed
Finnigan, J. (2006). The storage term in eddy flux calculations. Agriculture and Forest Meteorology, 136, 108–113.CrossRefGoogle Scholar
Finnigan, J.J., Clement, R., Malhi, Y., et al. (2003). A re-evaluation of long-term flux measurement techniques part I: averaging and coordinate rotation. Boundary Layer Meteorology, 107, 1–48.CrossRefGoogle Scholar
Finzi, A.C., Norby, R.J., Calfapietra, C., et al. (2007). Increases in nitrogen uptake rather than nitrogen-use efficiency support higher rates of temperate forest productivity under elevated CO2. Proceedings of the National Academy of Sciences of the USA, 104, 14014–14019.CrossRefGoogle ScholarPubMed
Fischer, B.B., Krieger-Liszkay, A., Hideg, E., et al. (2007). Role of singlet oxygen in chloroplast to nucleus retrograde signaling in Chlamydomonas reinhardtii. FEBS letters, 581, 5555–5560.CrossRefGoogle ScholarPubMed
Fischer, K. and Kluge, M. (1984). Studies on carbon flow in crassulacean acid metabolism during the initial light period. Planta, 160, 121–138.CrossRefGoogle ScholarPubMed
Fischer, M. and Kaldenhoff, R. (2008). On the pH regulation of plant aquaporins. Journal of Biological Chemistry, 283, 33889–33892.CrossRefGoogle Scholar
Fiscus, E., Booker, F. and Burkey, K. (2005). Crop responses to ozone: uptake, modes of action, carbon assimilation and partitioning. Plant, Cell and Environment, 28, 997–1011.CrossRefGoogle Scholar
Fiscus, E., Reid, C., Miller, J., et al. (1997). Elevated CO2 reduces O3 flux and O3-induced yield losses in soybeans: possible implications for elevated CO2 studies. Journal of Experimental Botany, 48, 307–313.CrossRefGoogle Scholar
Fiscus, E.L., Booker, F.L. and Burkey, K.O. (2005). Crop responses to ozone: uptake, modes of action, carbon assimilation and partitioning. Plant, Cell and Environment, 28, 997–1011.CrossRefGoogle Scholar
Fish, D.A. and Earl, H.J. (2009). Water-use efficiency is negatively correlated with leaf epidermal conductance in cotton (Gossypium spp.). Crop Science, 49, 1409–1415.CrossRefGoogle Scholar
Flaig, H. and Mohr, H. (1992). Assimilation of nitrate and ammonium by the Scots pine (Pinus sylvestris) seedling under conditions of high nitrogen supply. Physiologia Plantarum, 84, 568–576.CrossRefGoogle Scholar
Flamant, P.H., Loth, C., Bruneau, D., et al. (2005) FACTS: Future Atmospheric Carbon dioxide Testing from Space, Final report, available from ESA on request.
Flanagan, L.B., Bain, J.F. and Ehleringer, J.R. (1991b). Stable oxygen and hydrogen isotope composition of leaf water in C3 and C4 plant species under field conditions. Oecologia, 88, 394–400.CrossRefGoogle ScholarPubMed
Flanagan, L.B., Comstock, J.P. and Ehleringer, J.R. (1991a). Comparison of modeled and observed environmental influences on the stable oxygen and hydrogen isotope composition of leaf water in Phaseolus vulgaris L. Plant Physiology, 96, 588–596.CrossRefGoogle ScholarPubMed
Fleck, I., Hogan, K. P., Llorens, L., et al. (1998). Photosynthesis and photoprotection in Quercus ilex resprouts after fire. Tree Physiology, 18, 607–614.CrossRefGoogle ScholarPubMed
Fleck, I., Peña-Rojas, K. and Aranda, X. (2010). Mesophyll conductance to CO2 and leaf morphological characteristics under drought stress during Quercus ilex L. resprouting. Annals of Forest Science, 67, 308.CrossRefGoogle Scholar
Fleck, S., Niinemets, Ü., Cescatti, A., et al. (2003). Three-dimensional lamina architecture alters light harvesting efficiency in Fagus: a leaf-scale analysis. Tree Physiology, 23, 577–589.CrossRefGoogle ScholarPubMed
Fleischer, A., Titel, C., Ehwald, R. (1998). The boron requirement and cell wall properties of growing and stationary suspension-cultured Chenopodium album L. cells. Plant Physiology, 117, 1401–1410.CrossRefGoogle ScholarPubMed
Fleischmann, F., Koehl, J., Portz, R., et al. (2005). Physiological changes of Fagus sylvatica seedlings infected with Phytophthora citricola and the contribution of its elicitin “Citricolin” to pathogenesis. Plant Pathology, 7, 650–658.Google Scholar
Fleming, A.J. (2005). The control of leaf development. Tansley review. New Phytologist, 166, 9–20.CrossRefGoogle Scholar
Flexas, J. and Medrano, H. (2002a). Drought-inhibition of photosynthesis in C3 plants: stomatal and non-stomatal limitations revisited. Annals of Botany, 89, 183–189.CrossRefGoogle ScholarPubMed
Flexas, J. and Medrano, H. (2002b). Energy dissipation in C3 plants under drought. Functional Plant Biology, 29, 1209–1215.CrossRefGoogle Scholar
Flexas, J., Badger, M., Chow, W.S., et al. (1999b). Analysis of the relative increase in photosynthetic O2 uptake when photosynthesis in grapevine leaves is inhibited following low night temperatures and/or water stress. Plant Physiology, 121, 675–684.CrossRefGoogle ScholarPubMed
Flexas, J., Barón, M., Bota, J., et al. (2009). Photosynthesis limitations during water stress acclimation and recovery in the drought-adapted Vitis hybrid Richter-110 (V. berlandieri × V. rupestris). Journal of Experimental Botany, 60, 2361–2377.CrossRefGoogle Scholar
Flexas, J., Bota, J., Cifre, J., et al. (2004b). Understanding down-regulation of photosynthesis under water stress: future prospects and searching for physiological tools for irrigation management. Annals of Applied Biology, 144, 273–283.CrossRefGoogle Scholar
Flexas, J., Bota, J., Escalona, J.M., et al. (2002a). Effects of drought on photosynthesis in grapevines under field conditions: an evaluation of stomatal and mesophyll limitations. Functional Plant Biology, 29, 461–471.CrossRefGoogle Scholar
Flexas, J., Bota, J., Galmés, J., et al. (2006c). Keeping a positive carbon balance under adverse conditions: responses of photosynthesis and respiration to water stress. Physiologia Plantarum, 127, 343–352.CrossRefGoogle Scholar
Flexas, J., Bota, J., Loreto, F., et al. (2004a). Diffusive and metabolic limitations to photosynthesis under drought and salinity in C3 plants. Plant Biology, 6, 269–279.CrossRefGoogle Scholar
Flexas, J., Briantais, J.M., Cerovic, ZG., et al. (2000) Steady state and maximum chlorophyll fluorescence responses to water stress in grapevine leaves: a new remote sensing system. Remote Sensing of the Environment, 73, 283–297.CrossRefGoogle Scholar
Flexas, J., Díaz-Espejo, A., Berry, J.A., et al. (2007b). Analysis of leakage in IRGA’s leaf chambers of open gas exchange systems: quantification and its effects in photosynthesis parameterization. Journal Experimental Botany, 58, 1533–1543.CrossRefGoogle ScholarPubMed
Flexas, J., Diaz-Espejo, A., Galmés, J., et al. (2007a). Rapid variations of mesophyll conductance in response to changes in CO2 concentration around leaves. Plant, Cell and Environment, 30, 1284–1298.CrossRefGoogle ScholarPubMed
Flexas, J., Escalona, J.M. and Medrano, H. (1999a). Water stress induces different levels of photosynthesis and electron transport rate regulations in grapevines. Plant, Cell and Environment, 22, 39–48.CrossRefGoogle Scholar
Flexas, J., Escalona, J.M., Evain, S., et al. (2002b). Steady state chlorophyll fluorescence (Fs) measurements as a tool to follow varations of net CO2 assimilation and stomatal conductance during water-stress in C3 plants. Physiologia Plantarum, 114, 231–240.CrossRefGoogle Scholar
Flexas, J., Galmés, J., Gallé, A., et al. (2010). Improving water use efficiency in grapevines: potential physiological targets for biotechnological improvement. Australian Journal of Grape and Wine Research, 16, 106–121.CrossRefGoogle Scholar
Flexas, J., Gulías, J. and Medrano, H. (2003). Leaf photosynthesis in Mediterranean vegetation. In: Advances in Plant Physiology Vol. V. (ed. Hemantaranjan, A.), Scientific Publishers, Jodphur, India, pp. 181–226.Google Scholar
Flexas, J., Gulías, J., Jonasson, S., et al. (2001). Seasonal patterns and control of gas-exchange in local populations of the Mediterranean evergreen shrub Pistacia lentiscus L. Acta Oecologica, 22, 33–43.CrossRefGoogle Scholar
Flexas, J., Ortuño, M.F., Ribas-Carbó, M., et al. (2007c). Mesophyll conductance to CO2 in Arabidopsis thaliana. New Phytologist, 175, 501–511.CrossRefGoogle ScholarPubMed
Flexas, J., Ribas-Carbó, M., Bota, J., et al. (2006b). Decreased Rubisco activity during water stress is not induced by decreased relative water content but related to conditions of low stomatal conductance and chloroplast CO2 concentration. The New Phytologist, 172, 73–82.CrossRefGoogle Scholar
Flexas, J., Ribas-Carbó, M., Díaz-Espejo, A., et al. (2008). Mesophyll conductance to CO2: current knowledge and future prospects. Plant Cell and Environment, 31, 602–621.CrossRefGoogle ScholarPubMed
Flexas, J., Ribas-Carbó, M., Hanson, D.T., et al. (2006a). Tobacco aquaporin NtAQP1 is involved in mesophyll conductanve to CO2 in vivo. Plant Journal, 48, 427–439.CrossRefGoogle ScholarPubMed
Flinn, A.M., Atkins, C.A. and Pate, J.S. (1977). Significance of photosynthetic and respiratory exchanges in the carbon economy of the developing pea fruit. Plant Physiology, 60, 412–418.CrossRefGoogle ScholarPubMed
Flood, P.J., Harbinson, J. and Aarts, M.G.M. (2011). Natural genetic variation in plant photosynthesis. Trends in Plant Science, 16, 327–335.CrossRefGoogle ScholarPubMed
Florez-Sarasa, I.D., Flexas, J., Rasmusson, A.G. et al. (2011). In vivo cytochrome and alternative pathway respiration in leaves of Arabidopsis thaliana plants with altered alternative oxidase under different light conditions. Plant, Cell and Environment, 34, 1373–1383.CrossRefGoogle ScholarPubMed
Flor-Henry, M., McCabe, T.C., de Bruxelles, G.L., et al. (2004). Use of a highly sensitive two-dimensional luminescence imaging system to monitor endogenous bioluminescence in plant leaves. BMC Plant Biology, 4, 19.CrossRefGoogle ScholarPubMed
Flors, C., Fryer, M.J., Waring, J., et al. (2006). Imaging the production of singlet oxygen in vivo using a new fluorescent sensor, Singlet Oxygen Sensor Green (R). Journal of Experimental Botany, 57, 1725–1734.CrossRefGoogle Scholar
Flowers, M.D., Fiscus, E.L., Burkey, K.O., et al. (2007) Photosynthesis, chlorophyll fluorescence, and yield of snap bean (Phaseolus vulgaris L.) genotypes differing in sensitivity to ozone. Environmental and Experimental Botany, 61, 190–198.CrossRefGoogle Scholar
Flügge, U.I. (1999). Phosphate translocators in plastids. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 27–45.CrossRefGoogle ScholarPubMed
Flügge, U.I. and Heldt, H.W. (1991). Metabolite translocators of the chloroplast envelope. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 129–144.CrossRefGoogle Scholar
Foley, J.A., Prentice, I.C., Ramankutty, N., et al. (1996). An integrated biosphere model of land surface processes, terrestrial carbon balance, and vegetation dynamics. Global Biogeochemical Cycles, 10, 603–628.CrossRefGoogle Scholar
Forest, F., Grenyer, R., Rouget, M., et al. (2007). Preserving the evolutionary potential of floras in biodiversity hotspots. Nature, 445, 757–760.CrossRefGoogle ScholarPubMed
Fork, D.C. and Murata, N. (1990). The effect of light intensity on the assay of the low temperature limit of photosynthesis using msec delayed light emission. Photosynthesis Research, 23, 319–323.CrossRefGoogle ScholarPubMed
Fork, D.C., Mohanty, P. and Hoshina, S. (1985). The detection of early events in heat disruption of thylakoid membranes by delayed light emission. Physiologie Végétale, 23, 511–521.Google Scholar
Forseth, I.N. and Ehleringer, J.R. (1982). Ecophysiology of two solar tracking desert winter annuals. II. Leaf movements, water relations and microclimate. Oecologia, 54, 41–49.CrossRefGoogle ScholarPubMed
Forseth, I.N. and Ehleringer, J.R. (1983). Ecophysiology of two solar tracking desert winter annuals. III. Gas exchange responses to light, CO, and VPD in relation to long-term drought. Oecologia, 57, 344–351.CrossRefGoogle ScholarPubMed
Förstel, H. (1978). Contribution of oxygen isotope fractionation during the transpiration of plant leaves to the biogeochemical oxygen cycle. In: Environmental Biogeochemistry and Geomicrobiology, Vol 3 (ed. Krumbein, W.E.) Ann Arbor Science, Ann Arbor, MI, USA, pp. 811–824.Google Scholar
Förster, B., Mathesius, U. and Pogson, B.J. (2006). Comparative proteomics of high light stress in the model alga Chlamydomonas reinhardtii. Proteomics, 6, 4309–4320.CrossRefGoogle ScholarPubMed
Förster, B., Osmond, C.B. and Pogson, B.J. (2009). De novo synthesis and degradation of Lx and V cycle pigments during shade and sun acclimation in Avocado leaves. Plant Physiology, 149, 1179–1195.CrossRefGoogle Scholar
Forster, P., Ramaswamy, V., Artaxo, P., et al. (2007). Changes in atmospheric constituents and in eadiative forcing. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change(eds Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., M.Tignor, and Miller), H.L., Cambridge University Press, Cambridge, UK and New York, USA.Google Scholar
Fournier, A., Daumard, F., Champagne, S., et al. (2012). Effects of canopy structure on sun-induced chlorophyll fluorescence. Journal of Photogrammetry and Remote Sensing. IJPRS, 68: 112–120.CrossRefGoogle Scholar
Fowler, D., Cape, J.N., Coyle, M., et al. (1999a). Modelling photochemical oxidant formation, transport, deposition and exposure of terrestrial ecosystems. Environmental Pollution, 100, 43–55.CrossRefGoogle ScholarPubMed
Fowler, D., Cape, J.N., Coyle, M., et al. (1999b). The global exposure of forests to air pollutants. Water, Air, and Soil Pollution, 116, 5–32.CrossRefGoogle Scholar
Fowler, D., Flechard, C., Skiba, U.T.E., et al. (1998). The atmospheric budget of oxidized nitrogen and its role in ozone formation and deposition. New Phytologist, 139, 11–23.CrossRefGoogle Scholar
Fox, D.L. and Koch, P.L. (2003). Tertiary history of C4 biomass in the Great Plains, USA. Geology, 31, 809–812.CrossRefGoogle Scholar
Foy, C.D., Chaney, R.L. and White, M.C. (1978). The physiology of metal toxicity in plants. Annual Review of Plant Physiology, 29, 511–566.CrossRefGoogle Scholar
Foyer, C. and Spencer, C. (1986). The relationship between phosphate status and photosynthesis in leaves. Effects on intracellular orthophosphate distribution, photosynthesis and assimilate partitioning. Planta, 167, 369–375.CrossRefGoogle ScholarPubMed
Foyer, C., Furbank, R., Harbinson, J., et al. (1990). The mechanisms contributing to photosynthetic control of electron transport by carbon assimilation in leaves. Photosynthesis Research, 25, 83–100.CrossRefGoogle ScholarPubMed
Foyer, C., Lelandais, M. and Kunert, K.J. (1994). Photooxidative stress in plants. Physiologia Plantarum, 92, 696–717.CrossRefGoogle Scholar
Foyer, C.H., Noctor, G. and Verrier, P. (2006a). Photosynthetic carbon-nitrogen interactions: modelling inter-pathway control and signalling. In: Control of Primary Metabolism in Plants, Annual Plant Reviews, Volume 22 (eds McManus, M. and Plaxton, B.), Blackwell Publishing, Oxford, UK, pp. 325–347.Google Scholar
Foyer, C.H. and Harbinson, J. (1994). Oxygen metabolism and the regulation of photosynthetic electron transport. In: Causes of Photooxidative Stresses and Amelioration of Defense Systems in Plants (eds Foyer, C.H. and Mullineaux, P.), CRC Press, Florida, USA, pp.1–42.Google Scholar
Foyer, C.H. and Noctor, G. (2003). Redox sensing and signalling associated with reactive oxygen in chloroplasts, peroxisomes and mitochondria. Physiologia Plantarum, 119, 355–364.CrossRefGoogle Scholar
Foyer, C.H. and Noctor, G. (2005). Oxidant and antioxidant signalling in plants: a re-evaluation of the concept of oxidative stress in a physiological context. Plant, Cell and Environment, 28, 1056–1071.CrossRefGoogle Scholar
Foyer, C.H. and Noctor, G. (2009). Redox regulation in photosynthetic organisms: signaling, acclimation and practical implications. Antioxidants and Redox Signaling, 11, 861–905.CrossRefGoogle ScholarPubMed
Foyer, C.H., Bloom, A., Queval, G., et al. (2009). Photorespiratory metabolism: genes, mutants, energetics, and redox signaling. Annual Reviews of Plant Biology, 60, 455–484.CrossRefGoogle ScholarPubMed
Foyer, C.H., Lelandais, M. and Harbinson, J. (1992). Control of the quantum efficiencies of PSI and PSII, electron flow and enzyme activation following dark to light transitions in pea leaves; the relationship between NADP/NADPH ratios and NADP-malate dehydrogenase activation state. Plant Physiology, 99, 979–986.CrossRefGoogle Scholar
Foyer, C.H., Pellny, T.K., Locato, V., et al. (2008). Analysis of redox relationships in the plant cell cycle: determinations of ascorbate, glutathione and poly (ADPribose) polymerase (PARP) in plant cell cultures. Redox Mediated Signal Transduction. Methods in Molecular Biology Series, 476, 193–209.CrossRefGoogle Scholar
Foyer, C.H., Trebst, A. and Noctor, G. (2006b). Protective and signalling functions of ascorbate, glutathione and tocopherol in chloroplasts. In: Advances in Photosynthesis and Respiration, Volume 19, “Photoprotection, Photoinhibition, Gene Regulation, and Environment” (eds Demmig-Adams, B. and Adams, W.W.). Kluwer Academic Publishers. Dordrecht, Netherlands, pp. 241–268.Google Scholar
Fracheboud, Y., Luquez, V., Björkén, L., et al. (2009). The control of autumn senescence in European aspens (Populus tremula). Plant Physiology, 149, 1982–1991.CrossRefGoogle Scholar
Francey, R.J. (1985). Cape Grim isotope measurements-a preliminary assessment. Journal of Atmospheric Chemistry, 3, 247–260.CrossRefGoogle Scholar
Francey, R.J. and Tans, P.P. (1987). Latitudinal variation in oxygen-18 of atmospheric CO2. Nature, 327, 495–497.CrossRefGoogle Scholar
Franck, F., Juneau, P. and Popovic, R. (2002). Resolution of the photosystem I and photosystem II contributions to chlorophyll fluorescence of intact leaves at room temperature. Biochimica et Biophysica Acta, 1556, 239–246.CrossRefGoogle ScholarPubMed
Franco, A.C., de Soyza, A.G., Virginia, R.A., et al. (1994). Effects of plant size and water relations on gas exchange and growth of the desert shrub Larrea tridentata. Oecologia, 97, 171–178.CrossRefGoogle ScholarPubMed
Franke, J. and Menz, G. (2007). Multi-temporal wheat disease detection by multi-spectral remote sensing. Precision Agriculture, 8, 161–172.CrossRefGoogle Scholar
Frankenberg, C., Fisher, J.B., Worden, J., et al. (2011). New global observations of the terrestrial carbon cycle from GOSTA: Patterns of plant fluorescence with gross primary productivity. Geophys. Res. Lett. 38(17):L17706.
Franks, P.J. and Beerling, D.J. (2009). Maximum leaf conductance driven by CO2 effects on stomatal size and density over geologic time. Proceedings of the National Academy of Sciences USA, 106, 10343–10347.CrossRefGoogle ScholarPubMed
Franks, P.J. and Farquhar, G.D. (2001). The effect of exogenous abscisic acid on stomatal development, stomatal mechanics, and leaf gas exchange in Tradescantia virginiana. Plant Physiology, 125, 935–942.CrossRefGoogle ScholarPubMed
Franks, P.J. and Farquhar, G.D. (2007). The mechanical diversity of stomata and its significance in gas-exchange control. Plant Physiology, 143, 78–87.CrossRefGoogle ScholarPubMed
Frantz, J.M., Joly, R.J. and Mitchell, C.A. (2000). Intracanopy lighting influences radiation capture, productivity, and leaf senescence in cowpea canopies. Journal of the American Society of Horticultural Science, 125, 694–701.Google Scholar
Fravolini, A., Williams, D.G. and Thompson, T.L. (2002). Carbon isotope discrimination and bundle sheath leakiness in three C4 subtypes grown under variable nitrogen, water and atmospheric CO2 supply. Journal of Experimental Botany, 53, 2261–2269.CrossRefGoogle Scholar
Fredeen, A.L., Koch, G.W. and Field, C.B. (1995). Effects of atmospheric CO2 enrichment on ecosystem CO2 exchange in a nutrient and water limited grassland. Journal of Biogeography, 22, 215–219.CrossRefGoogle Scholar
Fredeen, A.L., Rao, I.M. and Terry, N. (1989). Influence of phosphorus nutrition on growth and carbon partition in Glycine max. Plant Physiology, 89, 225–230.CrossRefGoogle Scholar
Freitag, H. and Stichler, W. (2002). Bienertia cycloptera Bunge ex Boiss., Chenopodiaceae, another C4 plant without Kranz tissues. Plant Biology, 4, 121–131.CrossRefGoogle Scholar
Fridlyand, L., Backhausen, J. and Scheibe, R. (1998). Flux control of the malate valve in leaf cells. Archives of Biochemistry and Biophysics, 349, 290–298.CrossRefGoogle ScholarPubMed
Friedli, H., Siegenthaler, U., Rauber, D., et al. (1987). Measurements of concentration, 13C/12C and 18O/16O ratios of tropospheric carbon dioxide over Switzerland. Tellus, 39B, 80–88.CrossRefGoogle Scholar
Friedman, I. (1953). Deuterium content of natural water and other substances. Geochimica et Cosmochimica Acta, 4, 89–103.CrossRefGoogle Scholar
Friedman, W.E. (2008). Hydatellaceae are water lilies with gymnospermous tendencies. Nature, 453, 94–97.CrossRefGoogle ScholarPubMed
Friend, A.D. and Woodward, F.I. (1990). Evolutionary and ecophysiological responses of mountain plants to the growing season environment. Advances in Ecological Research, 20, 59–124.CrossRefGoogle Scholar
Friend, A.D. (2001). Modelling canopy CO2 fluxes: are ‘big-leaf’ simplifications justified?Global Ecology and Biogeography, 10, 603–619.CrossRefGoogle Scholar
Friend, A.D., Arneth, A., Kiang, N.Y., et al. (2007). FLUXNET and modelling the global carbon cycle. Global Change Biology, 13, 610–633.CrossRefGoogle Scholar
Friis, E.M., Pedersen, K.R. and Crane, P.R. (2005). When Earth started blooming: insights from the fossil record. Current Opinion in Plant Biology, 8, 5–12.CrossRefGoogle ScholarPubMed
Fromm, J. and Fei, H. (1998). Electrical signalling and gas exchange in maize plants of drying soil. Plant Science, 132, 203–213.CrossRefGoogle Scholar
Fromm, J. and Lautner, S. (2007). Electrical signals and their physiological significance in plants. Plant, Cell and Environment, 30, 249–257.CrossRefGoogle ScholarPubMed
Frost, D.L., Gurney, A.L., Press, M.C., et al. (1997). Striga hermonthica reduces photosynthesis in sorghum: the importance of stomatal limitations and a potential role for ABA. Plant, Cell and Environment, 20, 483–492.CrossRefGoogle Scholar
Frost-Christensen, H. and Floto, F. (2007). Resistance to CO2 diffusion in cuticular membranes of amphibious plants and the implication for CO2 acquisition. Plant, Cell and Environment, 30, 12–18.CrossRefGoogle ScholarPubMed
Fryer, M.J., Andrews, J.R., Oxborough, K., et al. (1998). Relationship between CO2 assimilation, photosynthetic electron transport, and active O2 metabolism in leaves of maize in the field during periods of low temperature. Plant Physiology, 116, 571–580.CrossRefGoogle ScholarPubMed
Fryer, M.J., Ball, L., Oxborough, K., et al. (2003). Control of Ascorbate Peroxidase 2 expression by hydrogen peroxide and leaf water status during excess light stress reveals a functional organisation of Arabidopsis leaves. The Plant Journal, 33, 691–705.CrossRefGoogle ScholarPubMed
Fryer, M.J., Oxborough, K., Mullineaux, P.M., et al. (2002). Imaging of photo-oxidative stress responses in leaves. Journal of Experimental Botany, 53, 1249–1254.Google ScholarPubMed
Fu, J., Momcilovic, I., Clemente, T.E., et al. (2008). Heterologous expression of a plastid EF-Tu reduces protein thermal aggregation and enhances CO2 fixation in wheat (Triticum aestivum) following heat stress. Plant Molecular Biology, 68, 277–288.CrossRefGoogle ScholarPubMed
Fuchs, E.E. and Livingston, N.J. (1996). Hydraulic control of stomatal conductance in Douglas fir [Pseudotsuga menziesii (Mirb.) Franco] and alder [Alnus rubra Bong.] seedlings. Plant, Cell and Environment, 19, 1091–1098.CrossRefGoogle Scholar
Fuchs, E.E., Livingston, N.J. and Rose, P.A. (1999). Structure-activity relationships of ABA analogs based on their effects on the gas exchange of clonal white spruce (Picea glauca) emblings. Physiologia Plantarum, 105, 246–256.CrossRefGoogle Scholar
Fuentes, D.A., Gamon, J.A., Cheng, Y., et al. (2006). Mapping carbon and water vapor fluxes in a chaparral ecosystem using vegetation indices derived from AVIRIS. Remote Sensing of Environment, 103, 312–323.CrossRefGoogle Scholar
Fuhrer, J. (2009). Ozone risk for crops and pastures in present and future climates. Naturwissenschaften, 96, 173–194.CrossRefGoogle ScholarPubMed
Fuhrmann, J., Johnen, T. and Heise, K.P. (1994). Compartmentation of fatty acid metabolism in zygotic rape embryo. Journal of Plant Physiology, 143, 565–569.CrossRefGoogle Scholar
Fujii, H., Chinnusamy, V., Rodrigues, A., et al. (2009). In vitro reconstitution of an abscisic acid signalling pathway. Nature, 462, 660–664.CrossRefGoogle ScholarPubMed
Fujita, K., Takagi, S. and Terashima, I. (2008). Leaf angle in Chenopodium album is determined by two processes: induction and cessation of petiole curvature. Plant, Cell and Environment, 31, 1138–1146.CrossRefGoogle ScholarPubMed
Fukshansky, L. (1981). Optical properties of plants. In: Plants and the Daylight Spectrum (ed Krumbein, W.E.), Academic Press, London, UK, pp. 21–40.Google Scholar
Fukuda, H. (2000). Programmed cell death of tracheary elements as a paradigm in plants. Plant Molecular Biology, 44, 245–253.CrossRefGoogle ScholarPubMed
Funayama, S., Hikosaka, K. and Yahara, T. (1997b). Effects of virus infection and growth irradiance on fitness components and photosynthetic properties of Eupatorium makinoi (Compositae). American Journal of Botany, 84, 823–829.CrossRefGoogle Scholar
Funayama, S., Sonoike, K. and Terashima, I. (1997a). Photosynthetic properties of leaves of Eupatorium makinoi infected by ageminivirus. Photosynthesis Research, 53, 253–261.CrossRefGoogle Scholar
Fung, I., Field, C.B., Berry, J.A., et al. (1997). Carbon-13 exchanges between the atmosphere and biosphere,Global Biogeochemical Cycles, 11, 507–533.CrossRefGoogle Scholar
Funk, J.L., Giardina, C.P., Knohl, A., et al. (2006). Influence of nutrient availability, stand age, and canopy structure on isoprene flux in a Eucalyptus saligna experimental forest. Journal of Geophysical Research – Biogeosciences 111:G02012.CrossRefGoogle Scholar
Furbank, R.T., Jenkins, C.L.D. and Hatch, M.D. (1989). CO2 concentrating mechanism of C4 photosynthesis. Permeability of isolated bundle sheath cells to inorganic carbon. Plant Physiology, 91, 1364–1371.CrossRefGoogle Scholar
Furley, P. (2006). Tropical savannas. Progress in Physical Geography, 30, 105–121.CrossRefGoogle Scholar
Furley, P.A. and Metcalfe, S.E. (2007). Dynamic changes in savanna and seasonally dry vegetation through time. Progress in Physical Geography, 31, 633–642.CrossRefGoogle Scholar
Gaastra, P. (1959). Photosynthesis of crop plants as influenced by light, carbon dioxide, temperature, and stomatal diffusion resistance. Mededelingen van Landbouwhogeschool te Wageningen, Nederland, 59, 1–68.Google Scholar
Gadjev, I., Vanderauwera, S., Gechev, T.S., et al. (2006). Transcriptomic footprints disclose specificity of reactive oxygen species signalling in Arabidopsis. Plant Physiology, 141, 436–445.CrossRefGoogle Scholar
Gale, J. (1972). Availability of carbon dioxide for photosynthesis at high altitudes. Ecology, 53, 494–497.CrossRefGoogle Scholar
Gallé, A. and Feller, U. (2007). Changes of photosynthetic traits in beech saplings (Fagus sylvatica) under severe drought stress and during recovery. Physiologia Plantarum, 131, 412–421.CrossRefGoogle ScholarPubMed
Gallé, A., Florez-Sarasa, I., Thameur, A, et al. (2010). Effects of drought stress and subsequent rewatering on photosynthetic and respiratory pathways in Nicotiana sylvestris wild type and the mitochondrial complex I-deficient CMSII mutant. Journal of Experimental Botany, 61, 765–775.CrossRefGoogle ScholarPubMed
Gallé, A., Florez-Sarasa, I., Tomás, M., et al. (2009). The role of mesophyll conductance during water stress and recovery in tobacco (Nicotiana sylvestris): acclimation or limitation?Journal of Experimental Botany, 60, 2379–2390.CrossRefGoogle ScholarPubMed
Gallé, A., Haldimann, P. and Feller, U. (2007). Photosynthetic performance and water relations in young pubescent oak (Quercus pubescens) trees during drought stress and recovery. New Phytologist, 174, 799–810.CrossRefGoogle ScholarPubMed
Gallo, K., Daughtry, C.S.T. and Wiegand, C.L. (1993). Errors in measuring absorbed radiation and computing crop radiation use efficiency. Agronomy Journal, 85, 1222–1228.CrossRefGoogle Scholar
Gallo, K.P. and Daughtry, C.S.T. (1986). Techniques for measuring intercepted and absorbed photosynthetically active radiation in corn canopies. Agronomy Journal, 78, 752–756.CrossRefGoogle Scholar
Galmés, J., Medrano, H. and Flexas, J. (2007a). Photosynthetic limitations in response to water stress and recovery in Mediterranean plants with different growth forms. New Phytologist, 175, 81–93.CrossRefGoogle ScholarPubMed
Galmés, J., Abadía, A., Cifre, J., et al. (2007b). Photoprotection processes under water stress and recovery in Mediterranean plants with different growth forms and leaf habits. Physiologia Plantarum, 130, 495–510.CrossRefGoogle Scholar
Galmés, J., Conesa, M.A., Ochogavía, J.M. et al. (2011). Physiological and morphological adaptations in relation to water use efficiency in Mediterranean accessions of Solanum lycopersicum. Plant, Cell and Environment, 34, 245–260.CrossRefGoogle ScholarPubMed
Galmés, J., Flexas, J., Keys, A.J., et al. (2005). Rubisco specificity factor tends to be larger in plant species from drier habitats and in species with persistent leaves. Plant, Cell and Environment, 28, 571–579.CrossRefGoogle Scholar
Galmés, J., Flexas, J., Ribas-Carbó, M., et al. (2010). Rubisco activity in Mediterranean species is regulated by chloroplastic CO2 concentration under water stress. Journal of Experimental Botany, 62, 653–665.CrossRefGoogle ScholarPubMed
Galmés, J., Medrano, H. and Flexas, J. (2006). Acclimation of Rubisco specificity factor to drought in tobacco: discrepancies between in vitro and in vivo estimations. Journal of Experimental Botany, 57, 3659–3667.CrossRefGoogle ScholarPubMed
Galmés, J., Medrano, H. and Flexas, J. (2007c) Water relations and stomatal characteristics of Mediterranean plants with different growth forms and leaf habits: responses to water stress and recovery. Plant and Soil, 290, 139–155.CrossRefGoogle Scholar
Galmés, J., Medrano, H. and Flexas, J. (2007f). Photosynthesis and photoinhibition in response to drought in a pubescent (var. minor) and a glabrous (var. palaui) variety of Digitalis minor. Environmental and Experimental Botany, 60, 105–111.CrossRefGoogle Scholar
Galmés, J., Pou, A., Alsina, M.M., et al. (2007d). Aquaporin expression in response to different water stress intensities and recovery in Richter-110 (Vitis sp.): relationship with ecophysiological status. Planta, 226, 671–681.CrossRefGoogle ScholarPubMed
Galmés, J., Ribas-Carbó, M., Medrano, H. and Flexas, J. (2007e). Response of leaf respiration to water stress in Mediterranean species with different growth forms. Journal of Arid Environments, 68, 206–222.CrossRefGoogle Scholar
Galvez, D. and Pearcy, R.W. (2003). Petiole twisting in the crowns of Psychotria limonensis: implications for light interception and daily carbon gain. Oecologia, 135, 22–29.CrossRefGoogle ScholarPubMed
Gamon, J.A. and Qiu, H.L. (1999). Ecological applications of remote sensing at multiple scales. In: Handbook of Functional Plant Ecology (eds Pugnaire, F.I. and Valladares, F.), Marcel Dekker, New York, USA, pp. 805–846.Google Scholar
Gamon, J.A. and Surfus, J.S. (1999). Assessing leaf pigment content and activity whith a reflectometer. New Phytologist, 143, 105–117.CrossRefGoogle Scholar
Gamon, J.A., Field, C.B., Bilger, W., et al. (1990). Remote sensing of the xanthophyll cycle and chlorophyll fluorescence in sunflower leaves and canopies. Oecologia, 85, 1–7.CrossRefGoogle ScholarPubMed
Gamon, J.A., Field, C.B., Goulden, M.L., et al. (1995). Relationships between NDVI, canopy structure, and photosynthesis in three Californian vegetation types. Ecological Applications, 5, 28–41.CrossRefGoogle Scholar
Gamon, J.A., Huemmrich, K.F., Peddle, D.R., et al. (2004). Remote sensing in BOREAS: lessons learned. Remote Sensing of Environment, 89, 139–162.CrossRefGoogle Scholar
Gamon, J.A., Peñuelas, J. and Field, C.B. (1992). A narrow-waveband spectral index that traces diurnal changes in photosynthetic efficiency. Remote Sensing of the Environment, 41, 35–44.CrossRefGoogle Scholar
Gamon, J.A., Serrano, L. and Surfus, J.S. (1997). The photochemical reflectance index: an optical indicator of photosynthetic radiation use efficiency across species, functional types, and nutrient levels. Oecologia, 112, 492–501.CrossRefGoogle ScholarPubMed
Gan, K.S., Wong, S.C., Yong, J.W.H., et al. (2002). 18O spatial patterns of vein xylem water, leaf water, and dry matter in cotton leaves. Plant Physiology, 130, 1008–1021.CrossRefGoogle Scholar
Gao, D., Gao, Q., Xu, H.Y., et al. (2009). Physiological responses to gradual drought stress in the diploid hybrid Pinus densata and its two parental species. Trees, 23, 717–728.CrossRefGoogle Scholar
Gao, Q., Zhao, P., Zeng, X., et al. (2002). A model of stomatal conductance to quantify the relationship between leaf transpiration, microclimate and soil water stress. Plant, Cell and Environment, 25, 1373–1381.CrossRefGoogle Scholar
Gao, Z., Bian, L. and Zhou, X. (2003). Measurements of turbulent transfer in the near-surface layer over a rice paddy in China. Journal of Geophysical Research, 108, 4387.CrossRefGoogle Scholar
Garbulsky, M.F., Peñuelas, J., Gamon, J. et al. (2011). The photochemical reflectance index (PRI) and the remote sensing of leaf, canopy and ecosystem radiation use efficiencies: a review and meta-analysis. Remote Sensing of the Environment, 115, 281–297.CrossRefGoogle Scholar
Garcia, R.L., Idso, S.B., Wall, G.W., et al. (1994). Changes in net photosynthesis and growth of Pinus eldarica seedlings in response to atmospheric CO2 enrichment. Plant Cell Environment, 17, 971–978.CrossRefGoogle Scholar
García-Plazaola, J.I. and Becerril, J.M. (2001). Seasonal changes in photosynthetic pigments and antioxidants in beech (Fagus sylvatica) in a Mediterranean climate: implications for tree decline diagnosis. Australian Journal of Plant Physiology, 28, 225–232.Google Scholar
García-Plazaola, J.I., Artetxe, U., Duñabeitia, M.K., et al. (1999). Role of photoprotective systems of holm-oak (Quercus ilex) in the adaptation to winter conditions. Journal of Plant Physiology, 155, 625–630.CrossRefGoogle Scholar
García-Plazaola, J.I., Becerril, J.M., Hernández, A., et al. (2004). Acclimation of antioxidant pools to the light environment in a natural forest canopy. The New Phytologist, 163, 87–97.CrossRefGoogle Scholar
García-Plazaola, J.I., Esteban, R., Hormaetxe, K., et al. (2008a). Seasonal reversibility of acclimation to irradiance in leaves of common box (Buxus sempervirens L.) in a deciduous forest. Flora, 203, 254–260.CrossRefGoogle Scholar
García-Plazaola, J.I., Esteban, R., Hormaetxe, K., et al. (2008b). Photoprotective responses of Mediterranean and Atlantic trees to the extreme heat-wave of summer 2003 in Southwestern Europe. Trees: Structure and Function, 3, 385–392.CrossRefGoogle Scholar
Garcia-Plazaola, J.I., Hernández, A. and Becerill, J.M. (2003a). Antioxidant and pigment composition during autumnal leaf senescence in woody deciduous species differing in their ecological traits. Plant Biology, 5, 557–566.CrossRefGoogle Scholar
García-Plazaola, J.I., Matsubara, S. and Osmond, C.B. (2007). The lutein epoxide cycle in higher plants: its relationships to other xanthophyll cycles and possible functions. Functional Plant Biology, 34, 759–773.CrossRefGoogle Scholar
García-Plazaola, J.I., Olano, J.M., Hernandez, J.M., et al. (2003b). Photoprotection in evergreen Mediterranean plants during sudden periods of intense cold weather. Trees, 17, 285–291.Google Scholar
Gardeström, P. and Wigge, B. (1988). Influence of photorespiration on ATP/ADP ratios in the chloroplasts, mitochondria and cytosol, studies by rapid fractionation of barley protoplasts. Plant Physiology, 88, 69–76.CrossRefGoogle Scholar
Gardeström, P., Igamberdiev, A.U. and Raghavendra, A.S. (2002). Mitochondrial functions in the light and significance to carbon-nitrogen interactions. In: Photosynthetic Nitrogen Assimilation and Associated Carbon and Respiratory Metabolism (eds Fouer, C.H. and Noctor, G.), Kluwer Academic Publisher, Dordrecht, Netherlands, pp. 151–172.Google Scholar
Garmier, M., Priault, P., Vidal, G., et al. (2008). Light and oxygen are not required for harpin-induced cell death. Journal of Biological Chemistry, 282, 37556–37566.CrossRefGoogle Scholar
Gates, D.M. (1962) Energy Exchange in the Biosphere. Harper and Row, New York, USA, pp. 151.Google Scholar
Gates, D.M. (1970). Physical and physiological properties of plants. In: Remote Sensing. Nat Acad of Sci., Washington DC, USA, pp. 224–252.Google Scholar
Gates, D.M. (1980). Biophysical Ecology. Springer-Verlag, New York – Heidelberg – Berlin.CrossRefGoogle Scholar
Gates, D.M. and Papian, L.E. (1971). Atlas of Energy Budgets of Plant Leaves. London: Academic Press.Google Scholar
Gauslaa, Y. (1984). Heat resistance and energy budget in different scandinavian plants. Holarctic Ecology, 7, 1–78.Google Scholar
Gaut, B.S. and Doebley, J.F. (1997). DNA sequence evidence for the segmental allotetraploid origin of maize. Proceedings of the National Academy of Sciences (USA), 94, 68090–68094.CrossRefGoogle ScholarPubMed
Gautier, H., Vavasseur, Al., Gans, P., et al. (1991). Relationship between respiration and photosynthesis in guard cell and mesophyll cell protoplasts of Commelina communis L. Plant Physiology, 95, 636–641.CrossRefGoogle ScholarPubMed
Geber, M.A. and Dawson, T.E. (1997). Genetic variation in stomatal and biochemical limitations to photosynthesis in the annual plant, Polygonum arenastrum. Oecologia, 109, 535–546.CrossRefGoogle ScholarPubMed
Gedney, N., Cox, P., Betts, R., et al. (2006) Detection of a direct carbon dioxide effect in continental river runoff records. Nature, 439, 835–838.CrossRefGoogle ScholarPubMed
Geiken, B., Masojídek, J., Rizzuto, M., et al. (1998). Incorporation of [35S]methionine in higher plants reveals that stimulation of the D1 reaction centre II protein turnover accompanies tolerance to heavy metal stress. Plant, Cell and Environment, 21, 1265–1273.CrossRefGoogle Scholar
Geissler, N., Hussin, S. and Koyro, H.W. (2009). Elevated atmospheric CO2 concentration ameliorates effects of NaCl salinity on photosynthesis and leaf structure of Aster tripolium L. Journal of Experimental Botany, 60, 137–151.CrossRefGoogle ScholarPubMed
Gent, M.P., Ferrandino, F.J. and Elmer, W.H. (1995). Effects of Verticillium wilt on gas exchange of entire eggplants. Canadian Journal of Botany, 73, 557–565.CrossRefGoogle Scholar
Gent, M.P.N., LaMondia, J.A., Ferrandino, F.J., et al. (1999). The influence of compost amendement or straw mulch on the reduction of gas exchange in potato by Verticillium dahliae and Pratylenchus penetrans. Plant Disease, 83, 371–376.CrossRefGoogle Scholar
Gentry, A.H. and Dodson, C.H. (1987). Diversity and evolution of Neotropical Vascular epiphytes. Annals of the Missouri Botanical Garden, 74, 205–233.CrossRefGoogle Scholar
Genty, B. and Harbinson, J. (1996). Regulation of light utilization for photosynthetic electron transport. In: Photosynthesis and the Environment (ed. Baker, N.R.), Kluwer, Amsterdam, Netherlands, pp. 69–99.Google Scholar
Genty, B. and Meyer, S. (1995). Quantitative mapping of leaf photosynthesis using chlorophyll fluorescence imaging. Australian Journal of Plant Physiology, 22, 277–284.CrossRefGoogle Scholar
Genty, B., Briantais, J.M. and Baker, N.R. (1989). The relationship between the quantum yield of photosynthetic electron transport and quenching of chlorophyll fluorescence. Biochimica et Biophysica Acta, 990, 87–92.CrossRefGoogle Scholar
Genty, B., Meyer, S., Pile, C., et al. (1998). CO2 diffusion inside leaf mesophyll of ligneous plants. In Photosynthesis: Mechanisms and Effects (ed. Garab, G.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 3961–3967.Google Scholar
Genty, B., Wonders, J. and Baker, N. (1990). Nonphotochemical quenching of F0 in leaves is emission wavelength dependent. Consequences for quenching analysis and its interpretation. Photosynthesis Research, 26, 133–139.CrossRefGoogle Scholar
Gepstein, S. (2004). Leaf senescence- not just a “wear and tear” phenomenon. Genome Biology, 5, 212.CrossRefGoogle Scholar
Gerbaud, A. and Andre, M. (1979). Photosynthesis and photorespiration in whole plants of wheat. Plant Physiology, 64, 735–738.CrossRefGoogle ScholarPubMed
Gerbens-Leenes, W., Hoekstra, A.Y. and van der Meer, T.H. (2009). The water footprint of bioenergy. Proceedings of the National Academy of Science USA, 106, 10219–10223.CrossRefGoogle ScholarPubMed
Germeraad, J.H., Hopping, C.A. and Muller, J. (1967). Palynology of tertiary sediments from tropical areas. Review of Palaeobotany and Palynology, 6, 189–348.CrossRefGoogle Scholar
Germino, M.J. and Smith, W.K. (2000). High resistance to low-temperature photoinhibition in two alpine, snowbank species. Physiologia Plantarum, 110, 89–95.CrossRefGoogle Scholar
Germino, M.J. and Smith, W.K. (2001). Relative importance of microhabitat, plant form and photosynthetic physiology to carbon gain in two alpine herbs. Functional Ecology, 15, 243–251.CrossRefGoogle Scholar
Gerrish, G. (1989). Comparing crown growth and phenology of juvenile, early mature and late mature Metrosideros polymorpha trees. Pacific Science, 43, 211–222.Google Scholar
Gerrish, G. (1990). Relating carbon allocation patterns to tree senescence in Metrosideros forests. Ecology, 71, 1176–1184.CrossRefGoogle Scholar
Gerst, U., Schönknecht, G. and Heber, U. (1994). ATP and NADPH as the driving force of carbon reduction in leaves in relation to thylakoid energization by light. Planta, 193, 421–429.CrossRefGoogle Scholar
Gerwing, J.J. and Farias, D.L. (2000). Integrating liana abundance and forest stature into an estimate of total aboveground biomass for an eastern Amazonian forest. Journal of Tropical Ecology, 16, 327–335.CrossRefGoogle Scholar
Gessler, A., Brandes, E., Buchmann, N., et al. (2009b). Tracing carbon and oxygen isotope signals from newly assimilated sugars in the leaves to the tree-ring archive. Plant, Cell and Environment, 32, 780–795.CrossRefGoogle ScholarPubMed
Gessler, A., Schrempp, S., Matzarakis, A., et al. (2001). Radiation modifies the effect of water availability on the carbon isotope composition of beech (Fagus sylvatica). New Phytologist, 150, 653–664.CrossRefGoogle Scholar
Gessler, A., Tcherkez, G., Karyanto, O., et al. (2009a). On the metabolic origin of the carbon isotope composition of CO2 evolved from darkened light-acclimated leaves in Ricinus communis. New Phytologist, 181, 374–386.CrossRefGoogle ScholarPubMed
Gessler, A., Tcherkez, G., Peuke, A.D., et al. (2008). Experimental evidence for diel variations of the carbon isotope composition in leaf, stem and phloem sap organic matter in Ricinus communis. Plant, Cell and Environment, 31, 941–953.CrossRefGoogle ScholarPubMed
Ghannoum, O. (2009). C4 photosynthesis and water stress. Annals of Botany, 103, 635–644.CrossRefGoogle ScholarPubMed
Ghannoum, O., Evans, J.R., Chow, W.S., et al. (2005). Faster rubisco is the key to superior nitrogen-use efficiency in NADP-malic enzyme relative to NAD-malic enzyme C-4 grasses. Plant Physiology, 137, 638–650.CrossRefGoogle Scholar
Ghannoum, O., Phillips, N.G., Conroy, J.P., et al. (2010a). Exposure to preindustrial, current and future atmospheric CO2 and temperature differentially affects growth and photosynthesis in Eucalyptus. Global Change Biology, 16, 303–319.CrossRefGoogle Scholar
Ghannoum, O., Phillips, N.G., Sears, M.A., et al. (2010b). Photosynthetic responses of two eucalypts to industrial-age changes in atmospheric (CO2) and temperature. Plant, Cell and Environment, 33, 1671–1681.CrossRefGoogle ScholarPubMed
Ghannoum, O., von Caemmerer, S. and Conroy, J.P. (2002). The effect of drought on plant water use efficiency on nine NAD-ME and nine NADP-ME Australian C4 grasses. Functional Plant Biology, 29, 1337–1348.CrossRefGoogle Scholar
Ghashghaie, J. and Cornic, C. (1994). Effect of temperature on partitioning of photosynthetic electron flow between CO2 assimilation and O2 reduction and on the CO2/O2 specificity of Rubisco. Journal of Plant Physiology, 143, 643–650.CrossRefGoogle Scholar
Ghashghaie, J., Badeck, F.W., Lanigan, G., et al. (2003). Carbon isotope fractionation during dark respiration and photorespiration in C3 plants. Phytochemistry Reviews, 2, 145–161.CrossRefGoogle Scholar
Ghashghaie, J., Duranceau, M., Badeck, F.W., et al. (2001). δ13C of CO2 respired in the dark in relation to leaf metabolites: comparisons between Nicotiana sylvestris and Helinathus annuus under drought. Plant, Cell and Environment, 24, 505–515.CrossRefGoogle Scholar
Gholz, H.L. (1982). Environmental limits on aboveground net primary production, leaf area, and biomass in vegetation zones of the Pacific Northwest. Ecology, 63, 469–481.CrossRefGoogle Scholar
Gibberd, M.R., Walker, R.R., Blackmore, D., et al. (2001). Transpiration efficiency and carbon-isotope discrimination of grapevines grown under well-watered conditions in either glasshouse or vineyard. Australian Journal of Grape and Wine Research, 7, 110–117.CrossRefGoogle Scholar
Gibert, F., Flamant, P.H., Bruneau, D., et al. (2006). 2-µm heterodyne differential absorption lidar measurements of at mospheric CO2 mixing ratio in the boundary layer. Applied Optics, 45, 4448–4458.CrossRefGoogle Scholar
Gibert, F., Schmidt, M., Cuesta, J., et al. (2007). Retrieval of average CO2 fluxes by combining in-situ CO2 measurements and backscatter lidar information. Journal of Geophysical Research, 112, D10301.CrossRefGoogle Scholar
Gibson, A.C. (1981). Vegetative anatomy of Pachycormus (Anacardiaceae). Linnean Society, Botany, 83, 273–284.CrossRefGoogle Scholar
Gibson, A.C. (1983). Anatomy of photosynthetic old stems of nonsucculent dicotyledons from North American deserts. Botanical Gazette, 144, 347–362.CrossRefGoogle Scholar
Gibson, A.C. (1996). Structure-Function Relations of Warm Desert Plants. Springer-Verlag, Berlin, Germany.CrossRefGoogle Scholar
Gibson, A.C. (1998). Photosynthetic organs of desert plants. BioScience, 48, 911–920.CrossRefGoogle Scholar
Gibson, A.C. and Nobel, P.S. (1986). The Cactus Primer. Harvard University Press, Cambridge, MA, USA.CrossRefGoogle Scholar
Gibson, A.C., Rundel, P.W. and Sharifi, M.R. (2008). Ecology and ecophysiology of a subalpine fellfield community on Mount Pinos, southern California. Madroño, 55, 41–51.CrossRefGoogle Scholar
Gielen, B. and Ceulemans, R. (2001). The likely impact of rising atmospheric CO2 on natural land and managed Populus: a literature review. Environmental Pollution, 115, 335–358.CrossRefGoogle Scholar
Gifford, R.M. (1995). Whole plant respiration and photosynthesis of wheat under increased CO2 concentration and temperature: long-term vs. short-term distinctions for modelling. Global Change Biology, 1, 385–396.CrossRefGoogle Scholar
Gifford, R.M. and Evans, L.T. (1981). Photosynthesis, carbon partitioning, and yield. Annual Review of Plant Physiology, 32, 485–509.CrossRefGoogle Scholar
Gil, P.M., Gurovich, L., Schaffer, B., et al. (2008). Root to leaf electrical signalling in avocado in response to light and soil water content. Journal of Plant Physiology, 165, 1070–1078.CrossRefGoogle Scholar
Gilbert, M., Wagner, H., Weingart, I., et al. (2004). A new type of thermoluminometer: a highly sensitive tool in applied photosynthesis research and and stress physiology. Journal of Plant Physiology, 161, 641–651.CrossRefGoogle ScholarPubMed
Gilbert, M.E., Zwieniecki, M.A. and Holbrook, N.M. (2011). Independent variation in photosynthetic capacity and stomatal conductance leads to differences in intrinsic water use efficiency in 11 soybean genotypes before and during mild drought. Journal of Experimental Botany, 62, 2875–2887.CrossRefGoogle ScholarPubMed
Gillon, J. and Yakir, D. (2001). Influence of carbonic anhydrase activity in terrestrial vegetation on the 18O content of atmospheric CO2. Science, 291, 2584–2587.CrossRefGoogle ScholarPubMed
Gillon, J.S. and Griffiths, H. (1997). The influence of (photo)respiration on carbon isotope discrimination in plants. Plant, Cell and Environment, 20, 1217–1230.CrossRefGoogle Scholar
Gillon, J.S. and Yakir, D. (2000a). Internal conductance to CO2 diffusion and C18OO discrimination in C3 leaves. Plant Physiology, 123, 201–213.CrossRefGoogle Scholar
Gillon, J.S. and Yakir, D. (2000b). Naturally low carbonic anhydrase activity in C4 and C3 plants limits discrimination against (COO)-18O during photosynthesis. Plant Cell and Environment, 23, 903–915.CrossRefGoogle Scholar
Gilmanov, T.G., Soussana, J.F., Aires, L., et al. (2007). Partitioning European grassland net ecosystem CO2 exchange into gross primary productivity and ecosystem respiration using light response function analysis. Agriculture, Ecosystems and Environment, 121, 93–120.CrossRefGoogle Scholar
Gilmore, A.M. and Ball, M.C. (2000). Protection and storage of chlorophyll in overwintering evergreens. Proc Natl Acad Sci USA, 97, 11098–11101.CrossRefGoogle ScholarPubMed
Gilmore, D.W., Seymour, R.S., Halteman, W.A., et al. (1995). Canopy dynamics and the morphological development of Abies balsamea: effects of foliage age on specific leaf area and secondary vascular development. Tree Physiology, 15, 47–55.CrossRefGoogle ScholarPubMed
Giordano, M., Beardall, J. and Raven, J.A. (2005). CO2 concentrating mechanisms in algae: mechanisms, environmental modulation, and evolution. Annual Review of Plant Biology, 56, 99–131.CrossRefGoogle ScholarPubMed
Giraud, E., Ho, L.H.M., Clifton, R., et al. (2008). The absence of alternative oxidase 1a in Arabidopsis results in acute sensitivity to combined light and drought stress. Plant Physiology, 147, 595–610.CrossRefGoogle Scholar
Giri, A.P., Wunsche, H., Mitra, S., et al. (2006). Molecular interactions between the specialist herbivore Manduca sexta (Lepidoptera, Sphingidae) and its natural host Nicotiana attenuata. VII. Changes in the plant’s proteome. Plant Physiology, 142, 1621–1641.CrossRefGoogle ScholarPubMed
Gitelson, A., Viña, A., Verma, S., et al. (2006). Relationship between gross primary production and chlorophyll content in crops: implications for the synoptic monitoring of vegetation productivity. Journal of Geophysical Research, 111, D08S11.CrossRefGoogle Scholar
Gitelson, A.A., Buschmann, C. and Lichtenthaler, H.K. (1998). Leaf chlorophyll fluorescence corrected for re-absorption by means of absorption and reflectance measurements. Journal of Plant Physiology, 152, 283–296.CrossRefGoogle Scholar
Givnish, T. (1979). On the adaptive significance of leaf form. In: Topics in Plant Population Biology (eds Solbrig, O.T., Jain, S., Johnson, G.B. and Raven, P.H.), Columbia University Press, New York, USA, pp. 375–407.CrossRef
Givnish, T.J. (1984). Leaf and canopy adaptations in tropical forests. In: Physiological Ecology of Plants of the Wet Tropics. Proceedings of an International Symposium held in Oxatepec and Los Tuxtlas, Mexico, June 29 to July 6, 1983 (eds Medina, E., Mooney, H.A., Vásquez-Yánes, C.), Dr. W. Junk Publishers, The Hague, Netherlands, pp. 51–84.Google Scholar
Givnish, T.J. (1988). Adaptation to sun and shade: a whole-plant perspective. Australian Journal of Plant Physiology, 15, 63–92.CrossRefGoogle Scholar
Gleadow, R.M., Foley, W.J. and Woodrow, I.E. (1998). Enhanced CO2 alters the relationship between photosynthesis and defence in cyanogenic Eucalyptus cladocalyx F. Muell. Plant, Cell and Environment, 21, 12–22.CrossRefGoogle Scholar
Glover, J.R. and Lindquist, S. (1998). Hspl04, Hsp70, and Hsp40: A novel chaperone system that rescues previously aggregated proteins. Cell, 94, 73–82.CrossRefGoogle Scholar
Godbold, D.L. and Hüttermann, A. (1988). Inhibition of photosynthesis and transpiration in relation to mercury induced root damage in spruce seedlings. Physiologia Plantarum, 74, 270–275.CrossRefGoogle Scholar
Goerner, A., Reichstein, M. and Rambal, S. (2009). Tracking seasonal drought effects on ecosystem light use efficiency with satellite-based PRI in a Mediterranean forest. Remote Sensing of the Environment, 113, 1101–1111.CrossRefGoogle Scholar
Goffeau, A. and Bové, J.M. (1965). Virus infection and photosyntesis. I. Increased photophosphorylation by chloroplasts from Chinese cabbage infected with turnip yellow mosaic virus. Virology, 27, 243–252.CrossRefGoogle Scholar
Gog, L., Berenbaum, M.R., de Lucia, E.H., et al. (2005). Autotoxic effects of essential oils on photosynthesis in parsley, parsnip, and rough lemon. Chemoecology, 15, 115–119.CrossRefGoogle Scholar
Goh, C.H., Schreiber, U. and Hedrich, R. (1999). New approach of monitoring changes in chlorophyll a fluorescence of single guard cells and protoplasts in response to physiological stimuli. Plant, Cell and Environment, 22, 1057–1070.CrossRefGoogle Scholar
Goicoechea, N., Aguirreolea, J., Cenoz, S., et al. (2001). Gas exchange and flowering in verticillium-wilted pepper plants. Journal of Phytopathology, 149, 281–286.CrossRefGoogle Scholar
Golding, A.J. and Johnson, G.N. (2003). Down-regulation of linear and activation of cyclic electron transport during drought. Planta, 218, 107–14.CrossRefGoogle ScholarPubMed
Goldstein, G. and Nobel, P.S. (1994). Water relations and low temperature acclimation for cactus species varying in freezing tolerance. Plant Physiology, 104, 675–681.CrossRefGoogle ScholarPubMed
Goldstein, G., Sharifi, M.R., Kohorn, L.U., et al. (1991). Photosynthesis by inflated pods of a desert shrub, Isomeris arborea. Oecologia, 85, 396–402.CrossRefGoogle ScholarPubMed
Golem, S. and Culver, J.N. (2003). Tobacco mosaic virus induced alterations in the gene expression profile of Arabidopsis thaliana. Molecular Plant-Microbe Interactions, 16, 681–688.CrossRefGoogle ScholarPubMed
Gollan, T., Schurr, U. and Schulze, E-D. (1992). Stomatal response to drying soil in relation to changes in the xylem sap composition of Helianthus annuus. I. The concentration of cations, anions, amino acids in, and pH of, the xylem sap. Plant, Cell and Environment, 15, 551–559.CrossRefGoogle Scholar
Golluscio, R.A. and Oesterheld, M. (2007). Water use efficiency of twenty five co-existing Patagonian species growing under different soil water availability. Oecologia, 154, 207–217.CrossRefGoogle ScholarPubMed
Gombos, Z., Wada, H., Hideg, E., et al. (1994). The unsaturation of membrane lipids stabilizes photosynthesis against heat stress. Plant Physiology, 104, 563–567.CrossRefGoogle ScholarPubMed
Gomes, F., Oliva, M.A., Mielke, M.S., et al. (2008). Photosynthetic limitations in leaves of young Brazilian Green Dwarf coconut (Cocos nucifera L. ‘nana’) palm under well-watered conditions or recovering from drought stress. Environmental and Experimental Botany, 62, 195–204.CrossRefGoogle Scholar
Gomes, F.P., Oliva, M.A., Mielke, M.S., et al. (2006). Photosynthetic irradiance-response in leaves of dwarf coconut palm (Cocos nucifera L. ‘nana’, Arecaceae): Comparison of three models. Scientia Horticulturae, 109, 101–105.CrossRefGoogle Scholar
Gómez-del-Campo, M., Baeza, P., Ruiz, C., et al. (2007). Effect of previous water conditions on vine response to rewatering. Vitis, 46, 51–55.Google Scholar
Gonfiantini, R., Gratziou, S. and Tongiorgi, E. (1965). Oxygen isotopic composition of water in leaves. In: Isotopes and Radiation in Soil-plant Nutrition Studies. International Atomic Energy Agency, Vienna, 405–410.Google Scholar
González, A., Steffen, K.L. and Lynch, J.P. (1998). Light and excess manganese. Implications for oxidative stress in common bean. Plant Physiology, 118, 493–504.CrossRefGoogle ScholarPubMed
Goodman, R.N., Király, Z. and Wood, K.R. (1986). The Biochemistry and Physiology of Plant Disease. Columbia: University of Missouri Press.Google Scholar
Gordon, A.H., Lomax, J.A., Dalgarno, K., et al. (1985). Preparation and composition of mesophyll, epidermis and fibre cell walls from leaves of perennial ryegrass (Lolium perenne) and italian ryegrass (Lolium multiflorum). Journal of the Science of Food and Agriculture, 36, 509–519.CrossRefGoogle Scholar
Gordon, D.R., Welker, J.M., Menke, J.W., et al. (1989). Competition for soil water between annual plants and blue oak (Quercus douglasii) seedlings. Oecologia, 79, 533–541.CrossRefGoogle ScholarPubMed
Gordon, J.C. and Wheeler, C.T. (1978). Whole plant studies on photosynthesis and acetylene reduction in Alnus glutinosa. New Phytologist, 80, 179–186.CrossRefGoogle Scholar
Gorton, H.L., Herbert, S.K. and Vogelmann, T.C. (2003). Photoacoustic analysis indicates that chloroplast movement does not alter liquid-phase CO2 diffusion in leaves of Alocasia brisbanensis. Plant Physiology, 132, 1529–1539.CrossRefGoogle Scholar
Goudarzi, S. (2006). Scientist Reading the Leaves to Predict Violent Weather. LiveScience, 13 March 2006.Google Scholar
Goudriaan, J., Van Laar, H.H., Van Keulen, H., et al. (1985). Photosynthesis, CO2, and plant production. In: Wheat Growth and Modelling. NATO ASI Series, Series A, Vol 86 (eds Day, W. and Atkin, R.K.), Plenum Press, New York, USA, pp. 107–122.Google Scholar
Goulas, Y., Camenen, L., Guyot, G., et al. (1997). Measurements of laser-induced fluorescence decay and reflectance of plant canopies. Remote Sensing Reviews, 15, 305–322.CrossRefGoogle Scholar
Gould, K.S. (2004). Nature’s swiss army knife: the diverse protective roles of anthocyanins in leaves. Journal of Biomedicine and Biotechnology, 5, 314–320.CrossRefGoogle Scholar
Gould, S.J. and Lewontin, R.C. (1979). The Spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proceedings of the Royal Society of London, 205B, 581–598.CrossRef
Goulden, M.L. (1996). Carbon assimilation and water-use efficiency by neighboring Mediterranean-climate oaks that differ in water access. Tree Physiology, 16, 417–424.CrossRefGoogle ScholarPubMed
Goulden, M.L. and Field, C.B. (1994). Three methods for monitoring the gas exchange of individual tree canopies: ventilated-chamber, sap-flow and Penman-Monteith measurements on evergreen oaks. Functional Ecology, 8, 125–135.CrossRefGoogle Scholar
Goulden, M.L., Miller, S.D., da Rocha, H.R., et al. (2004). Diel and seasonal patterns of tropical forest CO2 exchange. Ecological Applications, 14, S42–S54.CrossRefGoogle Scholar
Goulden, M.L., Munger, J.W., Fan, S.M., et al. (1996). Measurements of carbon sequestration by long-term eddy covariance: methods and critical evaluation of accuracy. Global Change Biology, 2, 169–181.CrossRefGoogle Scholar
Gounaris, K., Brain, A.P.R., Quinn, P.J., et al. (1984). Structural reorganization of chloroplast thylakoid membranes in response to heat stress. Biochimica et Biophysica Acta, 766, 198–208.CrossRefGoogle Scholar
Gout, E., Bligny, R. and Douce, R. (1992). Regulation of intracellular pH values in higher plant cells. Journal of Biological Chemistry, 267, 13903–13909.Google ScholarPubMed
Gout, E., Boisson, A.M., Aubert, S., et al. (2001). Origin of the cytoplasmic pH changes during anaerobic stress in higher plant cells: carbon-13 and phosphorus-31 nuclear magnetic resonance studies. Plant Physiology, 125, 912–925.CrossRefGoogle Scholar
Gouveia, A.C. and Freitas, H. (2009). Modulation of leaf attributes and water use efficiency in Quercus suber along a rainfall gradient. Trees, 23, 267–275.CrossRefGoogle Scholar
Govindachary, S., Bigras, C., Harnois, J., et al. (2007). Changes in the mode of electron flow to photosystem I following chilling-induced photoinhibition in a C3 plant Cucumis sativus L. Photosynthesis Research, 94, 333–345.CrossRefGoogle Scholar
Govindjee, J. (1995). Sixty three years since Kautsky: chlorophyll a fluorescence. Australian Journal of Plant Physiology, 22, 131–160.CrossRefGoogle Scholar
Gower, S.T. (2002). Productivity of terrestrial ecosystems. In: Encyclopedia of Climate Change, vol 2 (eds Mooney, H.A. and Canadell, J.), Blackwell Scientific, Oxford, UK, pp. 516–521.Google Scholar
Gower, S.T., Reich, P.B. and Son, Y. (1993). Canopy dynamics and aboveground production of five tree species with different leaf longevities. Tree Physiology, 12, 327–345.CrossRefGoogle ScholarPubMed
Gowik, U., Burscheidt, J., Akyildiz, M. et al. (2004). Cis-regulatory elements for mesophyll-specific gene expression in the C4 plant Flaveria trinervia, the promoter of the C4 phosphoenolpyruvate carboxylase gene. Plant Cell, 16, 1077–1090.CrossRefGoogle ScholarPubMed
Grabherr, G., Gottfried, M. and Paull, H. (1994). Climate effects on mountain plants. Nature, 369, 448–448.CrossRefGoogle ScholarPubMed
Grace, J. (1990) Cuticular water loss unlikely to explain tree-line in Scotland. Oecologia, 84, 64–68.CrossRefGoogle ScholarPubMed
Grace, J., Berninger, F. and Nagy, L. (2002). Impacts of climate change at the treeline. Annals of Botany, 90, 537–554.CrossRefGoogle Scholar
Grace, J., Jose, J.S., Meir, P., et al. (2006). Productivity and carbon fluxes of tropical savannas. Journal of Biogeography, 33, 387–400.CrossRefGoogle Scholar
Grace, J., Lloyd, J., McIntyre, J., et al. (1995). Carbon-dioxide uptake by an undistrubed tropical rain-forest in Southwest Amazonia, 1992 to 1993. Science, 270, 778–780.CrossRefGoogle Scholar
Grace, J., Lloyd, J., Miranda, A.C., et al. (1998). Fluxes of carbon dioxide and water vapour over a C4 pasture in South Western Amazonia (Brazil). Australian Journal of Plant Physiology, 25, 519–530.CrossRefGoogle Scholar
Grace, J., Nichol, C., Disney, M., et al. (2007). Can we measure terrestrial photosynthesis from space directly, using spectral reflectance and fluorescence?Global Change Biology, 13, 1484–1497.CrossRefGoogle Scholar
Graham, D. and Patterson, B.D. (1982). Responses of plants to low, nonfreezing temperatures: proteins, metabolism, and acclimation. Annual Review of Plant Physiology, 33, 347–372.CrossRefGoogle Scholar
Graham, E.A., Mulkey, S.S., Kitajima, K., et al. (2003). Cloud cover limits net CO2 uptake and growth of a rainforest tree during tropical rainy seasons. Proceedings of the National Academy of Sciences of the United States of America, 100, 572–576.CrossRefGoogle ScholarPubMed
Grainger, J. and Ring, J. (1962). Anomalous Fraunhofer lines profiles. Nature, 193, 762–764.CrossRefGoogle Scholar
Grams, T.E.E., Borland, A.M., Roberts, A., et al. (1997). On the mechanism of reinitiation of endogenous crassulacean acid metabolism by temperature changes. Plant Physiology, 113, 1309–1317.CrossRefGoogle ScholarPubMed
Grams, T.E.E., Koziolek, C., Lautner, S., et al. (2007). Distinct roles of electric and hydraulic signals on the reaction of leaf gas exchange upon re-irrigation in Zea mays L. Plant, Cell and Environment, 30, 79–84.CrossRefGoogle ScholarPubMed
Granier, A., Aubinet, M., Epron, D., et al. (2003). Deciduous forests: carbon and water fluxes, balances, ecological and ecophysiological determinants. In: Fluxes of Carbon, Water and Energy of European Forests, Ecological Studies 163 (ed. Valentini, R.), Springer-Verlag, Berlin Heidelberg, Germany, pp. 55–70.Google Scholar
Granier, A., Biron, P., Bréda, N., et al. (1996). Transpiration of trees and forest stands: short and long-term monitoring using sapflow methods. Global Change Biology, 2, 265–274.CrossRefGoogle Scholar
Granier, A., Reichstein, M., Bréda, N., et al. (2007) Evidence for soil water control on carbon and water dynamics in European forests during the extremely dry year: 2003. Agricultural and Forest Meteorology, 143, 123–145.CrossRefGoogle Scholar
Granier, C. and Tardieu, F. (1999). Leaf expansion and cell division are affected by reducing absorbed light before but not after the decline in cell division rate in the sunflower leaf. Plant Cell Environment, 22, 1365–1376.CrossRefGoogle Scholar
Granier, C., Turc, O. and Tardieu, F. (2000). Co-ordination of cell division and tissue expansion in sunflower, tobacco, and pea leaves: dependence or independence of both processes?Journal of Plant Growth Regulation, 19, 45–54.CrossRefGoogle ScholarPubMed
Grant, O.M., Tronina, L., Jones, H.G., et al. (2007). Exploring thermal imaging variables for the detection of stress responses in grapevine under different irrigation regimes. Journal of Experimental Botany, 58, 815–825.CrossRefGoogle ScholarPubMed
Grant, O.M., Tronina, Ł., Ramalho, J.C., et al. (2010). The impact of drought on leaf physiology of Quercus suber L. trees: comparison of an extreme drought event with chronic rainfall reduction. Journal of Experimental Botany, 61, 4361–4371.CrossRefGoogle ScholarPubMed
Grantz, D.A. and Assmann, S.M. (1991). Stomatal response to blue-light-water-use efficiency in sugarcane and soybean. Plant, Cell and Environment, 14, 683–690.CrossRefGoogle Scholar
Grassi, G. and Magnani, F. (2005). Stomatal, mesophyll conductance and biochemical limitations to photosynthesis as affected by drought and leaf ontogeny in ash and oak trees. Plant, Cell and Environment, 28, 834–849.CrossRefGoogle Scholar
Grassi, G., Ripullone, F., Borghetti, M., et al. (2009). Contribution of diffusional and non-diffusional limitations to midday depression of photosynthesis in Arbutus unedo L. Trees, 23, 1149–1161.CrossRefGoogle Scholar
Grassi, G., Vicinelli, E., Ponti, F., et al. (2005). Seasonal and interannual variability of photosynthetic capacity in relation to leaf nitrogen in a deciduous forest plantation in northern Italy. Tree Physiology, 25, 349–360.CrossRefGoogle Scholar
Gratani, L. and Bombelli, A. (2000). Correlation between leaf age and other leaf traits in three Mediterranean maquis shrub species: Quercus ilex, Phillyrea latifolia and Cistus incanus. Environmental and Experimental Botany, 43, 141–153.CrossRefGoogle Scholar
Gratani, L. and Ghia, E. (2002). Adaptive strategy at the leaf level of Arbutus unedo L. to cope with Mediterranean climate. Flora, 197, 275–284.CrossRefGoogle Scholar
Graves, J.D., Press, M.C. and Steward, G.R. (1989). A carbon balance model of the sorghum-Striga hermonthica host-parasite association. Plant, Cell and Environment, 12, 101–107.CrossRefGoogle Scholar
Gray, G.R. and Heath, D. (2005). A global reorganization of the metabolome in Arabidopsis during cold acclimation is revealed by metabolic fingerprinting. Physiologia Plantarum, 124, 236–248.CrossRefGoogle Scholar
Gray, G.R., Ivanov, A.G., Krol, M., et al. (1998). Adjustment of thylakoid plastoquinone content and electron donor pool size in response to growth temperature and growth irradiance in Winter Rye (Secale cereale L.). Photosynthesis Research, 56, 209–221.CrossRefGoogle Scholar
Gray, G.R., Chauvin, L.P., Sarhan, F., et al. (1997). Cold acclimation and freezing tolerance. A complex interaction of light and temperature. Plant Physiology, 114, 467–474.CrossRefGoogle ScholarPubMed
Gray, J.E., Holroyd, G.H., van der Lee, F.M., et al. (2000). The HIC signaling pathway links CO2 perception to stomatal development. Nature, 408, 713–716.CrossRefGoogle ScholarPubMed
Greer, D.H. (1996). Photosynthetic development in relation to leaf expansion in kiwifruit (Actinidia deliciosa) vines during growth in a controlled environment. Australian Journal of Plant Physiology, 23, 541–549.CrossRefGoogle Scholar
Greer, D.H. and Sicard, S.M. (2009). The net carbon balance in relation to growth and biomass accumulation of grapevines (Vitis vinifera cv. Semillon) grown in a controlled environment. Functional Plant Biology, 36, 645–653.CrossRefGoogle Scholar
Greer, D.H., Seleznyova, A.N. and Green, S.R. (2004). From controlled environments to field simulations: leaf area dynamics and photosynthesis in kiwifruit vines (Actinidia deliciosa). Functional Plant Biology, 31, 169–179.CrossRefGoogle Scholar
Greger, M. and Johansson, M. (1992). Cadmium effects on leaf transpiration of sugar beet (Beta vulgaris). Physiologia Plantarum, 86, 465–473.CrossRefGoogle Scholar
Greger, M. and Ögren, E. (1991). Direct and indirect effects of Cd2+ on photosynthesis in sugar beet (Beta vulgaris). Physiologia Plantarum, 83, 129–135.CrossRefGoogle Scholar
Gregg, J.W., Jones, C.G. and Dawson, T.E. (2003). Urbanization effects on tree growth in the vicinity of New York City. Nature, 424, 183–187.CrossRefGoogle ScholarPubMed
Grewe, V., Dameris, M., Hein, R., et al. (2001). Future changes of the atmospheric composition and the impact of climate change. Tellus B, 53, 103–121.CrossRefGoogle Scholar
Griffin, K.L., Anderson, O.R., Gastrich, M.D., et al. (2001). Plant growth in elevated CO2 alters mitochondrial number and chloroplast fine structure. Proceedings of the National Academy of Sciences USA, 98, 2473–2478.CrossRefGoogle Scholar
Griffin, K.L., Ross, P.D., Sims, D.A., et al. (1996). EcoCELLs: tools for mesocosm scale measurements of gas exchange. Plant Cell Environment, 19, 1210–1221.CrossRefGoogle ScholarPubMed
Griffis, T.J., Sargent, S.D., Baker, J.M., et al. (2008). Direct measurements of biosphere-atmosphere isotopic CO2 exchange using the eddy covariance technique. Journal of Geophysical Research, 113, D08304.CrossRefGoogle Scholar
Griffith, H. and Jarvis, P. (edit) (2005). The Carbon Balance of Forest Biomes. SEB series vol 57. Taylor & Francis, Oxford, UK, pp. 1–720.
Griffiths, H. (1989). Carbon dioxide concentrating mechanisms and the evolution of CAM in vascular epiphytes. In: Vascular Plants as Epiphytes: Evolution and Ecophysiology (ed. Lüttge, U), Springer Verlag, Berlin, Germany, pp. 42–86.Google Scholar
Griffiths, H. (1992). Carbon isotope discrimination and the integration of carbon assimilation pathways in terrestrial CAM plants. Plant, Cell and Environment, 15, 1051–1062.CrossRefGoogle Scholar
Griffiths, H., Broadmeadow, M.S.J., Borland, A.M., et al. (1990). Short-term changes in carbon-isotope discrimination identify transitions between C3 and C4 carboxylation during crassulacean acid metabolism. Planta, 181, 604–610.CrossRefGoogle ScholarPubMed
Griffiths, H., Cousins, A.B., Badger, M., et al. (2007). Discrimination in the dark. Resolving the interplay between metabolic and physical constraints to phosphoenolpyruvate carboxylase activity during the crassulacean acid metabolism cycle. Plant Physiology, 143, 1055–1067.CrossRefGoogle ScholarPubMed
Groom, Q.J., Baker, N.R. and Long, S.P. (1991). Photoinhibition of holly (Ilex aquifolium) in the field during the winter. Physiologia Plantarum, 83, 585–590.CrossRefGoogle Scholar
Grünzweig, J.M. and Körner, C. (2001). Growth, water and nitrogen relations in grassland model ecosystems of the semi-arid Negev of Israel exposed to elevated CO2. Oecologia, 128, 251–262.CrossRefGoogle ScholarPubMed
Grünzweig, J.M., Lin, T., Rotenberg, E., et al. (2003). Carbon sequestration in arid land forest. Global Change Biology, 9, 791–799.CrossRefGoogle Scholar
Grzesiak, M.T., Grzesiak, S. and Skoczowski, A. (2006). Changes of leaf water potential and gas exchange during and after drought in triticale and maize genotypes differing in drought tolerance. Photosynthetica, 44, 561–568.CrossRefGoogle Scholar
Gu, L., Baldocchi, D., Verma, S.B., et al. (2002). Advantages of diffuse radiation for terrestrial ecosystem productivity. Journal of Geophysical Research, 107, 1–23.CrossRefGoogle Scholar
Gu, L., Baldocchi, D.D., Wofsy, S.C., et al. (2003). Response of a deciduous forest to the Mount Pinatubo eruption: enhanced photosynthesis. Science, 299, 2035–2038.CrossRefGoogle ScholarPubMed
Gu, L., Hanson, P.J., Post, W.M., et al. (2008). The 2007 Eastern US Spring Freeze: Increased Cold Damage in a Warming World?Bioscience, 58, 253–262.CrossRefGoogle Scholar
Guanter, L., Alonso, L., Gomez-Chova, L., et al. (2010), Developments for vegetation fluorescence retrieval from spaceborne high-resolution spectrometry in the O2A and O2B absorption bands, J. Geophys. Res., 115, D19303, doi:.
Guderian, R. (1986). Terrestrial ecosystems: particulate deposition. In: Air Pollutants and their Effects on the Terrestrial Ecosystem. Advances in Environmental Science and Technology (eds Legge, A.H. and Krupa, S.V.), Wiley, New York, USA, pp. 339–363.Google Scholar
Guehl, J.M., Domenacg, A.M., Bereau, M., et al. (1998). Functional diversity in an Amazonian rainforest of Fench Guyana: a dual isotope approach (δ15N and δ 13C). Oecologia, 116, 316–330.CrossRefGoogle Scholar
Guenther, A., Hewitt, C.N., Erickson, D., et al. (1995). A global model of natural volatile compound emissions. Journal of Geophysical Research, 100, 8873–8892.CrossRefGoogle Scholar
Guenther, A.B., Zimmerman, P.R. and Harley, P.C. (1993). Isoprene and monoterpene emission rate variability: Model evaluations and sensitivity analysis. Journal of Geophysical Research, 98, 12,609–12,617.CrossRefGoogle Scholar
Guida dos Santo, M., Vasconcelos Ribeiro, R., Ferraz de Oliveira, R., et al. (2006). The role of inorganic phosphate on photosynthesis recovery of common bean after a mild water deficit. Plant Science, 170, 659–664.CrossRefGoogle Scholar
Guidi, L., Degl’Innocenti, E. and Soldatini, G.F., (2002). Assimilation of CO2, enzyme activation and photosynthetic electron transport in bean leaves as affected by high light and ozone. New Phytologist, 156, 377–388.CrossRefGoogle Scholar
Guissé, B., Srivastava, A. and Strasser, R.J. (1995). The polyphasic rise of the chlorophyll-a fluorescence (O-K-J-I-P) in heat-stressed leaves. Archives des Sciences, 48, 147–160.Google Scholar
Gulías, J., Cifre, J., Jonasson, S., et al. (2009). Seasonal and inter-annual variations of gas exchange in thirteen woody species along a climatic gradient in the Mediterranean island of Mallorca. Flora, 204, 169–181.CrossRefGoogle Scholar
Gulías, J., Flexas, J., Abadía, A., et al. (2002). Photosynthetic responses to water deficit in six Mediterranean sclerophyll species: possible factors explaining the declining of Rhamnus ludovici-salvatoris, an endemic Balearic species. Tree Physiology, 22, 687–697.CrossRefGoogle ScholarPubMed
Gulías, J., Flexas, J., Mus, M., et al. (2003). Relationship between maximum leaf photosynthesis, nitrogen content and specific leaf area in balearic endemic and non-endemic Mediterranean species. Annals of Botany, 92, 215–222.CrossRefGoogle Scholar
Gunasekera, D. and Berkowitz, G.A. (1992). Heterogeneous stomatal closure in response to leaf water deficits is not a universal phenomenon. Plant Physiology, 98, 660–665.CrossRefGoogle Scholar
Gunawardena, A.H.L.A.N. (2008). Programmed cell death and tissue remodelling in plants. Journal of Experimental Botany, 59, 445–451.CrossRefGoogle ScholarPubMed
Gunderson, C.A. and Wullschleger, S.D. (1994). Photosynthetic acclimation in trees to rising atmospheric CO2: a broader perspective. Photosynthesis Research, 39, 369–388.CrossRefGoogle ScholarPubMed
Gunderson, C.A., Norby, R.J. and Wullschleger, S.D. (2000). Acclimation of photosynthesis and respiration to simulated climatic warming in northern and southern populations of Acer saccharum: laboratory and field evidence. Tree Physiology, 20, 87–96.CrossRefGoogle ScholarPubMed
Gunderson, C.A., Sholtis, J.D., Wullschleger, S.D., et al. (2002). Environmental and stomatal control of photosynthetic enhancement in the canopy of a sweetgum (Liquidambar styraciflua L.) plantation during 3 years of CO2 enrichment. Plant, Cell and Environment, 25, 379–393.CrossRefGoogle Scholar
Gunnell, G.F., Morgan, M.E., Maas, M.C., et al. (1995). Comparative palaeoecology of Paleogene and Neogene mammalian faunas: trophic structure and composition. Palaeogeography, Palaeoclimatology, Palaeoecology, 115, 265–286.CrossRefGoogle Scholar
Günthardt-Goerg, M.S. and Vollenweider, P. (2007). Linking stress with macroscopic and microscopic leaf response in trees: new diagnostic perspectives. Environmental Pollution, 147, 467–488.CrossRefGoogle ScholarPubMed
Günthardt-Goerg, M.S., McQuattie, C.J., Scheidegger, C., et al. (1997). Ozone induced cytochemical and ultrastructural changes in leaf mesophyll cell walls. Canadian Journal of Forest Research, 27, 453–463.CrossRefGoogle Scholar
Guo, J. and Trotter, C.M. (2004). Estimating photosynthetic light-use efficiency using the photochemical reflectance index: variations among species. Functional Plant Biology, 31, 255–265.CrossRefGoogle Scholar
Guo, S.J., Zhou, H.Y., Zhang, X.S., et al. (2007). Overexpression of CaHSP26 in transgenic tobacco alleviates photoinhibition of PSII and PSI during chilling stress under low irradiance. Journal of Plant Physiology, 164, 126–136.CrossRefGoogle ScholarPubMed
Guralnick, L.J. and Jackson, M.D. (2001). The occurrence and phylogenetics of crassulacean acid metabolism in the Portulacaceae. International Journal of Plant Science, 162, 257–262.CrossRefGoogle Scholar
Guralnick, L.J., Ting, I.P. and Lord, E.M. (1986). Crassulacean acid metabolism in the Gesneriaceae. American Journal of Botany, 53, 336–345.CrossRefGoogle Scholar
Gurney, A.L., Press, M.C. and Ransom, J.K. (1995). The parasitic angiosperm Striga hermonthica can reduce photosynthesis of its sorghum and maize hosts in the field. Journal of Experimental Botany, 46, 1817–1823.CrossRefGoogle Scholar
Güsewell, S. (2004) Tansley review. N : P ratios in terrestrial plants: variation and functional significance. The New Phytologist, 164, 243–266.CrossRefGoogle Scholar
Gustafsson, H.G., Winter, K. and Bittrich, V. (2007). Diversity, phylogeny and classification of Clusia. In: Clusia: A Woody Neotropical Genus of Remarkable Plasticity and Diversity. Ecological Studies, Vol. 194 (ed. Lüttge, U.), Springer-Verlag, Berlin Heidelberg, Germany, pp. 95–116.Google Scholar
Gutschick, V.P. and BassiriRad, H. (2003). Tansely review: Extreme events as shaping physiology, ecology, and evolution of plants: toward a unified definition and evaluation of their consequences. The New Phytologist, 160, 21–42.CrossRefGoogle Scholar
Gutschick, V.P. and Simonneau, T. (2002). Modelling stomatal conductance of field-grown sunflower under varying soil water content and leaf environment: comparison of three models of stomatal response to leaf environment and coupling with an abscisic acid based model of stomatal response to soil drying. Plant, Cell and Environment, 25, 1423–1434.CrossRefGoogle Scholar
Gutschick, V.P. and Wiegel, F.W. (1988). Optimizing the canopy photosynthetic rate by patterns of investment in specific leaf mass. American Naturalist, 132, 67–86.CrossRefGoogle Scholar
Guy, C.L., Carter, J.V., Yelenosky, G., et al. (1984). Changes in glutathione content during cold acclimation in Cornus sericea and Citrus sinensis. Cryobiology, 21, 443–453.CrossRefGoogle Scholar
Guy, R.D., Berry, J.A., Fogel, M.L., et al. (1989). Differential fractionation of oxygen isotopes by cyanide-resistant and cyanide-sensitive respiration in plants. Planta, 177, 483–491.CrossRefGoogle ScholarPubMed
Guy, R.D., Fogel, M.L. and Berry, J.A. (1993). Photosynthetic fractionation of the stable isotopes of oxygen and carbon. Plant Physiology, 101, 37–47.CrossRefGoogle ScholarPubMed
Guy, R.D., Fogel, M.L., Berry, J.A., et al. (1987). Isotope fractionation during oxygen production and consumption by plants. In: Progress in Photosynthesis Research III (ed. Biggins, J.), Springer, Dordrecht, Netherlands, pp. 597–600.Google Scholar
Gwathmey, C.O., Hall, A.E. and Madore, M.A. (1992). Pod removal effects on cowpea genotypes contrasting in monocarpic senescence traits. Crop Science, 32, 1003–1009.CrossRefGoogle Scholar
Gyenge, J., Fernandez, M.E., Sarasola, M., et al. (2008). Testing a hypothesis of the relationship between productivity and water use efficiency in Patagonian forests with native and exotic species. Forest Ecology and Management, 255, 3281–3287.CrossRefGoogle Scholar
Haase, P., Pugnaire, F.J., Clark, S.C., et al. (1999a). Diurnal and seasonal changes in cladode photosynthetic rate in relation to canopy age structure in the leguminous shrub Retama sphaerocarpa. Functional Ecology, 14, 640–649.CrossRefGoogle Scholar
Haase, P., Pugnaire, P.I., Clark, S.C., et al. (1999b). Environmental control of canopy dynamics and photosynthetic rate in the evergreen tussock grass Stipa tenacissima. Plant Ecology, 145, 327–339.CrossRefGoogle Scholar
Habash, D., Percival, M.P. and Baker, N.R. (1985). Rapid chlorophyll fluorescence technique for the study of penetration of photosynthetically active herbidies into leaf tissue. Weed Research, 25, 389–395.CrossRefGoogle Scholar
Habermann, G., Machado, E.C., Rodrigues, J.D. et al. (2003a). Gas exchange rates at different vapour pressure deficits and water relations of ‘Pera’ sweet orange plants with citrus variegated chlorosis (CVC). Scientia Horticulturae, 98, 233–245.CrossRefGoogle Scholar
Habermann, G., Machado, E.C., Rodrigues, J.D., et al. (2003b). CO2 assimilation, photosynthetic light response curves, and water relations of ‘Pêra’ sweet orange plants infected with Xylella fastidiosa. Brazilian Journal of Plant Physiology, 15, 79–87.CrossRefGoogle Scholar
Haehnel, W. (1984). Photosynthetic electron transport in higher plants. Annual Reviews of Plant Physiology and Plant Molelcular Biology, 35, 659–683.CrossRefGoogle Scholar
Hafke, J.B., Hafke, Y., Smith, J.A.C., et al. (2003). Vacuolar malate uptake is mediated by an anion-selective inward rectifier. The Plant Journal, 35, 116–128.CrossRefGoogle ScholarPubMed
Haile, F.J. and Higley, L.G. (2003). Changes in soybean gas exchange after moisture stress and spider mite injury. Environmental Entomology, 32, 433–440.CrossRefGoogle Scholar
Hajirezaei, M., Sonnewald, U., Viola, R., et al. (1994). Transgenic potato plants with strongly decreased expression of PFP show no visible phenotype and only minor changes in metabolic fluxes in their tubers. Planta, 192, 16–30.Google Scholar
Hald, S., Nandha, B., Gallois, P., et al. (2007). Feedback regulation of photosynthetic electron transport by NADP(H) redox poise. Biochimica et Biophysica Acta, 1777, 433–440.CrossRefGoogle Scholar
Haldimann, P. and Feller, U. (2004). Inhibition of photosynthesis by high temperature in oak (Quercus pubescens L.) leaves grown under natural conditions closely correlates with a reversible heat-dependent reduction of the activation state of ribulose-1,5-bisphosphate carboxylase/oxygenase. Plant, Cell and Environment, 27, 1169–1183.CrossRefGoogle Scholar
Haldimann, P., Gallé, A. and Feller, U. (2008). Impact of an exceptionally hot dry summer on photosynthetic traits in oak (Quercus pubescens) leaves. Tree Physiology, 28, 785–795.CrossRefGoogle ScholarPubMed
Haldrup, A., Jensen, P.E., Lunde, C., et al. (2001). Balance of power: a view of the mechanism of photosynthetic state transitions. Trends in Plant Science, 6, 301–305.CrossRefGoogle ScholarPubMed
Hales, S. (1727). Vegetable Staticks, or an Account of Some Statistical Experiments on the Sap in Vegetation. W. Innys, London.
Hall, A.E. and Schulze, E.D. (1980). Stomatal responses to environment and a possible interrelation between stomatal effects on transpiration and CO2 assimilation. Plant, Cell and Environment, 3, 467–474.Google Scholar
Hall, D.A., Ptacek, J.J. and Snyder, M.M. (2007). Protein microarray technology. Mechanisms of Ageing and Development, 128, 161–167.CrossRefGoogle ScholarPubMed
Hall, N.M., Griffiths, H., Corlett, J.A., et al. (2005). Relationships between water-use traits and photosynthesis in Brassica oleracea resolved by quantitative genetic analysis. Plant Breeding, 124, 557–564.CrossRefGoogle Scholar
Hallik, L., Kull, O., Niinemets, Ü., et al. (2009). Contrasting correlation networks between leaf structure, nitrogen and chlorophyll in herbaceous and woody canopies. Basic and Applied Ecology, 10, 309–318.CrossRefGoogle Scholar
Halliwell, B. and Whiteman, M. (2004). Measuring reactive species and oxidative damage in vivo and in cell culture: how should you do it and what do the results mean?British Journal of Pharmacology, 142, 231–255.CrossRefGoogle Scholar
Halsted, M. and Lynch, J. (1996). Phosphorus responses of C3 and C4 species. Journal of Experimental Botany, 47, 497–505.CrossRefGoogle Scholar
Ham, J., Owensby, C., Coyne, P., et al. (1995). Fluxes of CO2 and water vapor from a prairie ecosystem exposed to ambient and elevated atmospheric CO2. Agricultural and Forest Meteorology, 77, 73–93.CrossRefGoogle Scholar
Ham, J.M. and Knapp, A.K. (1998). Fluxes of CO2, water vapor, and energy from a prairie ecosystem during the seasonal transition from carbon sink to carbon source. Agricultural and Forest Meteorology, 89, 1–14.CrossRefGoogle Scholar
Ham, J.M., Owensby, C.E. and Coyne, P.I. (1993). Technique for measuring air flow and carbon dioxide flux in large open-top chambers. Journal of Environmental Quality, 22, 759–766.CrossRefGoogle Scholar
Hamerlynck, E.P. and Knapp, A.K. (1994). Leaf-level responses to light and temperature in two co-occurring Quercus (Fagaceae) species: implications for tree distribution patterns. Forest Ecology and Management, 68, 149–159.CrossRefGoogle Scholar
Hamerlynck, E.P., Huxman, T.E., Loik, M.E., et al. (2000). Effects of extreme high temperature, drought and elevated CO2 on photosynthesis of the Mojave Desert evergreen shrub, Larrea tridentata. Plant Ecology, 148, 183–93.CrossRefGoogle Scholar
Han, H., Li, T. and Zhou, S. (2008). Overexpression of phytoene synthase gene from Salicornia europaea alters response to reactive oxygen species under salt stress in transgenic Arabidopsis. Biotechnology Letters, 30, 1501–1507.CrossRefGoogle ScholarPubMed
Han, Q. (2011). Height-related decreases in mesophyll conductance, leaf photosynthesis and compensating adjustments associated with leaf nitrogen concentrations in Pinus densiflora. Tree Physiology, 31, 976–984.CrossRef
Han, T., Vogelmann, T. and Nishio, J. (1999). Profiles of photosynthetic oxygen-evolution within leaves of Spinacia oleracea. The New Phytologist, 143, 83–92.CrossRefGoogle Scholar
Hanan, N.P., Berry, J.A., Verma, S.B., et al. (2005). Testing a model of CO2, water and energy exchange in Great Plains tallgrass prairie and wheat ecosystems. Agricultural and Forest Meteorology, 131, 162–179.CrossRefGoogle Scholar
Hanba, Y.T., Miyazawa, S.I., Kogami, H., et al. (2001). Effects of leaf age on internal CO2 transfer conductance and photosynthesis in tree species having different types of shoot phenology. Australian Journal Plant Physiology, 28, 1075–1084.Google Scholar
Hanba, Y.T., Shibasaka, M., Hayashi, Y., et al. (2004). Overexpression of the barley aquaporin HvPIP2;1 increases internal CO2 conductance and CO2 assimilation in the leaves of transgenic rice plants. Plant and Cell Physiology, 45, 521–529.CrossRefGoogle Scholar
Handa, I.T., Körner, C. and Hättenschweiler, S. (2005). A test of the tree-line carbon limitation hypothesis by in situ CO2 enrichment and defoliation. Ecology, 86, 1288–1300.CrossRefGoogle Scholar
Handford, M.G. and Carr, J.P. (2007). A defect in carbohydrate metabolism ameliorates symptom severity in virus-infected Arabidopsis thaliana. Journal of General Virology, 88, 337–341.CrossRefGoogle ScholarPubMed
Hänninen, H. (1989). Modelling bud dormancy release in trees from cool and temperate regions. Acta Forestalia Fennica, 213, 1–47.Google Scholar
Hanning, I. and Heldt, H.W. (1993). On the function of mitochondrial metabolism during photosynthesis in spinach (Spinacia oleracea L.) leaves. Plant Physiology, 103, 1147–1154.CrossRefGoogle ScholarPubMed
Hansen, H.C. (1959). Der Einfluß des Lichtes auf die Bildung von Licht- und Schattenblättern der Buche, Fagus silvatica. Physiologia Plantarum, 12, 545–550.CrossRefGoogle Scholar
Hansen, J. and Moller, I.B. (1975). Percolation of starch and soluble carbohydrates from plant tissue for quantitative determination. Biochemistry, 68, 87–94.Google ScholarPubMed
Hanson, P.J., McRoberts, R.E., Isebrands, J.G., et al. (1987). An optimal sampling strategy for determining CO2 exchange rate as a function of photosynthetic photon-flux density. Photosynthetica, 21, 98–101.Google Scholar
Harazono, Y., Mano, M., Zulueta, R.Y., et al. (2003). Inter-annual carbon dioxide uptake of a wet sedge tundra ecosystem in the Arctic. Tellus, 55B, 215–231.CrossRefGoogle Scholar
Harbinson, J. and Foyer, C.H. (1991). Relationships between the efficiencies of photosystems I and II and stromal redox state in CO2-free air: evidence for cyclic electron flow in Vivo. Plant Physiology, 97, 41–49.CrossRefGoogle Scholar
Harbinson, J. and Hedley, C.L. (1989). The kinetics of P700+ reduction in leaves: a novel in situ probe of thylakoid functioning. Plant, Cell and Environment, 12, 357–369.CrossRefGoogle Scholar
Harbinson, J. and Woodward, F.I. (1987). The use of light-induced absorbance changes at 820 nm to monitor the oxidation state of p-700 in leaves. Plant, Cell and Environment, 10, 131–140.Google Scholar
Hardie, D.G. (2007). AMP-activated / SNF1 protein kinases: conserved guardians of cellular energy. National Review of Molecular Cell Biology, 8, 774–785.CrossRefGoogle ScholarPubMed
Hari, P., Keronen, P., Bäck, J., et al. (1999). An improvement of the method for calibrating measurements of photosynthetic CO2 flux. Plant, Cell and Environment, 22, 1297–1301.CrossRefGoogle Scholar
Harley, P.C., Loreto, F., Di Marco, G., et al. (1992a). Theoretical considerations when estimating the mesophyll conductance to CO2 flux by the analysis of the response of photosynthesis to CO2. Plant Physiology, 98, 1429–1436.CrossRefGoogle Scholar
Harley, P.C. and Baldocchi, D.D. (1995). Scaling carbon dioxide and water vapour exchange from leaf to canopy in a deciduous forest. I. Leaf model parametrization. Plant Cell Environment, 18, 1146–1156.CrossRefGoogle Scholar
Harley, P.C. and Sharkey, T.D. (1991). An improved model of C3 photosynthesis at high CO2: reversed O2 sensitivity explained by lack of glycerate reentry into the chloroplast. Photosynthesis Research, 27, 169–178.Google ScholarPubMed
Harley, P.C., Thomas, R.B., Reynolds, J.F., et al. (1992b). Modelling photosynthesis of cotton grown in elevated CO2. Plant, Cell and Environment, 15, 271–282.CrossRefGoogle Scholar
Harley, P.C., Weber, J.A. and Gates, D.M. (1985). Interactive effects of light, leaf temperature, CO2 and O2 on photosynthesis in soybean. Planta, 165, 249–263.CrossRefGoogle Scholar
Harms, K.E. and Dalling, J.W. (1997). Damage and herbivory tolerance through resprouting as an advantage of large seed size in tropical trees and lianas. Journal of Tropical Ecology, 13, 617–621.CrossRefGoogle Scholar
Harris, S., Tapper, N., Packham, D., et al. (2008). The relationship between the monsoonal summer rain and dry season fire activity of northern Australia. International Journal of Wildland Fire, 17, 674–684.CrossRefGoogle Scholar
Harrison, M.T., Edwards, E.J., Farquhar, G.D., et al. (2009). Nitrogen in cell walls of sclerophyllous leaves accounts for little of the variation in photosynthetic nitrogen-use efficiency. Plant, Cell and Environment, 32, 259–270.CrossRefGoogle ScholarPubMed
Hartung, W. (2010). The evolution of abscisic acid (ABA) and ABA function in lower plants, fungi and lichen. Functional Plant Biology, 37, 806–812.CrossRefGoogle Scholar
Haselwandter, K., Hofmann, A., Holzmann, H.P., et al. (1983). Availability of nitrogen and phosphorus in the nival zone of the Alps. Oecologia, 57, 266–269.CrossRefGoogle ScholarPubMed
Hashimoto, M., Negi, J., Young, J., et al. (2006). Arabidopsis HT1 kinase controls stomatal movements in response to CO2. Nature Cell Biology, 8, 391–397.CrossRefGoogle Scholar
Haslam, E. (1985). Metabolites and Metabolism: A Commentary on Secondary Metabolism. Clarendon Press, Oxford, UK.Google Scholar
Haslam, E. (1986). Secondary metabolism – fact and fiction. Natural Product Reports, 4, 217–249.CrossRefGoogle Scholar
Hassiotou, F., Evans, J.R., Ludwig, M., et al. (2009b). Stomatal crypts may facilitate diffusion of CO2 to adaxial mesophyll cells in thick sclerophylls. Plant, Cell and Environment, 32, 1596–1611.CrossRefGoogle Scholar
Hassiotou, F., Ludwig, M., Renton, M., et al. (2009a). Influence of leaf dry mass per area, CO2 and irradiance on mesophyll conductance in sclerophylls. Journal of Experimental Botany, 60, 2303–2314.CrossRefGoogle ScholarPubMed
Hassiotou, F., Renton, M., Ludwig, M., et al. (2010). Photosynthesis at an extreme end of the leaf trait spectrum: how does it relate to high leaf dry mass per area and associated structural parameters?Journal of Experimental Botany, 61, 3015–3028.CrossRefGoogle ScholarPubMed
Hastings, S.J., Oechel, W.C. and Sionit, N. (1989). Water relations and photosynthesis of chaparral resprouts and seedlings following fire and hand clearing. In: The California Chaparral. Paradigms Re-examined (ed. Keeley, S.C.), Natural History Museum of Los Angeles County, Los Angeles, USA, pp. 107–113.Google Scholar
Hatch, M.D. and Slack, C.R. (1966). Photosynthesis by sugarcane leaves: a new carboxylation reaction and the pathway of sugar formation. Biochemical Journal, 101, 103–111.CrossRefGoogle Scholar
Hatch, M.D., Agostino, A. and Jenkins, C.L.D. (1995). Measurements of the leakage of CO2 from bundle-sheath cells of leaves during C4 photosynthesis. Plant Physiology, 108, 173–181.CrossRefGoogle Scholar
Hatier, J.H.B. and Gould, K.S. (2007). Black colouration of Ophiopogon planiscapus “nigrescens”. Leaf optics, chromaticity, and internal light gradients. Functional Plant Biology, 34, 130–138.CrossRefGoogle Scholar
Hättenschwiler, S. (2001). Tree seedling growth in natural deep shade: functional traits related to interspecific variation in response to elevated CO2. Oecologia, 129, 31–42.CrossRefGoogle ScholarPubMed
Hattersley, P.W. (1982). δ13C values of C4 types in grasses. Australian Journal of Plant Physiology, 9, 139–154.CrossRefGoogle Scholar
Hattersley, P.W., Watson, L. and Osmond, C.B. (1977). In situ immunofluorescent labelling of ribulose-1,5-bisphosphate carboxylase in leaves of C3 and C4 plants. Australian Journal of Plant Physiology, 4, 523–39.CrossRefGoogle Scholar
Hauben, M., Haesendonckx, B., Standaert, E., et al. (2009). Energy use efficiency is characterized by an epigenetic component that can be directed through artificial selection to increase yield. Proceedings of the National Academy of Sciences USA, 106, 20109–20114.CrossRefGoogle ScholarPubMed
Haug, A. (1984). Molecular aspects of aluminum toxicity. Critical Reviews in Plant Sciences, 1, 345–373.CrossRefGoogle Scholar
Haughn, G.W. and Gilchrist, E.J. (2006). TILLING in the botanical garden: a reverse genetic technique feasible for all plant species. Floriculture, Ornamental and Plant Biotechnology, 1, 476–482.Google Scholar
Hauglustaine, D. and Brasseur, G.P. (2001). Evolution of tropospheric ozone under anthropogenic activities and associated radiative forcing of climate. Journal of Geophysical Research D. Atmospheres, 106, 32,337–32,360.CrossRefGoogle Scholar
Haupt, W. and Scheuerlein, R. (1990). Chloroplast movement. Plant, Cell and Environment, 13, 595–614.CrossRefGoogle Scholar
Haupt-Herting, S., Klug, K. and Fock, H. (2001). A new approach to measure gross CO2 fluxes in leaves. Gross CO2 assimilation, photorespiration, and mitochondrial respiration in the light in tomato under drought stress. Plant Physiology, 126, 388–396.CrossRefGoogle ScholarPubMed
Havaux, M. (1992). Stress tolerance of photosystem II in vivo. Antagonistic effects of water, heat and photoinhibition stresses. Plant Physiology, 100, 424–432.CrossRefGoogle ScholarPubMed
Havaux, M. (1993). Rapid photosynthetic adaptation to heat stress triggered in potato leaves by moderately elevated temperatures. Plant, Cell and Environment, 16, 461–467.CrossRefGoogle Scholar
Havaux, M. (1996). Short-term responses of photosystem I to heat stress – Induction of a PS II-independent electron transport through PS I fed by stromal components. Photosynthesis Research, 47, 85–97.CrossRefGoogle ScholarPubMed
Havaux, M. (1998). Carotenoids as membrane stabilizers in chloroplasts. Trends in Plant Science, 3, 147–151.CrossRefGoogle Scholar
Havaux, M. (2003). Spontaneous and thermoinduced photon emission: new methods to detect and quantify oxidative stress in plants. Trends Plant Science, 8, 409–413.CrossRefGoogle ScholarPubMed
Havaux, M. and Davaud, A. (1994). Photoinhibition of photosynthesis in chilled potato leaves is not correlated with a loss of photosystem-II activity. Preferential inactivation of photosystem I. Photosynthesis Research, 40, 75–92.CrossRefGoogle Scholar
Havaux, M. and Lannoye, R. (1983). Temperature dependence of delayed chlorophyll fluorescence in intact leaves of higher plants. A rapid method for detecting the phase transition of thylakoid membrane lipids. Photosynthesis Research, 4, 257–263.Google ScholarPubMed
Havaux, M. and Niyogi, K.K. (1999). The violaxanthin cycle protects plants from photooxidative damage by more than one mechanism. Proceedings of the National Academy of Sciences USA, 96, 8762–8767.CrossRefGoogle ScholarPubMed
Havaux, M., Dall’Osto, L. and Bassi, R. (2007). Zeaxanthin has enhanced antioxidant capacity with respect to all other xanthophylls in Arabidopsis leaves and functions independent of binding of PSII antennae. Plant Physiology, 145, 1506–1520.CrossRefGoogle ScholarPubMed
Havaux, M., Greppin, H. and Strasser, R.J. (1991). Functioning of photosytems I and II in pea leaves exposed to heat stress in the presence or absence of light. Planta, 186, 88–98.CrossRefGoogle Scholar
Havaux, M., Tardy, F., Ravenel, J., et al. (1996). Thylakoid membrane stability to heat stress studied by flash spectroscopic measurements of the electrochromic shift in intact potato leaves: influence of the xanthophyll content. Plant, Cell and Environment, 19, 1359–1368.CrossRefGoogle Scholar
Havaux, M., Triantaphylidès, C. and Genty, B. (2006). Autoluminiscence imaging: a non-invasive tool for mapping oxidative stress. Trends in Plant Science, 11, 480–484.CrossRefGoogle Scholar
Haveman, J. and Mathis, P. (1976). Flash-induced absorption changes of the primary donor of photosystem II at 820 nm in chloroplasts inhibited by low ph or tris-treatment. Biochimica et Biophysica Acta (BBA) – Bioenergetics, 440, 346–355.CrossRefGoogle ScholarPubMed
Haverkort, A.J., Rouse, D.I. and Turkensteen, L.J. (1990). The influence of Verticillium dahliae and drought on potato crop growth. 1. Effects on gas exchange and stomatal behaviour of individual leaves and crop canopies. Netherlands Journal of Plant Pathology, 96, 273–289.CrossRefGoogle Scholar
Haworth, P. and Melis, A. (1983). Phosphorylation of chloroplast membrane proteins does not increase the absorption cross-section of photosystem I. FEBS Letters, 160, 277–280.CrossRefGoogle Scholar
He, C., Yan, J., Shen, G., et al. (2005). Expression of an Arabidopsis vacuolar sodium/proton antiporter gene in cotton improves photosynthetic performance under salt conditions and increases fiber yield in the field. Plant and Cell Physiology, 46, 1848–1854.CrossRefGoogle ScholarPubMed
He, C.R., Murray, F. and Lyons, T. (2000). Monoterpene and isoprene emissions from 15 Eucalyptus species in Australia. Atmospheric Environment, 34, 645–655.CrossRefGoogle Scholar
He, J., Ouyang, W. and Chia, T.F. (2004). Growth and photosynthesis of virus-infected and virus-eradicated orchid plants exposed to different growth irradiances under natural tropical conditions. Physiologia Plantarum, 121, 612–619.CrossRefGoogle Scholar
He, J.Y., Ren, Y.F., Zhu, C., et al. (2008). Effect of Cd on growth, photosynthetic gas exchange, and chlorophyll fluorescence of wild and Cd-sensitive mutant rice. Photosynthetica, 46(3), 466–470.CrossRefGoogle Scholar
Heath, R.L. (1996). The modification of photosynthetic capacity induced by ozone exposure. In: Photosynthesis and the Environment (ed. Baker, N.R.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 469–476.Google Scholar
Heaton, E., Voigt, T. and Long, S.P. (2004). A quantitative review comparing the yields of two candidate C4 perennial biomass crops in relation to nitrogen, temperature and water. Biomass and Bioenergy, 27, 21–30.CrossRefGoogle Scholar
Heber, U. (1969). Conformational changes of chloroplasts induced by illumination of leaves in vivo. Biochimica et Biophysica Acta, 180, 302–319.CrossRefGoogle ScholarPubMed
Heber, U., Bilger, W., Bligny, R., et al. (2000). Phototolerance of lichens, mosses and higher plants in an alpine environment: analysis of photoreactions. Planta, 211, 770–780.CrossRefGoogle Scholar
Heber, U., Bligny, R., Streb, P., et al. (1996). Photorespiration is essential for the protection of the photosynthetic apparatus of C3 plants against photoinactivation under sunlight. Botanica Acta, 109, 307–315.CrossRefGoogle Scholar
Heber, U., Neimanis, S., Dietz, K.J., et al. (1986). Assimilation power as a driving force in photosynthesis. Biochimica et Biophysica Acta, 852, 144–155.CrossRefGoogle Scholar
Heckathorn, S.A., Downs, C.A., Sharkey, T.D., et al. (1998). The small, methionine-rich chloroplast heat-shock protein protects photosystem II electron transport during heat stress. Plant Physiology, 116, 439–444.CrossRefGoogle ScholarPubMed
Heckathorn, S.A., Ryan, S.L., Baylis, J.A., et al. (2002). In vivo evidence from an Agrostis stolonifera selection genotype that chloroplast small heat-shock proteins can protect photosystem II during heat stress. Functional Plant Biology, 29, 933–944.CrossRefGoogle Scholar
Heckwolf, M., Pater, D., Hanson, D.T. et al. (2011). The Arabidopsis thaliana aquaporin AtPIP1;2 is a physiologically relevant CO2 transport facilitator. The Plant Journal, 67, 795–804.CrossRefGoogle Scholar
Hedberg, O. (1964). Features of afroalpine plant ecology. Acta Phytogeographica Suecia, 49, 1–144.Google Scholar
Hedrich, R., Marten, I., Lohse, G., et al. (1994). Malate-sensitive anion channels enable guard cells to sense changes in the ambient CO2. The Plant Journal, 6, 741–748.CrossRefGoogle Scholar
Heidorn, T. and Joern, A. (1984). Differential herbivory on C3 and C4 grasses by the grasshopper Ageneotettix deorum (Orthoptera: Acrididae). Oecologia, 65, 19–25.CrossRefGoogle Scholar
Heilmann, I., Mekhedov, S., King, B., et al. (2004). Identification of the Arabidopsis palmitoyl-monogalactosyldiacylglycerol D7-desaturase gene FAD5, and effects of plastidial retargeting of Arabidopsis desaturases on the fad5 mutant phenotype. Plant Physiology, 136, 4237–4245.CrossRefGoogle ScholarPubMed
Heimann, M. and Reichstein, M. (2008). Terrestrial ecosystem carbon dynamics and climate feedbacks. Nature, 451, 289–292.CrossRefGoogle ScholarPubMed
Heisel, F., Sowinska, M., Khalili, E., et al. (1997). Laser-induced fluorescence imaging for monitoring nitrogen fertilising treatements of wheat. In: Aerosense ‘97 (eds Narayanan, R.M. and Kalshoven, J.E.), SPIE, Bellingham, USA, pp. 10–21.Google Scholar
Held, A.A., Steduto, P., Orgaz, F., et al. (1990). Bowen-ratio energy balance technique for estimating crop net CO2 assimilation and comparison with a canopy chamber. Theoretical and Applied Climatology, 42, 203–213.CrossRefGoogle Scholar
Helliker, B.R. and Ehleringer, J.R. (2000). Establishing a grassland signature in veins: 18O in the leaf water of C3 and C4 grasses. Proceedings of the National Academy of Science USA, 97, 7894–7898.CrossRefGoogle ScholarPubMed
Helliker, B.R. and Richter, S.L. (2008). Subtropical to boreal convergence of tree-leaf temperatures. Nature, 454, 511–514.CrossRefGoogle ScholarPubMed
Hellkvist, J., Richards, G.P. and Jarvis, P.G. (1974). Vertical gradients of water potential and tissue water relations in Sitka spruce trees measured with the pressure chamber. J Appl Ecol, 11, 637–667.CrossRefGoogle Scholar
Henderson, S.A., von Caemmerer, S. and Farquhar, G.D. (1992). Short-term measurements of carbon isotope discrimination in several C4 species. Australian Journal of Plant Physiology, 19, 263–285.CrossRefGoogle Scholar
Hendrey, G., Lewin, K., Kolber, Z., et al. (1989). Control of ozone concentrations for plant effect studies. Report to the National Council of the Paper Industry for Air and Stream Improvement, BNL 43589.
Hendrickson, L., Chow, W.S. and Furbank, R.T. (2004b). Low temperature effects on grapevine photosynthesis: the role of inorganic phosphate. Functional Plant Biology, 31, 789–801.CrossRefGoogle Scholar
Hendrickson, L., Furbank, R.T. and Chow, W.S. (2004a). A simple alternative approach to assessing the fate of absorbed light energy using chlorophyll fluorescence. Photosynth Research, 82, 73–81.CrossRefGoogle ScholarPubMed
Hendrickson, L., Sharwood, R., Ludwig, M., et al. (2008). The effects of Rubisco activase on C4 photosynthesis and metabolism at high temperature. Journal of Experimental Botany, 59, 1789–1798.CrossRefGoogle ScholarPubMed
Hendrickson, L., Vlckova, A., Selstam, E., et al. (2006). Cold acclimation of the Arabidopsis dgd1 mutant results in recovery from photosystem I-limited photosynthesis. FEBS Letters, 580, 4959–4968.CrossRefGoogle ScholarPubMed
Heng-Moss, T., Macedo, T., Franzen, L., et al. (2006). Physiological responses of resistant and susceptible buffalo grasses to Blissus occiduus (Hemiptera: Blissidae) feeding. Journal of Economic Entomology, 99, 222–228.CrossRefGoogle Scholar
Henriques, F.S. (1989). Effects of copper deficiency on the photosynthetic apparatus of sugar beet (Beta vulgaris L.). Journal of Plant Physiology, 135, 453–458.CrossRefGoogle Scholar
Henriques, F.S. (2001). Loss of blade photosynthetic area and of chloroplasts’ photochemical capacity account for reduced CO2 assimilation rates in zinc-deficient sugar beet leaves. Journal of Plant Physiology, 158, 915–919.CrossRefGoogle Scholar
Henriques, F.S. (2003). Gas exchange, chlorophyll a fluorescence kinetics and lipid peroxidation of pecan leaves with varying manganese concentrations. Plant Science, 165, 239–244.CrossRefGoogle Scholar
Henry, B.K., Atkin, O.K., Day, D.A., et al. (1999). Calculation of the isotope discrimination factor for studying plant respiration. Australian Journal of Plant Physiology, 26, 773–780.CrossRefGoogle Scholar
Herbert, S.K., Fork, D.C. and Malkin, S. (1990). Photoacoustic measurements in vivo of energy storage by cyclic electron flow in algae and higher plants. Plant Physiology, 94, 926–934.CrossRefGoogle ScholarPubMed
Herbette, S., Le Menn, A., Rousselle, P., et al. (2007). Modification of photosynthetic regulation in tomato overexpressing glutathione peroxidise. Biochimica et Biophysica Acta, 1724, 108–118.CrossRefGoogle Scholar
Herde, O., Peña-Cortés, H., Willmitzer, L., et al. (1997). Stomatal responses to jasmonic acid, linolenic acid and abscisic acid in wild-type and ABA-deficient tomato plants. Plant, Cell and Environment, 20, 136–141.CrossRefGoogle Scholar
Hernández, J.A. and Almansa, M.S. (2002). Short-term effects of salt stress on antioxidant systems and leaf water relations of pea leaves. Physiologia Plantarum, 115, 251–257.CrossRefGoogle ScholarPubMed
Herold, A. (1980). Regulation of photosynthesis by sink activity – the missing link. New Phytologist, 86, 131–144.CrossRefGoogle Scholar
Herrera, A., Tezara, W., Marín, O., et al. (2008). Stomatal and non-stomatal limitations of photosynthesis in trees of a tropical seasonally flooded forest. Physiologia Plantarum, 134, 41–48.CrossRefGoogle ScholarPubMed
Herrick, J.D. and Thomas, R.B. (2003). Leaf senescence and late-season net photosynthesis of sun and shade leaves of overstory sweetgum (Liquidambar styraciflua) grown in elevated and ambient carbon dioxide concentrations. Tree Physiology, 23, 109–118.CrossRefGoogle Scholar
Herring, J.R. (1985). Charcoal fluxes into sediments of the North Pacific Ocean: the Cenozoic record of burning. In: The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present (eds Sundquist, E.T. and Broecker, W.S.), American Geophysical Union, Washington DC, USA, pp. 419–442.Google Scholar
Herrmann, K.M. and Weaver, L.M. (1999). The shikimate pathway. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 473–503.CrossRefGoogle ScholarPubMed
Herzog, B., Grams, T.E.E., Haag-Kerwer, A., et al. (1999a). Expression of modes of photosynthesis (C3, CAM) in Clusia criuva CAMB. in a cerrado/gallery forest transect. Plant Biology, 1, 357–364.CrossRefGoogle Scholar
Herzog, B., Hoffmann, S., Hartung, W., et al. (1999b). Comparison of photosynthetic responses of the sympatric tropical C3-species Clusia multiflora H.B.K. and the C3-CAM intermediate species Clusia minor L. to irradiance and drought stress in a phytotron. Plant Biology, 1, 460–470.CrossRefGoogle Scholar
Heschel, M.S., Donohue, K., Hausmann, N., et al. (2002). Population differentiation and natural selection for water-use efficiency in Impatiens capensis (Balsaminaceae). International Journal of Plant Sciences, 163, 907–912.CrossRefGoogle Scholar
Hetherington, A.M. and Woodward, F.I. (2003). The role of stomata in sensing and driving environmental change. Nature, 424, 901–908.CrossRefGoogle ScholarPubMed
Hetherington, S.E., Smillie, R.M. and Davies, W.J. (1998). Photosynthetic activities of vegetative and fruiting tissues of tomato. Journal of Experimental Botany, 15, 1173–1181.CrossRefGoogle Scholar
Hew, C.S., Lee, G.L. and Wong, S.C. (1980). Occurrence of non-functional stomata in the flowers of tropical orchids. Annals of Botany, 46, 195–201.CrossRefGoogle Scholar
Hew, C.S., Ye, Q.S. and Pau, R.C. (1991). Relation of respiration to CO2 fixation by Aranda orchid roots. Environmental and Experimental Botany, 31, 327–331.CrossRefGoogle Scholar
Hibberd, J.M. and Covshoff, S. (2010). The regulation of gene expression required for C4 photosynthesis. Annual Review of Plant Biology, 61, 181–207.CrossRefGoogle ScholarPubMed
Hibberd, J.M. and Quick, W.P. (2002). Characteristics of C4 photosynthesis in stems and petioles of C3 flowering plants. Nature, 145, 451–454.CrossRefGoogle Scholar
Hibberd, J.M., Bungard, R.A., Press, M.C., et al. (1998a). Localization of photosynthetic metabolism in the parasitic angiosperm Cuscuta reflexa. Planta, 205, 506–513.CrossRefGoogle Scholar
Hibberd, J.M., Quick, W.P., Press, M.C., et al. (1998b). Can source-sink relations explain responses of tobacco to infection by the root holoparasitic angiosperm Orobanche cernua. Plant, Cell and Environment, 21, 333–340.CrossRefGoogle Scholar
Hideg, E., Kálai, T., Kós, P.B., et al. (2006). Singlet oxygen in plants: its significance and possible detection with double (fluorescent and spin) indicator reagents. Photochemistry and Photobiology, 82, 1211–1218.CrossRefGoogle ScholarPubMed
Hideg, É., Ogawa, K., Kálai, T., et al. (2001). Singlet oxygen imaging in Arabidopsis thaliana leaves under photoinhibition by excess photosynthetically active radiation. Physiologia Plantarum, 112, 10–14.CrossRefGoogle Scholar
Hietz, P., Wanek, W. and Dünisch, O. (2005). Long-term trends in cellulose d13C and water-use efficiency of tropical Cedrela and Swietenia from Brazil. Tree Physiology, 25, 745–752.CrossRefGoogle Scholar
Higuchi, R.G., Dollinger, P.S., Walsh, S., et al. (1992). Simultaneous amplification and detection of specific DNA sequences. Biotechnology, 10, 413–417.CrossRefGoogle ScholarPubMed
Hikosaka, K. (1997). Modelling optimal temperature acclimation of the photosynthetic apparatus in C3 plants with respect to nitrogen use. Annals of Botany, 80, 721–730.CrossRefGoogle Scholar
Hikosaka, K. (2003). A model of dynamics of leaves and nitrogen in a plant canopy: an integration of canopy photosynthesis, leaf lifespan, and nitrogen use efficiency. American Naturalist, 162, 149–164.CrossRefGoogle Scholar
Hikosaka, K. and Shigeno, A. (2009). The role of Rubisco and cell walls in the interspecific variation in photosynthetic capacity. Oecologia, 160, 443–451.CrossRefGoogle ScholarPubMed
Hikosaka, K. and Terashima, I. (1996). Nitrogen partitioning among photosynthetic components and its consequence in sun and shade plants. Functional Ecology, 10, 335–343.CrossRefGoogle Scholar
Hikosaka, K., Hanba, Y.T., Hirose, T., et al. (1998). Photosynthetic nitrogen-use efficiency in leaves of woody and herbaceous species. Functional Ecology, 12, 896–905.CrossRefGoogle Scholar
Hikosaka, K., Ishikawa, K., Borjigidai, A., et al. (2006). Temperature acclimation of photosynthesis: mechanisms involved in the changes in temperature dependence of photosynthetic rate. Journal of Experimental Botany, 57, 291–302.CrossRefGoogle ScholarPubMed
Hikosaka, K., Murakami, A. and Hirose, T. (1999). Balancing carboxylation and regeneration of ribulose-1,5-bisphosphate in leaf photosynthesis: temperature acclimation of an evergreen tree, Quercus myrsinaefolia. Plant Cell Environment, 22, 841–849.CrossRefGoogle Scholar
Hilker, T., Hall, F.G., Coops, N.C. et al. (2010). Remote sensing of photosynthetic light-use efficiency across two forested biomes: spatial scaling. Remote Sensing of the Environment, 114, 2863–2874.CrossRefGoogle Scholar
Hill, S.A. and Bryce, J.H. (1992). Malate metabolism and LEDR in barley mesophyll protoplasts. In: Molecular, Biochemical and Physiological Aspects of Plant Respiration. (eds Lambers, H. and Van der Plas, L.H.W.), SPB Academic Publishing, The Hague.
Hiraki, M., van Rensen, J.J.S., Vredenberg, W.J., et al. (2003). Characterization of the alterations of chlorophyll a fluorescence induction curve after addition of Photosystem II inhibiting herbicides. Photosynthesis Research, 78, 35–46.CrossRefGoogle ScholarPubMed
Hirashima, M., Satoh, S., Tanaka, R., et al. (2006). Pigment shuffling in antenna systems achieved by expressing prokaryotic chlorophyllide a oxygenase in Arabidopsis. The Journal of Biological Chemistry, 281, 15385–15393.CrossRefGoogle ScholarPubMed
Hiroki, S., and Ichino, K. (1998). Comparison of growth habits under various light conditions between two climax species, Castanopsis sieboldii and Castanopsis cuspidata, with special reference to their shade tolerance. Ecological Research, 13, 65–72.CrossRefGoogle Scholar
Hirose, T. and Werger, M.J.A. (1987). Maximizing daily canopy photosynthesis with respect to the leaf nitrogen allocation pattern in the canopy. Oecologia, 72, 520–526.CrossRefGoogle ScholarPubMed
Hirschberg, J. (2001). Carotenoid biosynthesis in flowering plants. Current Opinion of Plant Biology, 4, 210–218.CrossRefGoogle ScholarPubMed
Hjelm, U. and Ögren, E. (2003). Is photosynthetic acclimation to low temperature controlled by capacities for storage and growth at low temperature? Results from comparative studies of grasses and trees. Physioogial Plantarum, 119, 113–120.CrossRefGoogle Scholar
Hlavácková, V., Spundova, M., Naus, J., et al. (2002). Mechanical wounding caused by inoculation influences the photosynthetic response of Nicotiana benthamiana plants to plum pox potyvirus. Photosynthetica, 40, 269–277.CrossRefGoogle Scholar
Hoch, G. and Körner, C. (2003). The carbon charging of pines at the climatic treeline: a global comparison. Oecologia, 135, 10–21.CrossRefGoogle ScholarPubMed
Hoch, G., Popp, M. and Körner, C. (2002). Altitudinal increase of mobile carbon pools in Pinus cembra suggests sink limitation of growth at the Swiss treeline. Oikos, 98, 361–374.CrossRefGoogle Scholar
Hoch, G., Richter, A. and Körner, C. (2003). Non-structural carbon compounds in temperate forest trees. Plant Cell Environment, 26, 1067–1081.CrossRefGoogle Scholar
Hoch, W.A., Sinsaas, E.L. and McCown, B.H. (2003b). Resorption protection. Anthocyanins facilitate nutrient recovery in autumn by shielding leaves from potentially damaging light levels. Plant Physiology, 133, 1296–1305.CrossRefGoogle ScholarPubMed
Hoch, W.A., Zeldin, E.L. and McCown, B.H. (2001). Physiological significance of anthocyanins during autumnal leaf senescence. Tree Physiology, 21, 1–8.CrossRefGoogle ScholarPubMed
Hodges, D.M., DeLong, J.M., Forney, C.F., et al. (1999). Improving the thiobarbituric acid-reactive-substances assay for estimating lipid peroxidation in plant tissues containing anthocyanin and other interfering compounds. Planta, 207, 604–611.CrossRefGoogle Scholar
Hodges, M. (2002). Enzyme redundancy and the importance of 2-oxoglutarate in plant ammonium assimilation. Journal of Experimental Botany, 53, 905–916.CrossRefGoogle ScholarPubMed
Hodges, M. and Moya, I. (1986). Time-resolved chlorophyll fluorescence studies of photosynthetic membranes: resolution and characterisation of four kinetics components. Biochimica et Biophysica Acta, 849, 193–202.CrossRefGoogle Scholar
Hodgson, A.J.R., Beachy, N.R., Pakrasi, B.H. (1989). Selective inhibition of photosystem II spinach by Tobacco mosaic virus: an effect of the viral coat protein. FEBS Letters, 245, 267–270.CrossRefGoogle ScholarPubMed
Hoefnagel, H.N., Atkin, O.K. and Wiskich, J.T. (1998). Interdependence between chloroplasts and mitochondria in the light and the dark. Biochimica et Biophysica Acta, 1366, 235–255.CrossRefGoogle Scholar
Hoekstra, H.E. and Coyne, J.A. (2007). The locus of evolution: evo devo and the genetics of adaptation. Evolution, 61, 995–1016.CrossRefGoogle ScholarPubMed
Hoff, C. and Rambal, S. (2003). An examination of the interaction between clime, soil and leaf-area index in a Quercus ilex ecosystem. Annals of Forest Science, 60, 153–161.CrossRefGoogle Scholar
Hoffmann, W.A. (1999). Fire and population dynamics of woody plants in a neotropical savanna: matrix model projections. Ecology, 80, 1354–1369.CrossRefGoogle Scholar
Hoffmann, W.A., Franco, A.C., Moreira, M.Z., et al. (2005). Specific leaf area explains differences in leaf traits between congeneric savanna and forest trees. Functional Ecology, 19, 932–940.CrossRefGoogle Scholar
Hoffmann, W.A., Orthen, B. and Do Nascimento, P.K.V. (2003). Comparative fire ecology of tropical savanna and forest trees. Functional Ecology, 17, 720–726.CrossRefGoogle Scholar
Hoffmann-Benning, S., Willmitzer, L. and Fisahn, J. (1997). Analysis of growth, composition and thickness of the cell walls of transgenic tobacco plants expressing a yeast-derived invertase. Protoplasma, 200, 146–153.CrossRefGoogle Scholar
Högberg, P., Nordgren, A., Buchmann, N., et al. (2001). Large-scale forest girdling shows that current photosynthesis drives soil respiration. Nature, 411, 789–792.CrossRefGoogle ScholarPubMed
Hohmann-Marriott, M.F. and Blankenship, R.E. (2011). Evolution of photosynthesis. Annual Review of Plant Biology, 62, 515–548.CrossRefGoogle ScholarPubMed
Holbrook, N.M., Shashidhar, V.R., James, R.A., et al. (2002). Stomatal control in tomato with ABA-deficient roots: response of grafted plants to soil drying. Journal of Experimental Botany, 53, 1503–1514.Google ScholarPubMed
Holden, M. (1976). Chlorophylls. In: Chemistry and Biochemistry of Plant Pigments (ed. Goodwin, T.W.), Academic Press, New York, USA, pp. 1–37.Google Scholar
Holladay, A.S., Martindale, W., Alred, R., et al. (1992). Changes in activities of enzymes of carbon metabolism in leaves during exposure of plants to low temperature. Plant Physiology, 98, 1105–114.CrossRefGoogle Scholar
Hollinger, S.E., Bernacchi, C.J. and Meyers, T. (2005). Carbon budget of mature no-till ecosystem in North Central Region of the United States. Agricultural and Forest Meteorology, 130, 59–69.CrossRefGoogle Scholar
Hollinger, D.Y., Goltz, S.M., Davidson, E.A., et al. (1999). Seasonal patterns and environmental control of carbon dioxide and water vapour exchange in an ecotonal boreal forest. Global Change Biology, 5, 891–902.CrossRefGoogle Scholar
Holmes, M.G. and Keiller, D.R. (2002). Effects of pubescence and waxes on the reflectance of leaves in the ultraviolet and photosynthetic wavebands: a comparison of a range of species. Plant, Cell and Environment, 25, 85–93.CrossRefGoogle Scholar
Holt, N.E., Zigmantas, D., Valkunas, L., et al. (2005). Carotenoid cation formation and the regulation of photosynthetic light harvesting. Science, 307, 433–436.CrossRefGoogle ScholarPubMed
Holtgrefe, S., Gohlke, J., Starmann, J., et al. (2007). Regulation of plant cytosolic glyceraldehyde 3-phosphate dehydrogenase isoforms by thiol modifications. Physiologia Plantarum, 133, 211–218.CrossRefGoogle Scholar
Holthe, P.A. and Szarek, S.R. (1985). Physiological potential for survival of propagules of crassulacean acid metabolism species. Plant Physiology, 79, 219–224.CrossRefGoogle ScholarPubMed
Holtum, J.A.M. and Winter, K. (1999). Degrees of crassulacean acid metabolism in tropical epiphytic and lithophytic ferns. Australian Journal of Plant Physiology, 26, 749–757.CrossRefGoogle Scholar
Holtum, J.A.M., Aranda, J.Virgo, A., et al. (2004). δ13C values and crassulacean acid metabolism in Clusia species from Panama. Trees: Structure and Function, 18, 658–668.CrossRefGoogle Scholar
Holub, O., Seufferheld, M.J., Gohlke, C., et al. (2000). Fluorescence lifetime imaging (FLI) in real time: a new technique in photosynthesis research. Photosynthetica, 38, 581–599.CrossRefGoogle Scholar
Holzwarth, A.R., Wendler, J. and Haehnel, W. (1985). Time-resolved picosecond fluorescence spectra of the antenna chlorophylls in Chlorella vulgaris. Resolution of photosystem I fluorescence. Biochimica et Biophysica Acta, 807, 155–167.CrossRefGoogle Scholar
Homann, P.H. (1999). Reliability of photosystem II thermoluminescence measurements after sample freezing: few artifacts with photosystem II membranes but gross distortions with certain leaves. Photosynthesis Research, 62, 219–229.CrossRefGoogle Scholar
Hong, S.W., Lee, U. and Vierling, E. (2003). Arabidopsis hot mutants define multiple functions required for acclimation to high temperatures. Plant Physiology, 132, 757–767.CrossRefGoogle ScholarPubMed
Hoorn, C., Ohja, T. and Quade, J. (2000). Palynological evidence for vegetation development and climatic change in the sub-Himalayan Zone (Neogene, central Nepal). Palaeogeography, Palaeoclimatology, Palaeoecology, 163, 133–161.CrossRefGoogle Scholar
Hope, A.B., Valente, P. and Matthews, D.B. (1994). Effects of pH on the kinetics of redox reactions in and around the cytochrome bf complex in an isolated system. Photosynthesis Research, 42, 111–120.CrossRefGoogle Scholar
Hopkins, D.L. (1989). Xylella fastidiosa: xylem-limited bacterial pathogen of plants. Annual Review of Phytopathology, 27, 271–290.CrossRefGoogle Scholar
Hopley, P.J., Marshall, J.D., Weedon, G.P., et al. (2007). Orbital forcing and the spread of C4 grasses in the late Neogene: stable isotope evidence from South African speleotherms. Journal of Human Evolution, 53, 620–634.CrossRefGoogle Scholar
Horken, K.M. and Tabita, F.R. (1999). Closely related form I ribulose bisphosphate carboxylase/oxygenase molecules that possess different CO2/O2 substrate specificities. Archives of Biochemistry and Biophysics, 361, 183–194.CrossRefGoogle Scholar
Horler, D.N.H., Barber, J. and Barringer, A.R. (1980). Effects of heavy metals on the absorbance and reflectance spectra of plants. International Journal of Remote Sensing, 1, 121–136.CrossRefGoogle Scholar
Hormaetxe, K., Becerril, J.M., Fleck, I., et al. (2005a). Functional role of red (retro)-carotenoids as passive light filters in the leaves of Buxus sempervirens L.: increased protection of photosynthetic tissues?Journal of Experimental Botany, 56, 2629–2636.CrossRefGoogle ScholarPubMed
Hormaetxe, K., Esteban, R., Becerril, J.M., et al. (2005b), Dynamics of the α-tocopherol pool as affected by external (environmental) and internal (leaf age) factors in Buxus sempervirens leaves. Physiologia Plantarum, 125, 333–344.CrossRefGoogle Scholar
Horn, H.S. (1971). The Adaptive Geometry of Trees. Princeton University Press, Princeton, New Jersey.Google Scholar
Hörtensteiner, S. (2006). Chlorophyll degradation during senescence. Annual Review of Plant Biology, 57, 55–77.CrossRefGoogle ScholarPubMed
Hörtensteiner, S. and Feller, U. (2002). Nitrogen metabolism and remobilization during senescence. Journal Experimental Botany, 53, 927–937.CrossRefGoogle ScholarPubMed
Horton, J.L., Kolb, T. and Hart, S.C. (2001). Leaf gas exchange characteristics differ among Sonoran Desert riparian tree species. Tree Physiology, 21, 233–241.CrossRefGoogle ScholarPubMed
Horton, P. (2000). Prospects for crop improvement through the genetic manipulation of photosynthesis: morphological and biochemical aspects of light capture. J. Experimental Botany, 51, 475–485.CrossRefGoogle ScholarPubMed
Horton, P., Mathew, P.J., Perez-Bueno, M.L., et al. (2008). Photosynthetic acclimation: does the dynamic structure and macro-organisation of photosystem II in higher plant grana membranes regulate light harvesting states?FEBS Journal, 275, 1069–1079.CrossRefGoogle ScholarPubMed
Horton, P., Ruban, A.V. and Young, A.J. (1999). Regulation of the structure and function of the light harvesting complexes of photosystem II by the xanthophyll cycle. In: Advances in Photosynthesis and Respiration: The Photochemistry of Carotenoids (eds Frank, H.A., Young, A.J., Britton, G. and Cogdell, R.J.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 271–291.Google Scholar
Horton, P., Wentworth, M. and Ruban, A. (2005). Control of the light harvesting function of chloroplast membranes: the LHCII-aggregation model for non-photochemical quenching. FEBS Letters, 579, 4201–4206.CrossRefGoogle ScholarPubMed
Horváth, E.M., Peter, S.O., Joët, T., et al. (2000). Targeted inactivation of the plastid ndhB gene in tobacco results in an enhanced sensitivity of photosynthesis to moderate stomatal closure. Plant Physiology, 123, 1337–1349.CrossRefGoogle Scholar
Hosler, J.P. and Yocum, C.F. (1987). Regulation of cyclic photophosphorylation during ferredoxin-mediated electron transport: effect of DCMU and the NADPH/NADP ratio. Plant Physiology, 83, 965–969.CrossRefGoogle ScholarPubMed
Housman, D.C., Naumburg, E., Huxman, T.E., et al. (2006). Increases in desert shrub productivity under elevated carbon dioxide vary with water availability. Ecosystems, 9, 374–385.CrossRefGoogle Scholar
Howe, G.A. and Jander, G. (2008). Plant immunity to insect herbivores. Annual Review of Plant Biology, 59, 41–66.CrossRefGoogle ScholarPubMed
Howe, G.T., Aitken, S.N., Neale, D.B., et al. (2003). From genotype to phenotype: unraveling the complexities of cold adaptation in forest trees. Canadian Journal of Botany, 81, 1247–1266.CrossRefGoogle Scholar
Hsiao, T.C., Steduto, P. and Fereres, E. (2007). A systematic and quantitative approach to improve water use efficiency in agriculture. Irrigation Sciences, 25, 209–231.CrossRefGoogle Scholar
Hsieh, T.H., Lee, J.T., Charng, Y.Y., et al. (2002). Tomato plants ectopically expressing Arabidopsis CBF1 show enhanced resistance to water deficit stress. Plant Physiology, 130, 618–626.CrossRefGoogle ScholarPubMed
Hu, H., Boisson-Dernier, A., Israelsson-Nordström, M., et al. (2010). Carbonic anhydrases are upstream regulators of CO2-controlled stomatal movements in guard cells. Nature Cell Biology, 12, 87–93.CrossRefGoogle ScholarPubMed
Hu, H., Dai, M., Yao, J., et al. (2006). Overexpressing a NAM, ATAF, and CUC (NAC) transcription factor enhances drought resistance and salt tolerance in rice. Proceedings of the National Academy of Sciences of USA, 103, 12987–12992.CrossRefGoogle Scholar
Hu, L., Wang, Z. and Huang, B. (2010). Diffusion limitations and metabolic factors associated with inhibition and recovery of photosynthesis from drought stress in a C3 perennial grass species. Physiologia Plantarum, 139, 93–106.CrossRefGoogle Scholar
Hu, X.L., Jiang, M.Y., Zhang, J.H., et al. (2007). Calcium-calmodulin is required for abscisic acid-induced antioxidant defense and functions both upstream and downstream of H2O2 production in leaves of maize (Zea mays) plants. New Phytologist, 173, 27–38.CrossRefGoogle ScholarPubMed
Huang, C.Y., Bazzaz, F.A. and Vanderhoef, L.N. (1974). The inhibition of soybean metabolism by cadmium and lead. Plant Physiology, 54, 122–124.CrossRefGoogle Scholar
Huang, H., Ger, M., Chen, C., et al. (2007). Disease resistance to bacterial pathogens affected by the amount of ferredoxin-I protein in plants. MolecularPlant Pathology, 8, 129–137.Google Scholar
Hubbard, R.M., Ryan, M.G., Stiller, V., et al. (2001). Stomatal conductance and photosynthesis vary linearly with plant hydraulic conductance in ponderosa pine. Plant, Cell and Environment, 24, 113–121.CrossRefGoogle Scholar
Hubick, K.T., Shorter, R. and Farquhar, G.D. (1988). Heritability and genotype × environment interactions of carbon isotope discrimination and transpiration efficiency in peanut (Arachis hypogaea L.). Australian Journal of Plant Physiology, 15, 799–813.CrossRefGoogle Scholar
Huc, R., Ferhi, A. and Guel, J.M. (1994). Pioneer and late stage tropical rainforest tree species (French Guyana) growing under common conditions differ in leaf gas exchange regulation, carbon isotope discrimination and leaf water potential. Oecologia, 99, 297–305.CrossRefGoogle Scholar
Huerta, L., Forment, J., Gadea, J., et al. (2008). Gene expression analysis in citrus reveals the role of gibberellins on photosynthesis and stress. Plant, Cell and Environment, 31, 1620–1633.CrossRefGoogle ScholarPubMed
Hughes, N.M., Vogelmann, T.C. and Smith, W.K. (2008). Optical effects of abaxial anthocyanin on absorption of red wavelengths by understory species: revisiting the back-scatter hypothesis. Journal of Experimental Botany, 59, 3435–3442.CrossRefGoogle Scholar
Hui, D.Q., Iqbal, J., Lehmann, K., et al. (2003). Molecular interactions between the specialist herbivore Manduca sexta (Lepidoptera, Sphingidae) and its natural host Nicotiana attenuata. V. Microarray analysis and further characterization of large-scale changes in herbivore-induced mRNAs. Plant Physiology, 131, 1877–1893.CrossRefGoogle ScholarPubMed
Hulbert, L.C. (1988). Causes of fire effects in tallgrass prairie. Ecology, 69, 46–58.CrossRefGoogle Scholar
Hull, R. (2002). Matthew’s Plant Virology, 4th edn. San Diego. California: Elsevier Academic Press.Google Scholar
Humbeck, K., Quast, S. and Krupinska, K. (1996). Functional and molecular changes in the photosynthetic apparatus during senescence of flag leaves from field-grown barley plants. Plant Cell Environment, 19, 337–344.CrossRefGoogle Scholar
Hummel, I., Pantin, F., Sulice, R. et al. (2010). Arabidopsis plants acclimate to water deficit at low cost through changes of carbon usage: an integrated perspective using growth, metabolite, enzyme, and gene expression analysis. Plant Physiology, 154, 357–372.CrossRefGoogle Scholar
Huner, N.P.A. and Ivanov, A.G. (2006). Photoprotection of photosystem II: Reaction center quenching versus antenna quenching. In: Advances in Photosynthesis and Respiration – Photoprotection, Photoinhibition, Gene Regulation, and Environment (eds Demmig-Adams, B., Adams, III, W.W. and Mattoo, A.), Springer, Dordrecht, Netherlands, pp. 155–173.Google Scholar
Huner, N.P.A., Öquist, G. and Melis, A. (2003). Photostasis in plants, green algae and cyanobacteria: the role of light harvesting antenna complexes. In: Advances in Photosynthesis and Respiration Light Harvesting Antennas in Photosynthesis (eds Green, B.R. and Parson, W.W.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 401–421.Google Scholar
Huner, N.P.A., Öquist, G. and Sarhan, F. (1998). Energy balance and acclimation to light and cold. Trends in Plant Science, 3, 224–230.CrossRefGoogle Scholar
Huner, N.P.A., Öquist, G., Hurry, V.M., et al. (1993). Photosynthesis, photoinhibition and low temperature acclimation in cold tolerant plants. Photosynthesis Research, 37, 19–39.CrossRefGoogle ScholarPubMed
Hungate, B.A., Reichstein, M., Dijkstra, P., et al. (2002). Evapotranspiration and soil water content in a scrub-oak woodland under carbon dioxide enrichment. Global Change Biology, 8, 289–298.CrossRefGoogle Scholar
Huppe, H. C. and Turpin, D. H. (1994). Integration of carbon and nitrogen metabolism in plant and algal cells. Annual Review of Plant Physiology and Plant Molecular Biology, 45, 577–607.CrossRefGoogle Scholar
Hura, T., Grzesiak, S., Hura, K., et al. (2006). Differences in the physiological state between triticale and maize plants during drought stress and followed rehydration expressed by the leaf gas exchange and spectrofluorimetric methods. Acta Physiologiae Plantarum, 28, 433–443.CrossRefGoogle Scholar
Hurry, V., Igamberdiev, A.U., Keergerg, O., et al. (2005). Respiration in photosynthetic cells: gas exchange components, interactions with photorespiration and the operation of mitochondria in the light. In: Plant Respiration: From Cell to Ecosystem (eds Lambers, H. and Ribas-Carbó, M.), Springer, Doordrecht, Netherlands, pp. 43–61.Google Scholar
Hurry, V.M., Malmberg, G., Garderstrom, P., et al. (1994). Effects of a short-term shift to low-temperature and of long-term cold hardening on photosynthesis and ribulose-1,5-bisphosphate carboxylase oxygenase and sucrose-phosphate synthase activity in leaves of winter rye (Secale cereale L). Plant Physiology, 106, 983–999.CrossRefGoogle Scholar
Hurry, V., Strand, A., Furbank, R., et al. (2000). The role of inorganic phosphate in the development of freezing tolerance and the acclimation of photosynthesis to low temperature is revealed by the pho mutants of Arabidopsis thaliana. Plant Journal, 24, 383–396.CrossRefGoogle ScholarPubMed
Hutchinson, B.A. and Matt, D.R. (1977). The distribution of solar radiation within a deciduous forest. Ecological Monographs, 47, 185–207.CrossRefGoogle Scholar
Hutchison, B.A., Matt, D.R. and McMillen, R.T. (1980). Effects of sky brightness distribution upon penetration of diffuse radiation through canopy gaps in a deciduous forest. Agricultural Meteorology, 22, 137–147.CrossRefGoogle Scholar
Hutchison, R.S., Groom, Q. and Ort, D.R. (2000). Differential effects of chilling-induced photooxidation on the redox regulation of photosynthetic enzymes. Biochemistry, 39, 6679–6688.CrossRefGoogle ScholarPubMed
Hutyra, L.R., Munger, J.W., Saleska, S.R., et al. (2007). Seasonal controls on the exchange of carbon and water in an Amazonian rain forest. Journal of Geophysical Research-Biogeosciences, 112.Google Scholar
Hüve, K., Bichele, I., Rasulov, B. et al. (2011). When it is too hot for photosynthesis: heat-induced instability of photosynthesis in relation to respiratory burst, cell permeability changes and H2O2 formation. Plant, Cell and Environment, 34, 113–126.CrossRefGoogle Scholar
Hüve, K., Bichele, I., Tobias, M., et al. (2006). Heat sensitivity of photosynthetic electron transport varies during the day due to changes in sugars and osmotic potential. Plant, Cell and Environment, 29, 212–228.CrossRefGoogle ScholarPubMed
Hüve, K., Christ, M.M., Kleist, E., et al. (2007). Simultaneous growth and emission measurements demonstrate an interactive control of methanol release by leaf expansion and stomata. Journal of Experimental Botany, 58, 1783–1793.CrossRefGoogle ScholarPubMed
Huxman, T.E., Barron-Gafford, G., Gerst, K.L., et al. (2008). Photosynthetic resource-use efficiency and demographic variability in desert winter annual plants. Ecology, 89, 1554–1563.CrossRefGoogle ScholarPubMed
Huxman, T.E., Hamerlynck, E.P., Moore, B.D., et al. (1998). Photosynthetic down-regulation in Larrea tridentata exposed to elevated atmospheric CO2: interaction with drought under glasshouse and field (FACE) exposure. Plant, Cell and Environment, 21, 1153–1161.CrossRefGoogle Scholar
Hwang, Y.H. and Morris, J.T. (1994). Whole-plant gas exchange responses of Spartina alterniflora (Poaceae) to a range of constant and transient salinities. American Journal of Botany, 81, 659–665.CrossRefGoogle Scholar
Ibrom, A., Dellwik, E., Flyvbjerg, H., et al. (2007). Strong low-pass filtering effects on water vapour flux measurements with closed-path eddy correlation systems. Agricultural and Forest Meteorology, 147, 140–156.CrossRefGoogle Scholar
ICNIRP (1996). International commission on non-ionizing radiation protection. Guidelines on limits of exposure to laser radiation of wavelengths between 180 nm and 1,000 µm. Health Physics, 71, 804–819.Google Scholar
Ida, K., Masamoto, K., Maoka, T., et al. (1995). The leaves of the common box, Buxus sempervirens (Buxaceae), become red as the level of a red carotenoid, anhydroeschscholtzxanthin, increases. Journal of Plant Research, 108, 369–376.CrossRefGoogle Scholar
Idle, D.B. and Proctor, C.W. (1983). An integrating sphere leaf chamber. Plant, Cell and Environment, 6, 437–439.CrossRefGoogle Scholar
Igamberdiev, A.U. and Gardeström, P. (2003). Regulation of NAD- and NADP-dependent isocitrate dehydrogenases by reduction levels of pyridine nucleotides in mitochondria and cytosol of pea leaves. Archives of Biochemistry and Biophysics, 1606, 117–125.Google ScholarPubMed
Igamberdiev, A.U. and Lea, P.J. (2006). Land plants equilibrate O2 and CO2 concentrations in the atmosphere. Photosynthesis Research, 87, 177–194.CrossRefGoogle ScholarPubMed
Igamberdiev, A.U., Mikkelsen, T.N., Ambus, P., et al. (2004). Photorespiration contributes to stomatal regulation and carbon isotope fractionation: a study with barley, potato and Arabidopsis plants deficient in glycine decarboxylase. Photosynthesis Research, 81, 139–152.CrossRefGoogle Scholar
Iida, S., Miyagi, A., Aoki, S., et al. (2009). Molecular adaptation of rbcL in the heterophyllous aquatic plant Potamogeton. PLoS ONE, 4, e4633.CrossRefGoogle ScholarPubMed
Inoue, Y. and Shibata, K. (1974). Comparative examination of terrestrial plant leaves in terms of light-induced absorption changes due to chloroplast rearrangements. Plant and Cell Physiology, 15, 717–721.CrossRefGoogle Scholar
IPCC (2007). Fourth Assessment Report of the Intergovernmental Panel on Climate Change (eds Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M. and Miller, H.L.), Cambridge University Press, Cambridge, UK and New York, USA, p. 976.Google Scholar
Ishida, H., Makino, A. and Mae, T. (1999). Fragmentation of the large subunit of ribulose-1,5-bisphosphate carboxylase by reactive oxygen species occurs near Gly 329. Journal of Biological Chemistry, 274, 5222–5226.CrossRefGoogle ScholarPubMed
Ishihara, K., Nishihara, T. and Ogura, T. (1971). The relationship between environmental factors and behaviour of stomata in the rice plant. I. On the measurement of the stomatal aperture. Proceedings of the Japanese Society of Crop Science, 40, 491–496.CrossRefGoogle Scholar
Ish-Shalom-Gordon, N., Lin, G. and Sternberg, L.D.S.L. (1992). Isotopic fractionation during cellulose synthesis in two mangrove species: salinity effects. Phytochemistry, 31, 2623–2626.CrossRefGoogle Scholar
Ismail, A.M. and Hall, A.E. (1993). Inheritance of carbon isotope discrimination and water-use efficiency in cowpea. Crop Science, 33, 498–503.CrossRefGoogle Scholar
Israel, A.A. and Nobel, P.S. (1995). Growth temperature versus CO2 uptake, Rubisco and PEPCase activities, and enzyme high-temperature sensitivities for a CAM plant. Plant Physiology and Biochemistry, 33, 345–351.Google Scholar
Iturbe-Ormaetxe, I., Morán, J. F., Arrese-Igor, C., et al. (1995). Activated oxygen and antioxidant defences in iron-deficient pea plants. Plant, Cell and Environment, 18, 421–429.CrossRefGoogle Scholar
Ivanov, A.G., Morgan, R.M., Gray, R.G., et al. (1998). Temperature/light dependent development of selective resistance to photoinhibition of photosystem I. FEBS Letters, 430, 288–292.CrossRefGoogle ScholarPubMed
Ivanov, A.G., Sane, P.V., Zeinalov, Y., et al. (2001). Photosynthetic electron transport adjutments in overwintering Scots pine (Pinus sylvestris L). Planta, 213, 575–585.CrossRefGoogle Scholar
Ivanov, A.G., Sane, P.V., Zeinalov, Y., et al. (2002). Seasonal responses of photosynthetic electron transport in Scots pine (Pinus sylvestris L.) studies by thermoluminescence. Planta, 215, 457–465.Google Scholar
Jack, S.B. and Long, J.N. (1992). Forest production and the organization of foliage within crowns and canopies. Forest Ecology and Management, 49, 233–245.CrossRefGoogle Scholar
Jackson, J.E. and Palmer, J.W. (1979). A simple model of light transmission and interception by discontinuous canopies. Annals of Botany, 44, 381–383.CrossRefGoogle Scholar
Jackson, M.B. (2002). Long-distance signalling from roots to shoots assessed: the flooding story. Journal of Experimental Botany, 53, 175–181.CrossRefGoogle ScholarPubMed
Jackson, M.B. and Hall, K.C. (1987). Early stomatal closure in water-logged pea plants is mediated by abscisic acid in the absence of foliar water deficits. Plant, Cell and Environment, 10, 121–130.Google Scholar
Jackson, P.C., Meinzer, F.C., Bustamante, M., et al. (1999). Partitioning of soil water among tree species in a Brazilian cerrado ecosystem. Tree Physiology, 19, 717–724.CrossRefGoogle Scholar
Jacobs, B.F. (2004). Palaeobotanical studies from tropical Africa: relevance to the evolution of forest, woodland and savannah biomes. Philosophical Transations of the Royal Society of London, Series B, 359, 1573–1583.CrossRefGoogle ScholarPubMed
Jacoby, G.C. and D’Arrigo, R.D. (1997). Tree rings, carbon dioxide, and climatic change. Proceedings of the National Academy of Sciences of the United States of America, 94, 8350–8353.CrossRefGoogle ScholarPubMed
Jacquemoud, S. and Baret, F. (1990). PROSPECT: a model of leaf optical properties spectra. Remote Sensing of Environment, 34, 75–91.CrossRefGoogle Scholar
Jahnke, S. (2001). Atmospheric CO2 concentration does not directly affect leaf respiration in bean or poplar. Plant, Cell and Environment, 24, 1139–1151.CrossRefGoogle Scholar
Jahnke, S. and Krewitt, M. (2002). Atmospheric CO2 concentration may directly affect leaf respiration measurement in tobacco, but not respiration itself. Plant, Cell and Environment, 25, 641–651.CrossRefGoogle Scholar
Jahnke, S. and Pieruschka, R. (2006). Air-pressure in clamp-on leaf chambers: a neglected issue in gas exchange measurements. Journal of Experimental Botany, 57, 2553–2561.CrossRefGoogle ScholarPubMed
Jalink, H., Frandas, A., van der Schoor, R., et al. (1998). Chlorophyll fluorescence of the testa of Brassica oleracea seeds as an indicator of maturity and seed quality. Scientia Agricola, 55, 88–93.CrossRefGoogle Scholar
James, S.A. and Bell, D. (2000). Leaf orientation, light interception and stomatal conductance of Eucalyptus globulus ssp. globulus leaves., Tree Physiology, 20, 815–823.CrossRefGoogle ScholarPubMed
Janda, T., Szalai, G., Papp, N., et al. (2004). Effect of freezing on thermoluminescence in various plant species. Photochemistry and Photobiology, 80, 525–530.2.0.CO;2>CrossRefGoogle ScholarPubMed
Janda, T., Szalai, G., Rios-Gonzalez, K., et al. (2003). Comparative study of frost tolerance and antioxidant activity in cereals. Plant Science, 164, 301–306.CrossRefGoogle Scholar
Janis, C.M., Damuth, J. and Theodor, J.M. (2000). Miocene ungulates and terrestrial primary productivity: where have all the browsers gone?Proceedings of the National Academy of Sciences, 97, 7899–7904.CrossRefGoogle ScholarPubMed
Jarvis, P.G. (1976). The interpretation of the variations in leaf water potential and stomatal conductance found in canopies in the field. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 273, 593–610.CrossRefGoogle Scholar
Jarvis, P.G. (1995). Scaling processes and problems. Plant, Cell and Environment, 18, 1079–1089.CrossRefGoogle Scholar
Jarvis, P.G. and Leverenz, J.W. (1983). Productivity of temperate, deciduous and evergreen forests. In: Physiological Plant Ecology, vol IV (eds Lange, O.L., Nobel, P.S., Osmond, C.B. and Ziegler, H.), Springer-Verlag, Berlin, Germany, pp. 233–280.Google Scholar
Jarvis, P.G. and McNaughton, K.G. (1986). Stomatal control of transpiration: scaling up from leaf to region. Advances in Ecological Research, 15, 1–49.CrossRefGoogle Scholar
Jarvis, P.G. and Sandford, A.P. (1986). Temperate forests. In: Photosynthesis in Contrasting Environments (eds Baker, N.R. and Long, S.P.), Elsevier Science Publishers B.V., New York, USA, pp. 199–236.Google Scholar
Jarvis, P.G., James, G.B. and Landsberg, J.J. (1976). Coniferous forests. In: Vegetation and the Atmosphere, vol 2 (ed. Monteith, J.L.). Academic Press, New York, USA, pp. 171–240.Google Scholar
Jasoni, R.L, Smith, S.D. and Arnone, J.A. (2005). Net ecosystem CO2 exchange in Mojave Desert shrublands during the eighth year of exposure to elevated CO2. Global Change Biology, 11, 749–756.CrossRefGoogle Scholar
Jayasekera, R. and Schleser, G.H. (1991). Seasonal changes in organic carbon content of leaves of deciduous trees. Journal of Plant Physiology, 138, 507–510.CrossRefGoogle Scholar
Jenkins, C.L.D., Furbank, R.T. and Hatch, M.D. (1989). Mechanisms of C4 photosynthesis. A model describing the inorganic carbon pool in bundle sheath cells. Plant Physiology, 91, 1372–1381.CrossRefGoogle Scholar
Jenny, H. (1941). Factors of Soil Formation. McGraw Hill, New York, USA.CrossRef
Jensen, P.E., Bassi, R., Boekema, E.J., et al. (2007). Structure, function and regulation of plant photosystem I. Biochimica et Biophysica Acta, 1767, 335–352.CrossRefGoogle ScholarPubMed
Jensen, P.E., Haldrup, A., Zhang, S., et al. (2004). The PSI-O subunit of plant photosystem I is involved in balancing the excitation pressure between the two photosystems. Journal of Biological Chemistry, 279, 24212–24217.CrossRefGoogle ScholarPubMed
Jensen, S.G. (1972). Metabolism and carbohydrate composition in barley yellow dwarf virus infected wheat. Phytopathology, 62, 587–592.CrossRefGoogle Scholar
Jeong, W.J., Park, Y.I., Suh, K., et al. (2002). A large population of small chloroplasts in tobacco leaf cells allows more effective chloroplast movement than a few enlarged chloroplasts. Plant Physiology, 129, 112–121.CrossRefGoogle ScholarPubMed
Jeschke, W.D. and Hilpert, A. (1997). Sink-stimulated photosynthesis and sink-dependent increase in nitrate uptake: nitrogen and carbon relations of the parasitic association Cuscuta reflexa-Ricinus communis. Plant, Cell and Environment, 20, 47–56.CrossRefGoogle Scholar
Jeschke, W.D., Baig, A. and Hilpert, A. (1997). Sink-stimulated photosynthesis, increased transpiration and increased demand-dependent stimulation of nitrate uptake: nitrogen and carbon relations in the parasitic association Cuscuta reflexa-Coleus blumei. Journal of Experimental Botany, 48, 915–925.CrossRefGoogle Scholar
Jia, G., Peng, P., Zhao, Q., et al.(2003). Changes in terrestrial ecosystem since 30 Ma in East Asia: stable isotope evidence from black carbon in the South China Sea. Geology, 31, 1093–1096.CrossRefGoogle Scholar
Jia, H., Oguchi, R., Hope, A.B., et al. (2008). Differential effects of severe water stress on linear and cyclic electron fluxes through photosystem I in spinach leaf discs in CO2-enriched air. Planta, 228, 803–812.CrossRefGoogle Scholar
Jiang, C.D., Gao, H.Y. and Zou, Q. (2001). Enhanced thermal energy dissipation depending on xanthophyll cycle and D1 protein turnover in iron-deficient maize leaves under high irradiance. Photosynthetica, 39, 269–274.CrossRefGoogle Scholar
Jiang, C.Z., Rodermel, S.R. and Shibles, R.M. (1997). Regulation of photosynthesis in developing leaves of soybean chlorophyll-deficient mutants. Photosynthesis Research, 51, 185–192.CrossRefGoogle Scholar
Jiang, G., Tang, H., Yu, M., et al. (1999). Response of photosynthesis of different plant functional types to environmental changes along Northeast China transect. Trees, 14, 72–82.CrossRefGoogle Scholar
Jiang, H.X., Chen, L.S., Zheng, J.G., et al. (2008). Aluminum-induced effects on photosystem II photochemistry in Citrus leaves assessed by the chlorophyll a fluorescence transient. Tree Physiology, 28, 1863–1871.CrossRefGoogle ScholarPubMed
Jiao, D., Hang, X., Li, X., et al. (2002). Photosynthetic characteristics and tolerance to photo-oxidation of transgenic rice expressing C4 photosynthesis enzymes. Photosynthesis Research, 72, 85–93.CrossRefGoogle Scholar
Jiao, J., Goodwin, P. and Grodzinski, B. (1999). Inhibition of photosynthesis and export in geranium grown at two CO2 levels and infected with Xanthomonas campestris pv. Pelargonii. Plant, Cell and Environment, 22, 15–25.CrossRefGoogle Scholar
Jiménez, I., Lopez, L., Alamillo, J.M., et al. (2006). Identification of a plum pox virus CI-interacting protein from chloroplast that has a negative effect in virus infection. Molecular Plant-Microbe Interactions, 19, 350–358.CrossRefGoogle Scholar
Job, C., Raijou, L., Lovigny, Y., et al. (2005). Patterns of protein oxidation in Arabidopsis seeds and during germination. Plant Physiology, 138, 790–802.CrossRefGoogle ScholarPubMed
Joel, G., Chapin, F.S. and Chiariello, N.R. (2001). Species-specific responses of plant communities to altered carbon and nutrient availability. Global Change Biology, 7, 435–450.CrossRefGoogle Scholar
Joët, T., Cournac, L., Horvath, E. M, et al. (2001). Increased sensitivity of photosynthesis to antimycin A induced by inactivation of the chloroplast ndhB gene. Evidence for a participation of the NADH-dehydrogenase complex to cyclic electron flow around photosystem I. Plant Physiology, 125, 1919–1929.CrossRefGoogle ScholarPubMed
Joët, T., Cournac, L, Peltier, G., et al. (2002). Cyclic electron flow around photosystem I in C-3 plants. In vivo control by the redox state of chloroplasts and involvement of the NADH-dehydrogenase complex. Plant Physiology, 128, 760–769.CrossRefGoogle ScholarPubMed
Johansson, E., Olsson, O. and Nystrom, T. (2004). Progression and specificity of protein oxidation in the lifecycle of Arabidopsis thaliana. Journal of Biological Chemistry, 279, 22204–22208.CrossRefGoogle Scholar
Johansson, J., Andersson, M., Edner, H., et al. (1996). Remote fluorescence measurements of vegetation spectrally resolved and by multi-colour fluorescence imaging. Journal of Plant Physiology, 148, 632–637.CrossRefGoogle Scholar
Johnson, D.A., Richards, R.A. and Turner, N.C. (1983). Yield, water relations, gas exchange, and surface reflectance of near-isogenic wheat lines differing in glaucousness. Crop Science, 24, 1168–1173.Google Scholar
Johnson, F.H., Eyring, H. and Williams, R.W. (1942). The nature of enzyme inhibitions in bacterial luminescence: sulfanilamide, urethane, temperature and pressure. Journal of Cellular and Comparative Physiology, 20, 247–268.CrossRefGoogle Scholar
Johnson, S.C. and Brown, W.V. (1973). Grass leaf ultrastructural variations. American Journal of Botany, 60, 727–735.CrossRefGoogle Scholar
Joiner, J., Yoshida, Y., Vasilkov, A., et al. (2010), First observations of global and seasonal terrestrial chlorophyll fluorescence from space, Biogeosci. Discuss., 7(6), 8281–8318.
Joliot, P., Lavergne, J. and Beal, D. (1992). Plastoquinone compartmentation in chloroplasts. I: evidence for domains with different rates of photo-reduction. Biochimica et Biophysica Acta (Bioenergetics), 1101, 1–12.CrossRefGoogle Scholar
Jones, A.M.E., Thomas, V., Bennett, M.H., et al. (2006). Modifications to the Arabidopsis defense proteome occur prior to significant transcriptional change in response to inoculation with Pseudomonas syringae. Plant Physiology, 142, 1603–1620.CrossRefGoogle ScholarPubMed
Jones, H.G. (1973). Moderate-term water stresses and associated changes in some photosynthetic parameters in cotton. The New Phytologist, 72, 1095–1105.CrossRefGoogle Scholar
Jones, H.G. (1985). Partitioning stomatal and non-stomatal limitations to photosynthesis. Plant, Cell and Environment, 8, 95–104.CrossRefGoogle Scholar
Jones, H.G. (1992). Plants and Microclimate. Cambridge University Press, New York, USA, pp. 323.
Jones, H.G. (1998). Stomatal control of photosynthesis and transpiration. Journal of Experimental Botany, 49, 387–398.CrossRefGoogle Scholar
Jones, H.G. (1999). Use of thermography for quantitative studies of spatial and temporal variation of stomatal conductance over leaf surfaces. Plant, Cell and Environment, 22, 1043–1055.CrossRefGoogle Scholar
Jones, H.G. (2004a). Applications of thermal imaging and infrared sensing in plant physiology and ecophysiology. Advances in Botanical Research Incorporating Advances in Plant Pathology, 41, 107–163.CrossRefGoogle Scholar
Jones, H.G. (2004b). Irrigation scheduling: advantages and pitfalls of plant-based methods. Journal of Experimental Botany, 55, 2427–2436.CrossRefGoogle ScholarPubMed
Jones, H.G. and Leinonen, I. (2003). Thermal imaging for the study of plant water relations. Journal of Agricultural Meteorology, Japan, 59, 205–217.CrossRefGoogle Scholar
Jones, H.G. and Morison, J. (eds) (2007). Special issue: imaging stress responses in plants. Journal of Experimental Botany, 58, 743–898.Google Scholar
Jones, H.G. and Schofield, P. (2008). Thermal and other remote sensing of plant stress. General and Applied Plant Physiology, 34, 19–32.Google Scholar
Jones, H.G. and Slatyer, R.O. (1971). Effects of intercellular resistances on estimates of the intracellular resistance to CO2 uptake by plant leaves. Australian Journal of Biological Sciences, 25, 443–453.CrossRefGoogle Scholar
Jones, H.G. and Slatyer, R.O. (1972). Estimation of the transport and carboxylation components of the intracellular limitation to leaf photosynthesis. Plant Physiology, 50, 283–288.CrossRefGoogle ScholarPubMed
Jones, H.G., Stoll, M., Santos, T., et al. (2002). Use of infrared thermography for monitoring stomatal closure in the field: application to grapevine. Journal of Experimental Botany, 53, 2249–2260.CrossRefGoogle ScholarPubMed
Jones, M.G.K., Outlaw, W.H. and Lowry, O.H. (1977). Enzymic assay of 10–7 to 10–14 moles of sucrose in plant tissues. Plant Physiology, 60, 379–383.CrossRefGoogle ScholarPubMed
Jones, T.L., Tucker, D.E. and Ort, D.R. (1998). Chilling delays circadian pattern of sucrose phosphate synthase and nitrate reductase activity in tomato. Plant Physiology, 118, 149–158.CrossRefGoogle ScholarPubMed
Jones, T.P. (1994). 13C enriched Lower Carboniferous fossil plants from Donegal, Ireland: carbon isotope constraints on taphonomy, diagenesis and palaeoenvironment. Review of Palaeobotany and Palynology, 81, 53–64.CrossRefGoogle Scholar
Jongebloed, U., Szederkényi, J., Hartig, K., et al. (2004). Sequence of morphological and physiological events during natural ageing and senescence of a castor bean leaf: sieve tube occlusion and carbohydrate back-up precede chlorophyll degradation. Physiologia Plantarum, 120, 338–346.CrossRefGoogle ScholarPubMed
Jongschaap, R.E.E., Blesgraaf, R.A.R., Bogaard, T.A., et al. (2009). The water footprint of bioenergy from Jatropha curcas L. Proceedings of the National Academy of Sciences USA, 106, E92.CrossRefGoogle ScholarPubMed
Jordan, C.F. (1969). Derivation of leaf-area index from quality of light in the forest floor. Ecology, 50, 663–666.CrossRefGoogle Scholar
Jordan, D.B. and Ögren, W.L. (1981a). A sensitive assay procedure for simultaneous determination of ribulose-1,5-bisphosphate carboxylase and oxygenase activities. Plant Physiology, 67, 237–245.CrossRefGoogle ScholarPubMed
Jordan, D.B. and Ögren, W.L. (1981b). Species variation in the specificity of ribulose bisphosphate carboxylase/oxygenase. Nature, 291, 513–515.CrossRefGoogle Scholar
Jordan, D.B. and Ögren, W.L. (1983). Species variation in kinetic properties of ribulose 1,5-bisphosphate carboxylase oxygenase. Archives of Biochemistry and Biophysics, 227, 425–433.CrossRefGoogle ScholarPubMed
Jordan, D.B. and Ögren, W.L. (1984). The CO2/O2 specificity of ribulose 1,5-bisphosphate carboxylase/oxygenase. Planta, 161, 308–313.CrossRefGoogle ScholarPubMed
Jordan, G.J., Dillon, R.A. and Weston, P.H. (2005). Solar radiation as a factor in the evolution of scleromorphic leaf anatomy in Proteaceae. American Journal of Botany, 92, 789–796.CrossRefGoogle ScholarPubMed
June, T., Evans, J.R. and Farquhar, G.D. (2004). A simple new equation for the reversible temperature dependence of photosynthetic electron transport: a study on soybean leaf. Functional Plant Biology, 31, 275–283.CrossRefGoogle Scholar
Jung, M., Reichstein, M., Ciais, P. et al. (2010). Recent decline in the global land evapotranspiration trend due to limited moisture supply. Nature, 467, 951–954.CrossRefGoogle ScholarPubMed
Junge, W. and Witt, H.T. (1968). On the ion transport system in photosynthesis: investigations on a molecular level. Zeitschrift für Naturforschung, 23b, 244–254.Google Scholar
Jurik, T.W. (1986). Temporal and spatial patterns of specific leaf weight in successional northern hardwood tree species. American Journal of Botany, 73, 1083–1092.CrossRefGoogle Scholar
Jurik, T.W. and Chabot, B.F. (1986). Leaf dynamics and profitability in wild strawberries. Oecologia, 69, 296–304.CrossRefGoogle ScholarPubMed
Jurik, T.W. and Kliebenstein, H. (2000). Canopy architecture, light extinction and self-shading of a prairie grass, Andropogon gerardii. American Midland Naturalist, 144, 51–65.CrossRefGoogle Scholar
Jurik, T.W., Weber, J.A. and Gates, D.M. (1984). Short-term effects of CO2 on gas exchange of leaves of bigtooth aspen (Populus grandidentata) in the field. Plant Physiology, 75, 1022–1026.CrossRefGoogle Scholar
Kacperska, A. (2004). Sensor types in signal transduction pathways in plant cells responding to abiotic stressors: do they depend on stress intensity?Physiologia Plantarum, 122, 159–168.CrossRefGoogle Scholar
Kaimal, J.C. and Finnigan, J.J. (1994). Atmospheric Boundary Layer Flows: Their Structure and Measurement. Oxford University Press, Oxford, UK.
Kaiser, H. (2009). The relation between stomatal aperture and gas exchange under consideration of pore geometry and diffusional resistance in the mesophyll. Plant, Cell and Environment, 32, 1091–1098.CrossRefGoogle ScholarPubMed
Kaiser, H. and Grams, T.E.E. (2006). Rapid hydropassive opening and subsequent active stomatal closure follow heat-induced electrical signals in Mimosa pudica. Journal of Experimental Botany, 57, 2087–2092.CrossRefGoogle ScholarPubMed
Kaiser, H. and Kappen, L. (2000). In situ observation of stomatal movements and gas exchange of Aegopodium podagraria L. in the understory. Journal of Experimental Botany, 51, 1741–1749.CrossRefGoogle Scholar
Kalapos, T., van den Boogaard, R. and Lambers, H. (1996). Effect of soil drying on growth, biomass allocation and leaf gas exchange of two annual grass species. Plant and Soil, 185, 137–149.CrossRefGoogle Scholar
Kalberer, S.R., Wisniewski, M. and Arora, R. (2006). Deacclimation and reacclimation of cold-hardy plants: current understanding and emerging concepts. Plant Science, 171, 3–16.CrossRefGoogle Scholar
Kamal-Eldin, A., Görgen, S., Pettersson, J., et al. (2000). Normal-phase high-performance liquid chromatography of tocopherols and tocotrienols. Comparison of different chromatographic columns. Journal of Chromatography, 881, 217–227.CrossRefGoogle Scholar
Kamaluddin, M. and Grace, J. (1992). Acclimation in seedlings of a tropical tree, Bischofia javanica, following a stepwise reduction in light. Annals of Botany, 69, 557–562.CrossRefGoogle Scholar
Kampfenkel, K., Van Montagu, M. and Inzé, D. (1995). Effect of iron excess on Nicotiana plumbaginifolia plants. Implications to oxidative stress. Plant Physiology, 107, 725–735.CrossRefGoogle ScholarPubMed
Kanai, R. and Edwards, G.E. (1999) The biochemistry of C4 photosynthesis. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, CA, USA, pp. 49–87.Google Scholar
Kanai, S., Ohkura, K., Adu-Gyamfi, J.J., et al. (2007). Depression of sink activity precedes the inhibition of biomass production in tomato plants subjected to potassium deficiency stress. Journal of Experimental Botany, 58, 2917–2928.CrossRefGoogle ScholarPubMed
Kandil, F.E., Grace, M.H., Seigler, D.S., et al. (2004). Polyphenolics in Rhizophora mangle L. leaves and their changes during leaf development and senescence. Trees, 18, 518–528.CrossRefGoogle Scholar
Kane, H.J., Viil, J., Entsch, B., et al. (1994). An improved method for measuring the CO2/O2 specificity of ribulose-bisphosphate carboxylase-oxygenase. Australian Journal of Plant Physiology, 21, 449–461.CrossRefGoogle Scholar
Kanervo, E., Tasaka, Y., Murata, N., et al. (1997). Membrane lipid unsaturation modulates processing of the photosystem II reaction-center protein d1 at low temperatures. Plant Physiology, 114, 841–849.CrossRefGoogle ScholarPubMed
Kangasjärvi, J., Talvinen, M. and Karjalainen, R. (1994). Plant defence systems induced by ozone. Plant, Cell and Environment, 17, 783–94.CrossRefGoogle Scholar
Kaplan, A. and Reinhold, L. (1999). CO2 concentrating mechanisms in photosynthetic microorganisms. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 539–570.CrossRefGoogle ScholarPubMed
Kappen, L., Andresen, G. and Lösch, R. (1987). In situ observations of stomatal movements. Journal of Experimental Botany, 38, 126–141.CrossRefGoogle Scholar
Kapralov, M.V. and Filatov, D.A. (2006). Molecular adaptation during adaptive radiation in the Hawaiian endemic genus Schiedea. PLoS ONE, 1, e8.CrossRefGoogle ScholarPubMed
Kapralov, M.V. and Filatov, D.A. (2007). Widespread positive selection in the photosynthetic Rubisco enzyme. BMC Evolutionary Biology, 7, 73.CrossRefGoogle ScholarPubMed
Karabourniotis, G. and Bornman, J.F. (1999). Penetration of UV-A, UV-B and blue light through the leaf trichome layers of two xeromorphic plants, olive and oak, measured by optical fibre microprobes. Physiologia Plantarum, 105, 655–661.CrossRefGoogle Scholar
Karabourniotis, G., Papastergiou, N., Kabanopoulou, E., et al. (1994). Foliar sclereids of Olea europaea may function as optical fibres. Canadian Journal of Botany, 72, 330–336.CrossRefGoogle Scholar
Karam., F., Lahoud, R., Massad, R., et al. (2007). Evapotranspiration, seed yield and water use efficiency of drip irrigated sunflower under full and deficit irrigation conditions. Agricultural Water Management, 90, 213–223.CrossRefGoogle Scholar
Kargul, J. and Barber, J. (2008). Photosynthetic acclimation: Structural reorganisation of light harvesting antenna: role of redox-dependent phosphorylation of major and minor chlorophyll a/b binding proteins. FEBS J, 275, 1056–1068.CrossRefGoogle ScholarPubMed
Karnosky, D.F., Mankovska, B., Percy, K., et al. (1999). Effects of tropospheric O3 on trembling aspen and interaction with CO2: results from an O3-gradient and a FACE experiment. Water Air and Soil Pollution, 116, 311–322.CrossRefGoogle Scholar
Karnosky, D.F., Zak, D.R., Pregitzer, K.S., et al. (2003). Tropospheric O3 moderates responses of temperate hardwood forests to elevated CO2: a synthesis of molecular to ecosystem results from the Aspen FACE project. Functional Ecology, 17, 289–304.CrossRefGoogle Scholar
Karolin, A.Moldau, H. (1976). A controlled environment chamber for recording of transpiration and CO2 exchange of plants shoots and roots. Fiziologia Rastenii (Sov. Plant Physiol.), 23, 630–634.Google Scholar
Karpinski, S., Reynolds, H., Karpinska, B., et al. (1999). Systemic signalling and acclimation in response to excess excitation energy in Arabidopsis. Science, 284, 654–657.CrossRefGoogle Scholar
Kasahara, M., Kagawa, T., Oikawa, K., et al. (2002). Chloroplast avoidance movement reduces photodamage in plants. Nature, 420, 829–832.CrossRefGoogle ScholarPubMed
Kasimova, M.R., Grigiene, J., Krab, K., et al. (2006). The free NADH concentration is kept constant in plant mitochondria under different metabolic conditions. Plant Cell, 18, 688–698.CrossRefGoogle ScholarPubMed
Katahata, S., Naramoto, M., Kakubari, Y., et al. (2005). Photosynthetic acclimation to dynamic changes in environmental conditions associated with deciduous overstory phenology in Daphniphyllum humile, an evergreen understory shrub. Tree Physiology, 25, 437–445.CrossRefGoogle ScholarPubMed
Katahata, S.I., Naramoto, M., Kakubari, Y., et al. (2007). Seasonal changes in photosynthesis and nitrogen allocation in leaves of different ages in evergreen understory shrub Daphniphyllum humile. Trees, 21, 619–629.CrossRefGoogle Scholar
Kato, T., Tang, Y., Gu, S., et al. (2004). Carbon dioxide exchange between the atmosphere and an alpine meadow ecosystem on the Qinghai–Tibetan Plateau, China Agricultural and Forest Meteorology, 124, 121–134.Google Scholar
Kato, M.C., Hikosaka, K. and Hirose, T. (2002). Photoinactivation and recovery of photosystem II in Chenopodium album leaves grown at different levels of irradiance and nitrogen availability. Functional Plant Biology, 29, 787–795.CrossRefGoogle Scholar
Kato, T. and Tang, Y.H. (2008). Spatial variability and major controlling factors of CO2 sink strength in Asian terrestrial ecosystems: evidence from eddy covariance data. Global Change Biology, 14, 2333–2348.CrossRefGoogle Scholar
Kattge, J. and Knorr, W. (2007). Temperature acclimation in a biochemical model of photosynthesis: a reanalysis of data from 36 species. Plant, Cell and Environment, 30, 1176–1190.CrossRefGoogle Scholar
Kautsky, H. and Hirsch, A. (1931). Neue Versuche zur Kohlensäureassimilation. Naturwissenschaften, 19, 964.CrossRefGoogle Scholar
Kaya, C., Higgs, D., Kirnak, H., et al. (2003). Mycorrhizal colonisation improves fruit yield and water use efficiency in watermelon (Citrullus lanatus Thunb.) grown under well-watered and water-stressed conditions. Plant and Soil, 253, 287–292.CrossRefGoogle Scholar
Kebeish, R., Niessen, M., Thiruveedhi, K., et al. (2007). Chloroplastic photorespiratory bypass increases photosynthesis and biomass production in Arabidopsis thaliana. Nature Biotechnology, 25, 593–599.CrossRefGoogle ScholarPubMed
Kee, S.C., Martin, B. and Ort, D.R. (1986). The effects of chilling in the dark and in the light on photosynthesis of tomato – electron-transfer reactions. Photosynthesis Research, 8, 41–51.CrossRefGoogle ScholarPubMed
Keeley, J.E. (1990). Photosynthetic pathways in freshwater aquatic plants. Trends in Ecology and Evolution, 5, 330–333.CrossRefGoogle ScholarPubMed
Keeley, J.E. (1996). Aquatic CAM photosynthesis. In: Crassulacean Acid Metabolism. Biochemistry, Ecophysiology and Evolution (eds Winter, K. and Smith, J.A.C.) Ecological Studies vol. 114. Springer-Verlag, Berlin – Heidelberg – New York, pp. 281–295.Google Scholar
Keeley, J.E. (1998a). CAM photosynthesis in submerged aquatic plants. Botanical Review, 64, 121–175.CrossRefGoogle Scholar
Keeley, J.E. (1998b). C4 photosynthetic modifications in the evolutionary transition from land to water in aquatic grasses. Oecologia, 116, 85–97.CrossRefGoogle ScholarPubMed
Keeley, J.E. and Busch, G. (1984). Carbon assimilation characteristics of the aquatic CAM plant, Isoetes howellii. Plant Physiology, 76, 525–530.CrossRefGoogle ScholarPubMed
Keeley, J.E. and Keeley, S.C. (1989). Crassulacean acid metabolism (CAM) in high elevation tropical cactus. Plant, Cell and Environment, 12, 331–336.CrossRefGoogle Scholar
Keeley, J.E. and Rundel, P.W. (2003). Evolution of CAM and C4 carbon-concentrating mechanisms. International Journal of Plant Sciences, 164 (3 Suppl.), S55–S77.CrossRefGoogle Scholar
Keeley, J.E. and Rundel, P.W. (2005). Fire and the Miocene expansion of C4 grasslands. Ecology Letters, 8, 683–690.CrossRefGoogle Scholar
Keeley, J.E., Osmond, C.B. and Raven, J.A. (1984). Stylites, a vascular land plant without stomata absorbs CO2 via its roots. Nature, 310, 694–695.CrossRefGoogle Scholar
Keeling, C.D. (1958). The concentration and isotopic abundance of atmospheric carbon dioxide in rural and marine areas. Geochimica et Cosmochimica Acta, 13, 322–334.CrossRefGoogle Scholar
Keeling, C.D., Chin, J.F.S. and Whorf, T.P. (1996). Increased activity of northern vegetation inferred from atmospheric CO2 measurements. Nature, 382, 146–149.CrossRefGoogle Scholar
Keeling, C.D., Mook, W.G. and Tans, P.P. (1979). Recent trends in the 13C/12C ratio of atmospheric carbon dioxide. Nature, 277, 121–123.CrossRefGoogle Scholar
Keeling, R.F., Piper, S.C., Bollenbacher, , et al. (2009). Atmospheric CO2 records from sites in the SIO air sampling network. In: Trends: A Compendium of Data on Global Change. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn., U.S.A.
Keenan, T., Sabate, S. and Gracia, C. (2010a). The importance of mesophyll conductance in regulating forest ecosystem productivity during drought periods. Global Change Biology, 16, 1019–1034.CrossRefGoogle Scholar
Keenan, T., Sabate, S. and Gracia, C. (2010b). Soil water stress and coupled photosynthesis–conductance models: bridging the gap between conflicting reports on the relative roles of stomatal, mesophyll conductance and biochemical limitations to photosynthesis. Agricultural and Forest Meteorology, 150, 443–453.CrossRefGoogle Scholar
Keitel, C., Adams, M.A., Holst, T., et al. (2003). Carbon and oxygen isotope composition of organic compounds in the phloem sap provides a short-term measure of stomatal conductance of European beech (Fagus sylvatica L.). Plant, Cell and Environment, 26, 1157–1168.CrossRefGoogle Scholar
Keitel, C., Matzarakis, A., Rennenberg, H., et al. (2006). Carbon isotopic composition and oxygen isotopic enrichment in phloem and total leaf organic matter of European beech (Fagus sylvatica L.) along a climate gradient. Plant, Cell and Environment, 29, 1492–1507.CrossRefGoogle ScholarPubMed
Kellogg, E.A. (1999). Phylogenetic aspects of the evolution of C4 photosynthesis. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.) Academic Press, San Diego, CA, USA, pp. 411–444.Google Scholar
Kellogg, E.A. (2001). Evolutionary history of the grasses. Plant Physiology, 125, 1198–1205.CrossRefGoogle ScholarPubMed
Kellomaki, S., Wang, K.Y. and Lemettinen, M. (2000). Controlled environment chambers for investigating tree response to elevated CO2 and temperature under boreal conditions. Photosynthetica, 38, 69–81.CrossRefGoogle Scholar
Kemp, P.R. and Williams, G.J., III (1980), A physiological basis for niche separation between Agropyron smithii (C3) and Bouteloua gracilis (C4). Ecology, 61, 846–858.CrossRefGoogle Scholar
Kempema, L.A., Cui, X., Holzer, F.M., et al. (2007). Arabidopsis transcriptome changes in response to phloem-feeding silverleaf whitefly nymphs. Similarities and distinctions in responses to aphids. Plant Physiology, 143, 849–865.CrossRefGoogle ScholarPubMed
Kenrich, P. and Crane, P.R. (1997). The origin and early evolution of plants on land. Nature, 389, 33–39.CrossRefGoogle Scholar
Kent, S.S. and Tomany, M.J. (1995). The differential of the ribulose 1,5-bisphosphate carboxylase/oxygenase specificity factor among higher plants and the potential for biomass enhancement. Plant Physiology and Biochemistry, 33, 71–80.Google Scholar
Kerstiens, G. (1996). Cuticular water permeability and its physiological significance. Journal of Experimental Botany, 47, 1813–1832.CrossRefGoogle Scholar
Kerstiens, G. (2006). Water transport in plant cuticles: an update. Journal of Experimental Botany, 57, 2493–2499.CrossRefGoogle ScholarPubMed
Keskitalo, J., Bergquist, G., Gardeström, P., et al. (2005). A cellular timetable of autumn senescence. Plant Physiology, 139, 1635–1648.CrossRefGoogle ScholarPubMed
Kesselmeier, J. and Staudt, M. (1999). Biogenic volatile organic compounds (VOC): an overview on emission, physiology and ecology. Journal of Atmospheric Chemistry, 33, 23–88.CrossRefGoogle Scholar
Kessler, A. and Baldwin, I.T. (2002). Plant responses to insect herbivory: the emerging molecular analysis. Annual Review of Plant Biology, 53, 299–328.CrossRefGoogle ScholarPubMed
Kessler, M., Siorak, Y., Wunderlich, M., et al. (2007). Patterns of morphological leaf traits among pteridophytes along humidity and temperature gradients in the Bolivian Andes. Functional Plant Biology, 34, 963–971.CrossRefGoogle Scholar
Kessler, S. and Sinha, N. (2004). Shaping up: the genetic control of leaf shape. Current Opinion in Plant Biology, 7, 65–72.CrossRefGoogle ScholarPubMed
Kettle, A.J., Kuhn, U., von Hobe, M., et al. (2002). Global budget of atmospheric carbonyl sulfide: temporal and spatial variations of the dominant sources and sinks. Journal of Geophysical Research – Atmospheres, 107, Article number 4658.CrossRefGoogle Scholar
Khalil, A.A.M. and Grace, J. (1993). Does xylem sap ABA control the stomatal behaviour of water-stressed sycamore (Acer pseudoplatanus L.) seedlings?Journal of Experimental Botany, 44, 1127–1134.CrossRefGoogle Scholar
Khamis, S., Lamaze, T., Lemoine, Y., et al. (1990). Adaptation of the photosynthetic apparatus in maize leaves as a result of nitrogen limitation. Relationships between electron transport and carbon assimilation. Plant Physiology, 94, 1436–1443.CrossRefGoogle ScholarPubMed
Khan, M.S. (2007). Engineering photorespiration in chloroplasts: a novel strategy for increasing biomass production. Trends in Biotechnology, 25, 437–40.CrossRefGoogle ScholarPubMed
Khandelwal, A., Elvitigala, T., Ghosh, B., et al. (2008). Arabidopsis transcriptome reveals control circuits regulating redox homeostasis and the role of an AP2 transcription factor. Plant Physiology, 148, 2050–2058.CrossRefGoogle ScholarPubMed
Kikuzawa, K. (1983). Leaf survival of woody plants in deciduous broad-leaved forests. 1. Tall trees. Canadian Journal of Botany, 61, 2133–2139.CrossRefGoogle Scholar
Kikuzawa, K. (1991). A cost-benefit analysis of leaf habit and leaf longevity of trees and their geographical pattern. American Naturalist, 138, 1250–1260.CrossRefGoogle Scholar
Kilian, J., Whitehead, D., Horak, J., et al. (2007). The AtGenExpress global stress expression data set: protocols, evaluation and model data analysis of UV-B light, drought and cold stress responses. The Plant Journal, 50, 347–363.CrossRefGoogle ScholarPubMed
Killingbeck, K.T. (1996). Nutrients in senesced leaves: keys to the search for potential resorption and resorption proficiency. Ecology, 77, 1716–1727.CrossRefGoogle Scholar
Kiltie, R.A. (1993). New light on forest shade. Trends in Ecology and Evolution, 8, 39–40.CrossRefGoogle ScholarPubMed
Kim, C.S. and Jung, J. (1993). The susceptibility of mung bean chloroplasts to photoinhibition is increased by an excess supply of iron to plants: a photobiological aspect of iron toxicity in plant leaves. Photochemistry and Photobiology, 58, 120–126.CrossRefGoogle Scholar
Kim, J. (2007). Perception, transduction, and networks in cold signaling. Journal of Plant Biology, 50, 139–147.CrossRefGoogle Scholar
Kim, M.S., Lefcourt, A.M. and Chen, Y.C. (2003). Multispectral laser-induced fluorescence imaging system for large biological samples. Applied Optics, 42, 3927–3934.CrossRefGoogle ScholarPubMed
Kim, T.H., Böhmer, M., Hu, H., et al. (2010). Guard cell signal transduction network: advances in understanding abscisic acid, CO2, and Ca2+ signaling. Annual Review of Plant Biology, 61, 561–591.CrossRefGoogle ScholarPubMed
Kim, Y.X. and Steudle, E. (2009). Gating of aquaporins by light and reactive oxygen species in leaf parenchyma cells of the midrib of Zea mays. Journal of Experimental Botany, 60, 547–556.CrossRefGoogle ScholarPubMed
Kim, S.H, Sicher, R.C., Bae, H., et al. (2006). Canopy photosynthesis, evapotranspiration, leaf nitrogen, and transcription profiles of maize in response to CO2 enrichment. Global Change Biology, 12, 588–600.CrossRefGoogle Scholar
Kimball, B. and Bernacchi, C.J. (2006). Evapotranspiration, canopy temperature, and plant water relations. In: Managed Ecosystems and Rising CO2: Case Studies, Processes, and Perspectives (eds Nösberger, J. and Blum, H.), Springer Verlag, Berlin, Germany, pp. 311–324.Google Scholar
Kimball, B.A., Pinter, P.J., Garcia, R.L., et al. (1995). Productivity and water use of wheat under free-air CO2 enrichment. Global Change Biology, 1, 429–442.CrossRefGoogle Scholar
Kimura, K., Ishida, A., Uemura, A., et al. (1998). Effects of current-year and previous-year PPFDs on shoot gross morphology and leaf properties in Fagus japonica. Tree Physiology, 18, 459–466.CrossRefGoogle ScholarPubMed
King, J.S., Hanson, P.J., Bernhardt, E., et al. (2004). A multiyear synthesis of soil respiration responses to elevated atmospheric CO2 from four forest FACE experiments. Global Change Biology, 10, 1027–1042.CrossRefGoogle Scholar
King, S.P., Badger, M.R. and Furbank, R.T. (1998). CO2 characteristics of developing canola seeds and silique wall. Australian Journal of Plant Physiology, 25, 377–386.CrossRefGoogle Scholar
Kingston-Smith, A.H., Harbinson, J., Williams, J., et al. (1997). Effect of chilling on carbon assimilation, enzyme activation, and photosynthetic electron transport in the absence of photoinhibition in maize leaves. Plant Physiology, 114, 1039–1046.CrossRefGoogle ScholarPubMed
Kiniry, J.R. (1998). Biomass accumulation and radiation use efficiency of honey mesquite and eastern red cedar. Biomass and Bioenergy, 15, 467–73.CrossRefGoogle Scholar
Kiniry, J.R., Jones, C.A., O’Toole, J.C., et al. (1989). Radiation-use efficiency in biomass accumulation prior to grain-filling for five grain-crop species. Field Crops Research, 20, 51–64.CrossRefGoogle Scholar
Kira, T. (1975). Primary production of forests. In: Photosynthesis and Productivity in Different Environments (eds Winter, K. and Smith, J.A.C.), Cambridge University Press, Cambridge, UK, pp. 5–40.
Kirschbaum, M.U. (2004). Direct and indirect climate change effects on photosynthesis and transpiration. Plant Biology, 6, 242–53.CrossRefGoogle ScholarPubMed
Kirschbaum, M.U.F. (1987). Water-stress in Eucalyptus pauciflora – comparison of effects on stomatal conductance with effects on the mesophyll capacity for photosynthesis, and investigation of a possible involvement of photoinhibition. Planta, 171, 466–473.CrossRefGoogle ScholarPubMed
Kirschbaum, M.U.F. (1988). Recovery of photosynthesis from water stress in Eucalyptus pauciflora – a process in two stages. Plant, Cell and Environment, 11, 685–694.CrossRefGoogle Scholar
Kirschbaum, M.U.F. and Pearcy, R.W. (1988). Concurrent measurements of oxygen-and carbon-dioxide exchange during lightflecks in Alocasia macrorrhiza (L.) G. Don. Planta, 174, 527–533.CrossRefGoogle Scholar
Kirschbaum, M.U.F. and Tompkins, D. (1990). Photosynthetic responses to phosphorus nutrition in Eucalyptus grandis seedlings. Australian Journal of Plant Physiology, 17, 527–535.CrossRefGoogle Scholar
Kirschbaum, M.U.F., Keith, H., Leuning, R., et al. (2007) Modelling net ecosystem carbon and water exchange of a temperate Eucalyptus delegatensis forest using multiple constraints. Agricultural and Forest Meteorology, 145, 48–68.CrossRefGoogle Scholar
Kitajima, K. (1994). Relative importance of photosynthetic traits and allocation patterns as correlates of seedling shade tolerance of 13 tropical trees. Oecologia, 98, 419–428.CrossRefGoogle ScholarPubMed
Kitajima, K. and Hogan, K.P. (2003). Increases of chlorophyll a/b ratios during acclimation of tropical woody seedlings to nitrogen limitation and high light. Plant, Cell and Environment, 26, 857–865.CrossRefGoogle ScholarPubMed
Kitajima, K., Mulkey, S.S., Samaniego, M., et al. (2002). Decline of photosynthetic capacity with leaf age and position in two tropical pioneer tree species. American Journal of Botany, 89, 1925–1932.CrossRefGoogle ScholarPubMed
Kitajima, K., Mulkey, S.S., and Wright, S.J. (1997). Decline of photosynthetic capacity with leaf age in relation to leaf longevities for five tropical canopy tree species. American Journal of Botany, 84, 702–708.CrossRefGoogle ScholarPubMed
Kitao, M., Utsugi, H., Kuramoto, S., et al. (2003) Light-dependent photosynthetic characteristics indicated by chlorophyll fluorescence in five mangrove species native to Pohnpei Island, Micronesia. Physiologia Plantarum, 117, 376–382.CrossRefGoogle ScholarPubMed
Kitaoka, S. and Koike, T. (2004). Invasion of broad-leaf tree species into a larch plantation: seasonal light environment, photosynthesis and nitrogen allocation. Physiologia Plantarum, 121, 604–611.CrossRefGoogle Scholar
Kjellström, E., Bärring, L., Jacob, D., et al. (2007). Modelling daily temperature extremes: recent climate and future changes over Europe. Climatic Change, 81, S249–S265.CrossRefGoogle Scholar
Klein, T., Hemming, D., Lin, T., et al. (2005). Association between tree-ring and needle delta 13C and leaf gas exchange in Pinus halepensis under semi-arid conditions. Oecologia, 144, 45–54.CrossRefGoogle Scholar
Kleine, T., Kindgren, P., Benedict, C., et al. (2007). Genome-wide gene expression analysis reveals a critical role for CRYPTOCHROME1 in the response of Arabidopsis to high irradiance. Plant Physiology 144, 1391–1406.CrossRefGoogle ScholarPubMed
Kleinert, K. and Strecker, M.R. (2001). Climate change in response to orographic barrier uplift: paleosol and stable evidence from the late Neogene Santa Maria Basin, norwestern Argentina. Bulletin of the Geological Society of America, 113, 728–742.2.0.CO;2>CrossRefGoogle Scholar
Kliemchen, A., Schomburg, M., Galla, H.J., et al. (1993). Phenotypic changes in the fluidity of the tonoplast membrane of crassulacean acid metabolism plants in response to temperature and salinity stress. Planta, 189, 403–409.CrossRefGoogle ScholarPubMed
Klingeman, W.E., Buntin, G.D., van Iersel, M.W., et al. (2000). Whole-plant gas exchange, not individual-leaf measurements, accurately assesses azalea response to insecticides. Crop Protection, 19, 407–415.CrossRefGoogle Scholar
Klingeman, W.E., van Iersel, M.W., Kang, J.G., et al. (2005). Whole-plant gas exchange measurements of mycorrhizal ‘iceberg’ roses exposed to cyclic drought. Crop Protection, 24, 309–317.CrossRefGoogle Scholar
Klingler, J.P., Batelli, G. and Zhu, J.-K. (2010). ABA receptors: the START of a new paradigm in phytohormone signalling. Journal of Experimental Botany, 61, 3199–3210.CrossRefGoogle ScholarPubMed
Kluge, M. and Brulfert, J. (2000). Ecophysiology of vascular plants on inselbergs. In: Inselbergs. Ecological Studies, Vol. 146 (eds Porembski, S. and Barthlott, W.), Springer-Verlag, Berlin – Heidelberg – New York, pp. 143–174.Google Scholar
Kluge, M., Kliemchen, A. and Galla, H.J. (1991). Temperature effects on crassulacean acid metabolism: EPR spectroscopic studies on the thermotropic phase behaviour of the tonoplast membranes of Kalanchoë daigremontiana. Botanica Acta, 104, 355–360.CrossRefGoogle Scholar
Klughammer, C. and Schreiber, U. (1994). An improved method, using saturating light pulses, for the determination of photosystem I quantum yield via P700+-absorbance changes at 830 nm. Planta, 192, 261–268.CrossRefGoogle Scholar
Knapp, A.K. (1993). Gas exchange dynamics in C3 and C4 grasses: consequences of differences in stomatal conductance. Ecology, 74, 113–123.CrossRefGoogle Scholar
Knapp, A.K. and Smith, M.D. (2001). Variation among biomes in temporal dynamics of aboveground primary production. Science, 291, 481–484.CrossRefGoogle ScholarPubMed
Knapp, A.K., Briggs, J.M. and Koelliker, J.K. (2001). Frequency and extent of water limitation to primary production in a mesic temperate grassland. Ecosystems, 4, 19–28.CrossRefGoogle Scholar
Knapp, A.K., Fay, P.A., Blair, J.M., et al. (2002). Rainfall variability, carbon cycling, and plant species diversity in a mesic grassland. Science, 298, 202–2205.CrossRefGoogle Scholar
Knight, C.A. and Ackerly, D.D. (2003a). Evolution and plasticity of photosynthetic thermal tolerance, specific leaf area and leaf size: congeneric species from desert and coastal environments.. New Phytologist, 160, 337–47.CrossRefGoogle Scholar
Knight, C.A. and Ackerly, D.D. (2003b). Small heat shock protein responses of a closely related pair of desert and coastal Encelia. International Journal of Plant Sciences, 164, 53–60.CrossRefGoogle Scholar
Knight, C.A. and Ackerly, D.D. (2002). An ecological and evolutionary analysis of photosynthetic thermotolerance using the temperature-dependent increase in fluorescence. Oecologia, 130, 505–514.CrossRefGoogle ScholarPubMed
Knoop, W. and Walker, B. (1985). Interactions of woody and herbaceous vegetation in a southern African savanna. Journal of Ecology, 73, 235–253.CrossRefGoogle Scholar
Knorre, A.A., Kirdyanov, A.V. and Vaganov, E.A. (2006). Climatically induced interannual variability in aboveground production in forest-tundra and northern taiga of central Siberia. Oecologia, 147, 86–95.CrossRefGoogle ScholarPubMed
Kobayashi, M., Sasaki, K., Enomoto, M., et al. (2007). Highly sensitive determination of transient generation of biophotons during hypersensitive response to cucumber mosaic virus in cowpea. Journal of Experimental Botany, 58, 465–472.CrossRefGoogle ScholarPubMed
Koç, M., Barutçular, C. and Genç, I. (2003). Photosynthesis and productivity of old and modern durum wheats in a Mediterranean environment. Crop Science, 43, 2089–2098.CrossRefGoogle Scholar
Koch, C., Noga, G., and Strittmatter, G. (1994). Photosynthetic electron transport is differentially affected during early stages of cultivar/race specific interactions between potato and Phytophthora infestans. Planta, 193, 551–557.CrossRefGoogle Scholar
Koch, G.W., Sillett, S.C., Jennings, G.M., et al. (2004). The limits to tree height. Nature, 428, 851–854.CrossRefGoogle ScholarPubMed
Koch, W., Lange, O.L. and Schulze, E.D. (1971). Ecophysiological investigations on wild and cultivated plants in the Negev desert. Oecologia, 8, 296–309.CrossRefGoogle ScholarPubMed
Kocsy, G., Galiba, G. and Brunold, C. (2001). Role of glutathione in adaptation and signalling during chilling and cold acclimation in plants. Physiologia Plantarum, 113, 158–164.CrossRefGoogle ScholarPubMed
Koenning, S.R. and Barker, K.R. (1995). Soybean photosynthesis and yield as influenced by Heterodera glycines, soil type and irrigation. Journal of Nematology, 27, 51–62.Google ScholarPubMed
Koeppe, D.E. and Miller, R.J. (1970). Lead effects on corn mitochondrial respiration. Science, 167, 1376–1378.CrossRefGoogle ScholarPubMed
Kogami, H., Hanba, Y.T., Kibe, T., et al. (2001). CO2 transfer conductance, leaf structure and carbon isotope composition of Polygonum cuspidatum leaves from low and high altitudes. Plant, Cell and Environment, 24, 529–538.CrossRefGoogle Scholar
Kohen, E., Santus, R. and Hirschberg, J.G. (1995). Photobiology. Academic Press, London.
Kohzuma, K., Cruz, J.A., Akashi, K., et al. (2009). The long-term responses of the photosynthetic proton circuit to drought. Plant, Cell and Environment, 32, 209–219.CrossRefGoogle ScholarPubMed
Koike, T. (1987). Photosynthesis and expansion in leaves of early, mid, and late successional tree species, birch, ash, and maple. Photosynthetica, 21, 503–508.Google Scholar
Koike, T. (1990). Autumn colouring, photosynthetic performance and leaf development of deciduous broad-leaved trees in relation to forest succession. Tree Physiology, 7, 21–32.CrossRefGoogle Scholar
Koike, T., Kitaoka, S., Masyagina, O.V., et al. (2007). Nitrogen dynamics of leaves of deciduous broad-leaved tree seedlings grown in summer green forests in northern Japan. Eurasian Journal of Forestry Research, 10, 115–119.Google Scholar
Koiwa, H., Kojima, M., Ikeda, T., et al. (1992). Fluctuations of particles on chloroplast thylakoid membranes in tomato plants infected with virulent or attenuated strain of tobacco mosaic virus. Annals of the Phytopathological Society of Japan, 58, 58–64.CrossRefGoogle Scholar
Koizumi, H. and Oshima, Y. (1985). Seasonal changes in photosynthesis of four understory herbs in deciduous forests. The Botanical Magazine, Tokyo, 98, 1–13.CrossRefGoogle Scholar
Koizumi, M., Takahashi, K., Mineuchi, K., et al. (1998). Light gradients and the transverse distribution of chlorophyll fluorescence in mangrove and Camellia leaves. Annals of Botany London, 81, 527–533.CrossRefGoogle Scholar
Kok, B. (1948). A critical consideration of the quantum yield of Chlorella photosynthesis. Enzymologia, 13, 1–56.Google Scholar
Kokkinos, C.N., Clark, C.A.,. McGregorl, C.E., et al. (2006). The effect of sweet potato virus disease and its viral components on gene expression levels in sweetpotato. Journal of American Society for Horticultural Science, 13, 657–666.Google Scholar
Kolari, P., Pumpanen, J., Rannik, U., et al. (2004). Carbon balance of different aged Scots pine forests in Southern Finland. Global Change Biology, 10, 1106–1119.CrossRefGoogle Scholar
Kolb, C.A., Kopecký, J., Riederer, R., et al. (2003). UV screening by phenolics in berries of grapevine (Vitis vinifera). Functional Plant Biology, 30, 1177–1186.CrossRefGoogle Scholar
Kolbe, A., Tiessen, A., Schluepmann, H., et al. (2005). Trehalose-6-phosphate regulates starch synthesis via post-translational redox activation of ADP-glucose pyrophosphorylase. Proceedings of the National Academy of Sciences USA, 102, 11118–11123.CrossRefGoogle ScholarPubMed
Kolber, Z., Klimov, D., Ananyev, G., et al. (2005), Measuring photosynthetic parameters at a distance: laser induced fluorescence transient (LIFT) method for remote measurements of photosynthesis in terrestrial vegetation. Photosynthesis Research, 84, 121–129.CrossRefGoogle Scholar
Kolber, Z. S., Prasil, O. and Falkowski, P.G. (1998). Measurements of variable chlorophyll fluorescence using fast repetition rate techniques. I. Defining methodology and experimental protocols. Biochimica et Biophysica Acta, 1367, 88–106.CrossRefGoogle Scholar
Kollist, T., Moldau, H., Rasulov, B., et al. (2007). A novel device detects a rapid ozone-induced transient stomatal closure in intact Arabidopsis and its absence in abi2 mutant. Physiologia Plantarum, 129, 796–803.CrossRefGoogle Scholar
Koltai, H. and Mckenzie Bird, D. (2000). High throughput cellular localization of specific plant mRNAs by liquid-phase in situ reverse transcription-polymerase chain reaction of tissue sections. Plant Physiology, 123, 1203–1212.CrossRefGoogle ScholarPubMed
Koop, H.U., Herz, S., Golds, T.J., et al. (2008). Top Current Genetics, 19, 457–510.CrossRef
Köppen, W. (1936). Das geographische System der Klimate. In: Handbuch der Klimatologie (eds Köppen, W. and Geiger, G.), Gebrüder Borntraeger, Berlin, Germany, pp. 1–46.Google Scholar
Koricheva, J., Larsson, S., Haukioja, E., et al. (1998). Regulation of woody plant secondary metabolism by resource availability: hypothesis testing by means of metaanalysis. Oikos, 83, 212–226.CrossRefGoogle Scholar
Körner, C, Asshoff, R., Bignucolo, O., et al. (2005). Carbon flux and growth in mature deciduous forest trees exposed to elevated CO2. Science, 309, 1360–1362.CrossRefGoogle ScholarPubMed
Körner, C. (1991). Some often overlooked plant characteristics as determinants of plant growth: a reconsideration. Functional Ecology, 5, 162–173.CrossRefGoogle Scholar
Körner, C. (1998). A re-assessment of high elevation treeline positions and their explanation. Oecologia, 115, 445–459.Google ScholarPubMed
Körner, C. and Pelaez Menendez-Riedl, S. (1990). The significance of developmental aspects in plant growth analysis. In: Causes and Consequences of Variation in Growth Rate and Productivity of Higher Plants (eds Köppen, W. and Geiger, G.). SPB Academic Publishing, The Hague, Netherlands, pp. 141–157.Google Scholar
Körner, C., Scheel, J.A. and Bauer, H. (1979). Maximum leaf diffusive conductance in vascular plants. Photosynthetica, 13, 45–82.Google Scholar
Körner, C. (1999). Alpine plant life. Functional plant ecology of high mountain ecosystems. Springer Verlag, Berlin, Heidelberg.CrossRef
Körner, C. (2007). The use of ‘altitude’ in ecological research. Trends in Ecology and Evolution, 22, 569–574.CrossRefGoogle ScholarPubMed
Körner, C. and Diemer, M. (1987). In situ photosynthetic responses to light, temperature and carbon dioxide in herbaceous plants from low and high altitude. Functional Ecology, 1, 179–194.CrossRefGoogle Scholar
Körner, C., Allison, A. and Hilscher, H. (1983). Altitudinal variations in leaf diffusive conductance and leaf anatomy in heliophytes of montane New Guinea and their interrelation with microclimate. Flora, 174, 91–135.CrossRefGoogle Scholar
Körner, C., Farquhar, G.D. and Wong, S.C. (1991). Carbon isotope discrimination by plants follows latitudinal and altitudinal trends. Oecologia, 88, 30–40.CrossRefGoogle ScholarPubMed
Kornyeyev, D., Logan, B.A., Payton, P.R., et al. (2003). Elevated chloroplastic glutathione reductase activities decrease chilling-induced photoinhibition by increasing rates of photochemistry, but not thermal energy dissipation, in transgenic cotton. Functional Plant Biology, 30, 101–110.CrossRefGoogle Scholar
Kortschak, H.P., Hartt, C.E. and Burr, G.O. (1965). Carbon dioxide fixation in sugarcane leaves. Plant Physiology, 40, 209–213.CrossRefGoogle ScholarPubMed
Köstner, B., Falge, E. and Tenhunen, J.D. (2002). Age-related effects on leaf area/sapwood area relationships, canopy transpiration and carbon gain of Norway spruce stands (Picea abies) in the Fichtelgebirge, Germany. Tree Physiology, 22, 567–574.CrossRefGoogle ScholarPubMed
Köstner, B., Granier, A. and Chermák, J. (1998). Sapflow measurements in forest stands: methods and uncertainties. Annals of Forest Science, 55, 13–27.CrossRefGoogle Scholar
Kosugi, Y., Takanashi, S., Ohkubo, S., et al. (2008). CO2 exchange of a tropical rainforest at Pasoh in Peninsular Malaysia. Agricultural and Forest Meteorology, 148, 439–452.CrossRefGoogle Scholar
Kottapalli, K.R., Rakwal, R., Shibato, J., et al. (2009). Physiology and proteomics of the water-deficit stress response in three contrasting peanut genotypes. Plant, Cell and Environment, 32, 380–407.CrossRefGoogle ScholarPubMed
Koussevitzky, S., Nott, A., Mockler, T.C., et al. (2007). Signals from chloroplasts converge to regulate nuclear gene expression. Science, 316, 715–719.CrossRefGoogle ScholarPubMed
Kowalski, A., Loustau, D., Berbigier, P., et al. (2004). Paired comparison of carbon exchange between undisturbed and regenerating stands in four managed forests in Europe. Global Change Biology, 10, 1707–1723.CrossRefGoogle Scholar
Kozaki, A. and Takeba, G. (1996). Photorespiration protects C3 plants from photooxidation. Nature, 384, 557–560.CrossRefGoogle Scholar
Kozela, C. and Regan, S. (2003). How plants make tubes. Annals of Forest Science, 8, 159–164.Google ScholarPubMed
Kozlowski, T.T. (1984). Plant responses to flooding of soil. BioScience, 34, 162–167.CrossRefGoogle Scholar
Krall, J.P. and Edwards, G.E. (1992). Relationship between photosystem II activity and CO2 fixation in leaves. Physiologia Plantarum, 86, 180–187.CrossRefGoogle Scholar
Kramer, D.M., Avenson, T.J. and Edwards, G.E. (2004b). Dynamic flexibility in the light reactions of photosynthesis governed by both electron and proton transfer reactions. Trends in Plant Sciences, 9, 349–357.CrossRefGoogle ScholarPubMed
Kramer, D.M., Cruz, J.A. and Kanazawa, A. (2003). Balancing the central roles of the thylakoid proton gradient. Trends in Plant Sciences, 8, 27–32.CrossRefGoogle ScholarPubMed
Kramer, D.M., Johnson, G., Kiirats, O., et al. (2004a). New fluorescence parameters for the determination of Qa redox state and excitation energy fluxes. Photosynthesis Research, 79, 209–218.CrossRefGoogle ScholarPubMed
Kramer, K., Friend, A. and Leinonen, I. (1996). Modelling comparison to evaluate the importance of phenology and spring frost damage for the effects of climate change on growth of mixed temperate-zone deciduous forests. Climate Research, 7, 31–41.CrossRefGoogle Scholar
Krapp, A. and Stitt, M. (1995). An evaluation of direct and indirect mechanisms for the sink-regulation of photosynthesis in spinach – changes in gas-exchange, carbohydrates, metabolites, enzyme-activities and steady state transcript levels after cold-girdling source leaves. Planta, 195, 313–323.CrossRefGoogle Scholar
Krause, G.H. (1988). Photoinhibition of photosynthesis. An evaluation of damaging and protective mechanisms. Physiologia Plantarum, 74, 566–574.CrossRefGoogle Scholar
Krause, G.H. and Weis, E. (1991). Chlorophyll fluorescence and photosynthesis: the basics. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 313–349.CrossRefGoogle Scholar
Krause, G.H. and Winter, K. (1996). Photoinhibition of photosynthesis in plants growing in natural tropical forest gaps. A chlorophyll fluorescence study. Botanica Acta, 109, 456–462.CrossRefGoogle Scholar
Krause, G.H., Galle, A., Gademann, R., et al. (2003b). Capacity of protection against ultraviolet radiation in sun and shade leaves of tropical forest plants. Functional Plant Biology, 30, 533–542.CrossRefGoogle Scholar
Krause, G.H., Grube, E., Koroleva, O.Y., et al. (2004). Do mature shade leaves of tropical tree seedlings acclimate to high sunlight and UV radiation?Functional Plant Biology, 31, 743–756.CrossRefGoogle Scholar
Krause, G.H., Grube, E., Virgo, A., et al. (2003a). Sudden exposure to solar UV-B radiation reduces net CO2 uptake and photosystem I efficiency in shade-acclimated tropical tree seedlings. Plant Physiology, 131, 745–752.CrossRefGoogle Scholar
Krause, G.H., Jahns, P., Virgo, A., et al. (2007). Photoprotection, photosynthesis and growth of tropical tree seedlings under near-ambient and strongly reduced solar ultraviolet-B radiation. Journal of Plant Physiology, 164, 1311–1322.CrossRefGoogle ScholarPubMed
Krause, G.H., Koroleva, O.Y., Dalling, J.W., et al. (2001). Acclimation of tropical tree seedlings to excessive light in simulated tree-fall gaps. Plant, Cell and Environment, 24, 1345–1352.CrossRefGoogle Scholar
Krause, G.H., Schmude, C., Garden, H., et al. (1999). Effects of solar ultraviolet radiation on the potential efficiency of photosystem II in leaves of tropical plants. Plant Physiology, 121, 1349–1358.CrossRefGoogle ScholarPubMed
Krause, G.H., Virgo, A. and Winter, K. (1995). High susceptibility to photoinhibition of young leaves of tropical forest trees. Planta, 197, 583–591.CrossRefGoogle Scholar
Krauss, K.W. and Allen, J.A. (2003). Influences of salinity and shade on seedling photosynthesis and growth of two mangrove species, Rhizophora mangle and Bruguiera sexangula, introduced to Hawaii. Aquatic Botany, 77, 311–324.CrossRefGoogle Scholar
Krauss, K.W., Lovelock, C.E., McKee, K.L., et al. (2008). Environmental drivers in mangrove establishment and early development: a review. Aquatic Botany, 89, 105–127.CrossRefGoogle Scholar
Krestov, P.V. (2003). Forest Vegetation of Easternmost Russia (Russian Far East). In: Forest Vegetation of Northeast Asia (eds Kolbek, J., Srutek, M. and Box, E.). Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 93–180.Google Scholar
Kriedemann, P.E. and Anderson, J.E. (1988). Growth and photosynthetic responses to manganese and copper deficiencies in wheat (Triticum aestivum) and barley grass (Hordeum glaucum and H. leporinum). Australian Journal of Plant Physiology, 15, 429–446.CrossRefGoogle Scholar
Kriedemann, P.E. and Downton, W.J.S. (1981). Photosynthesis. In: The Physiology and Biochemistry of Drought Resistance in Plants (eds Paleg, L.G. and Aspinall, D.), Academic Press, Sydney, Australia, pp. 283–314.Google Scholar
Kriedemann, P.E., Graham, R.D. and Wiskich, J.T. (1985). Photosynthetic dysfunction and in vivo changes in chlorophyll a fluorescence from manganese-deficient wheat leaves. Australian Journal of Agricultural Research, 36, 157–169.CrossRefGoogle Scholar
Krieger, A., Bolte, S., Dietz, K.J., et al. (1998). Thermoluminescence studies on the facultative CAM plant Mesembryanthenum crystallinum L. Planta, 205, 587–594.CrossRefGoogle Scholar
Krishnan, P., Kruger, N.J. and Ratcliffe, R.G. (2005). Metabolite fingerprinting and profiling using NMR. Journal of Experimental Botany, 56, 255–265.CrossRefGoogle ScholarPubMed
Krömer, S. (1995). Respiration during photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology, 46, 45–70.CrossRefGoogle Scholar
Krömer, S. and Heldt, H.W. (1991). Respiration of pea leaf mitochondria and redox transfer between the mitochondrial and extramitochondrial compartment. Biochim Biophys Acta, 1057, 42–50.CrossRefGoogle Scholar
Krömer, S., Malmberg, G., and Gardeström, P. (1993). Mitochondrial contribution to photosynthetic metabolism. Plant Physiology, 102, 947–955.Google ScholarPubMed
Kroopnick, P. and Craig, H. (1972). Atmospheric oxygen: isotopic composition and solubility fractionation. Science, 175, 54–55.CrossRefGoogle ScholarPubMed
Kruckeberg, A.L., Neuhaus, H.E., Feil, R., et al. (1989). Decreased-activity mutants of phosphoglucose isomerase in the cytosol and chloroplast of Clarkia xantiana. Impact on mass-action ratios and fluxes to sucrose and starch, and estimation of flux control coefficients and elasticity coefficients. Biochemical Journal, 261, 457–467.CrossRefGoogle ScholarPubMed
Krupa, S., Nosal, M. and Legge, A. (1998). A numerical analysis of the combined open-top chamber data from the USA and Europe on ambient ozone and negative crop responses. Environmental Pollution, 101, 157–160.CrossRefGoogle ScholarPubMed
Krupa, S.V. (2003). Effects of atmospheric ammonia (NH3) on terrestrial vegetation: a review. Environmental Pollution, 124, 179–221.CrossRefGoogle ScholarPubMed
Krupa, Z., Öquist, G. and Huner, N.P.A. (1993). The effect of cadmium on photosynthesis of Phaseolus vulgaris – a fluorescence analysis. Physiologia Plantarum, 88, 626–630.CrossRefGoogle Scholar
Kruse, A., Fieuw, S., Heinke, D., et al. (1998). Antisense inhibition of cytosolic NADP-isocitrate dehydrogenase in transgenic potato plants. Planta, 205, 82–91.CrossRefGoogle Scholar
Kubien, D.S. and Sage, R.F. (2004). Low-temperature photosynthetic performance of a C-4 grass and a co-occurring C-3 grass native to high latitudes. Plant Cell Environment, 27, 907–916.CrossRefGoogle Scholar
Kubiske, M.E., Quinn, V.S., Marquardt, P.E., et al. (2006). Effects of elevated atmospheric CO2 and/or O3 on intra- and interspecific competitive ability of aspen. Plant Biology, 9, 342–355.CrossRefGoogle Scholar
Kucharik, C.J., Norman, J.M. and Gower, S.T. (1999). Characterization of radiation regimes in nonrandom forest canopies: theory, measurements, and a simplified modeling approach. Tree Physiology, 19, 695–706.CrossRefGoogle Scholar
Külheim, C., Ågren, J. and Jansson, S. (2002). Rapid regulation of light harvesting and plant fitness in the field. Science, 297, 91–93.CrossRefGoogle ScholarPubMed
Kull, O. and Jarvis, P.G. (1995). The role of nitrogen in a simple scheme to scale up photosynthesis from leaf to canopy. Plant Cell Environment, 18, 1174–1182.CrossRefGoogle Scholar
Kull, O. and Niinemets, U. (1998). Distribution of leaf photosynthetic properties in tree canopies: comparison of species with different shade tolerance. Functional Ecology, 12, 472–479.CrossRefGoogle Scholar
Kumagai, T., Tateishi, M., Shimizu, T., et al. (2008). Transpiration and canopy conductance at two slope positions in a Japanese cedar forest watershed. Agricultural and Forest Meteorology, 148, 1444–1455.CrossRefGoogle Scholar
Kumar, P., Kumar Tewari, R. and Sharma, P.N. (2008). Modulation of copper toxicity induced oxidative damage by excess supply of iron in maize plants. Plant Cell Reports, 27, 399–409.CrossRefGoogle ScholarPubMed
Kume, A., Tsuboi, N., Satomura, T., et al. (2000). Physiological characteristics of Japanese red pine, Pinus densiflora Sieb. et Zucc., in declined forests at Mt. Gokurakuji in Hiroshima Prefecture, Japan. Trees, 141, 305–311.Google Scholar
Kump, L.R. (2008). The rise of atmospheric oxygen. Nature, 451, 277–278.CrossRefGoogle ScholarPubMed
Kumudini, S., Hume, D.J. and Chu, G. (2001). Genetic improvement in short season soybeans: 1. Dry matter accumulation, partitioning, and leaf area duration. Crop Science, 41, 391–398.CrossRefGoogle Scholar
Kumutha, D., Ezhilmathi, K., Sairam, R.K., et al. (2009). Waterlogging induced oxidative stress and antioxidant activity in pigeonpea genotypes. Biologia Plantarum, 53, 75–84.CrossRefGoogle Scholar
Kunkel, K.E., Angel, J.R., Changnon, S.A., et al. (2006). The 2005 Illinois Drought, Illinois State Water Survey, Champaign, IL.Google Scholar
Küpper, H., Setlík, I., Trtílek, M., et al. (2000). A microscope for two-dimensional measurements of in vivo chlorophyll fluorescence kinetics using pulsed measuring radiation, continuous actinic radiation, and saturating flashes. Photosynthetica, 38, 553–570.CrossRefGoogle Scholar
Kurepa, J., Bueno, P., Kampfenkel, K., et al. (1997). Effects of iron deficiency on iron superoxide dismutase expression in Nicotiana tabacum. Plant Physiology and Biochemistry, 35, 467–474.Google Scholar
Kurimoto, K., Millar, A.H., Lambers, H., et al. (2004). Maintenance of growth rate at low temperature in rice and wheat cultivars with a high degree of respiratory homeostasis is associated with a high efficiency of respiratory ATP production. Plant and Cell Physiology, 45, 1015–1022.CrossRefGoogle ScholarPubMed
Kursar, T.A. and Coley, P.D. (1992a). The consequences of delayed greening during leaf development for light absorption and light use efficiency. Plant Cell Environment, 15, 901–909.CrossRefGoogle Scholar
Kursar, T.A. and Coley, P.D. (1992b). Delayed development of the photosynthetic apparatus in tropical rain forest species. Functional Ecology, 6, 411–422.CrossRefGoogle Scholar
Kursar, T.A. and Coley, P.D. (1992c). Delayed greening in tropical leaves: an antiherbivore defense?Biotropica, 24, 256–262.CrossRefGoogle Scholar
Kyparissis, A. and Manetas, Y. (1993). Seasonal leaf dimorphism in a semi-deciduous Mediterranean shrub: ecophysiological comparisons between winter and summer leaves. Acta Oecologica, 14, 23–32.Google Scholar
Kyparissis, A., Drilias, P. and Manetas, Y. (2000). Seasonal fluctuations in photoprotective (xanthophyll cycle) and photoselective (chlorophylls) capacity in eight Mediterranean plant species belonging to two different growth forms. Australian Journal of Plant Physiology, 27, 265–272.Google Scholar
Kyparissis, A., Petropoulou, Y. and Manetas, Y. (1995). Summer survival of leaves in a soft-leaved shrub (Phlomis fruticosa L., Labiatae) under Mediterranean field conditions: avoidance of photoinhibitory damage through decreased chlorophyll contents. Journal of Experimental Botany, 46, 1825–1831.CrossRefGoogle Scholar
Labrecque, M., Bellefleur, P., Simon, J.P., et al. (1989). Influence of light conditions on the predetermination of foliar characteristics in Betula alleghaniensis Britton. Annals of Forest Science, 46, 497–501.CrossRefGoogle Scholar
Laidler, K.J. and King, M.C. (1983). Development of transition-state theory. The Journal of Physical Chemistry, 87, 2657–2664.CrossRefGoogle Scholar
Laisk, A. (1983). Calculation of photosynthetic parameters considering the statistical distribution of stomatal apertures. Journal of Experimental Botany, 34, 1627–1635.CrossRefGoogle Scholar
Laisk, A. and Loreto, F. (1996). Determining photosynthetic parameters from leaf CO2 exchange and chlorophyll fluorescence. Ribulose-1,5-bisphosphate carboxylase/oxygenase specificity factor, dark respiration in the light, excitation distribution between photosystems, alternative electron transport rate, and mesophyll diffusion resistance. Plant Physiology, 110, 903–912.CrossRefGoogle ScholarPubMed
Laisk, A. and Oja, V. (1998). Dynamics of Leaf Photosynthesis: Rapid-Response Measurements and their Interpretations. CSIRO Publishing, Canberra, Australia.Google Scholar
Laisk, A., Kull, O. and Moldau, H. (1989). Ozone concentration in leaf intercellular air spaces is close to zero. Plant Physiology, 90, 1163–1167.CrossRefGoogle ScholarPubMed
Laisk, A., Oja, V. and Kull, K. (1980). Statistical distribution of stomatal apertures of Vicia faba and Hordeum vulgare and Spannungsphase of stomatal opening. Journal of Experimental Botany, 31, 49–58.CrossRefGoogle Scholar
Laisk, A., Oja, V., Rasulov, B., et al. (2002). A computer-operated routine of gas exchange and optical measurements to diagnose photosynthetic apparatus in leaves. Plant, Cell and Environment, 25, 923–943.CrossRefGoogle Scholar
Laisk, A.K. (1977). Kinetics of photosynthesis and photorespiration in C3-plants. [In Russian] Nauka, Moscow.
Lake, J.A. and Wade, R.N. (2009) Plant-pathogen interactions and elevated CO2: morphological changes in favour of pathogens. Journal of Experimental Botany, 60, 3123–3131.CrossRefGoogle ScholarPubMed
Lal, R. (2004a) Agricultural activities and the global carbon cycle. Nutrient Cycling in Agroecosystems, 70, 103–116.CrossRefGoogle Scholar
Lal, R. (2004b). Soil carbon sequestration impacts on global climate change and food security. Science, 304, 1623–1627.CrossRefGoogle ScholarPubMed
Lal, R. (2004c). Soil carbon sequestration to mitigate climate change. Geoderma, 123, 1–22.CrossRefGoogle Scholar
Lal, R. and Edwards, G.E. (1996). Analysis of inhibition of photosynthesis under water stress in the C4 species Amaranthus cruentus and Zea mays: electron transport, CO2 fixation and carboxylation capacity. Functional Plant Biology, 23, 403–412.Google Scholar
Laloi, C., Mestres-Ortega, D., Marco, Y., et al. (2004). The Arabidopsis cytosolic thioredoxin h5 gene induction by oxidative stress and its W-box-mediated response to pathogen elicitor. Plant Physiology, 134, 1006–1016.CrossRefGoogle ScholarPubMed
Lambers, H. and Poorter, H. (1992). Inherent variation in growth rate between higher plants: a search for physiological causes and ecological consequences. In: Advances in Ecological Research (eds Begon, M. and Fitter, A.H.), Academic Press, London, UK, pp. 187–261.Google Scholar
Lambers, H., Chapin III, F.S. and Pons, T.L. (2008). Plant Physiological Ecology, 2nd ed. Springer Science + Business Media, LLC, New York, USA, pp. 604.CrossRefGoogle Scholar
Lambert, G., Monfray, P., Ardouin, B., et al. (1995). Year-to-year changes in atmo-spheric CO2. Tellus, 47B, 53–55.CrossRefGoogle Scholar
Lamont, B.B., Groom, P.K. and Cowling, R.M. (2002). High leaf mass per area of related species assemblages may reflect low rainfall and carbon isotope discrimination rather than low phosphorus and nitrogen concentrations. Functional Ecology, 16, 403–412.CrossRefGoogle Scholar
Landlot, W., Bühlmann, U., Bleuler, P., et al. (2000). Ozone exposure-response relationships for biomass and root/shoot ratio of beech (Fagus sylvatica), ash (Fraxinus excelsior), Norway spruce (Picea abies) and Scots pine (Pinus sylvestris). Environmental Pollution, 109, 473–478.CrossRefGoogle Scholar
Landsberg, J.J., Prince, S.D., Jarvis, P.G., et al. (1996). Energy conversion and use in forests: an analysis of forest production in terms of radiation utilisation efficiency (epsilon). In: The Use of Remote Sensing in the Modeling of Forest Productivity (eds Gholz, H.L., Nakane, K. and Shimoda, H.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 273–298.Google Scholar
Lane, G.A. and Dole, M. (1956). Fractionation of oxygen isotopes during respiration. Science, 123, 574–576.CrossRefGoogle ScholarPubMed
Lane, M.D., Maruyama, H. and Easterday, R.L. (1969). Phosphoenolpyruvate carboxylase from peanut cotyledons. Methods in Enzymology, 13, 277–283.CrossRefGoogle Scholar
Lange, O.L. (1969). CO2-Gaswechsel von Moosen nach Wasserdamnfaufnahme aus dem Luftraum. Planta, 89, 90–94.CrossRefGoogle 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., Schulze, E.D., Evenari, M., et al. (1974). The temperature-related photosynthetic capacity of plants under desert conditions I. Seasonal changes of the photosynthetic response to temperature. Oecologia, 17, 97–110.CrossRefGoogle ScholarPubMed
Lange, O.L., Schulze, E.D., Kappen, L., et al. (1975). CO2 exchange pattern under natural conditions of Caralluma negevensis, a CAM plant of the Negev desert. Photosynthetica, 9, 318–326.Google Scholar
Langenheim, J.H., Osmond, C.B., Brooks, A., et al. (1984). Photosynthetic responses to light in seedlings of selected Amazonian and Australian rainforest tree species. Oecologia, 63, 215–224.CrossRefGoogle ScholarPubMed
Lanigan, G., Betson, N., Griffiths, H., et al. (2008). Carbon isotope fractionation during photorespiration and carboxylation in Senecio. Plant Physiology, 148, 2013–2020.CrossRefGoogle ScholarPubMed
Larbi, A., Abadía, A., Abadía, J., et al. (2006). Down co-regulation of light absorption, photochemistry, and carboxylation in Fe-deficient plants growing in different environments. Photosynthesis Research, 89, 113–126.CrossRefGoogle ScholarPubMed
Larbi, A., Morales, F., Abadía, A., et al. (2002). Effects of Cd and Pb in sugar beet plants grown in nutrient solution: induced Fe deficiency and growth inhibition. Functional Plant Biology, 29, 1453–1464.CrossRefGoogle Scholar
Larcher, W., Wagner, J. and Lütz, C. (1997). The effect of heat on photosynthesis, dark respiration and cellular ultrastructure of the arctic-alpine psychrophyte Ranunculus glacialis. Photosynthetica, 34, 219–232.CrossRefGoogle Scholar
Larom, S., Salama, F., Schuster, G. et al. (2010). Engineering of an alternative electron transfer path in photosystem II. Proceedings of the National Academy of Sciences USA, 107, 9650–9655.CrossRefGoogle ScholarPubMed
Lasslop, G., Reichstein, M., Kattge, J., et al. (2008). Influences of observation errors in eddy flux data on inverse model parameter estimation. Biogeosciences, 5, 1311–1324.CrossRefGoogle Scholar
Lasslop, G., Reichstein, M., Papale, D., et al. (2010). Separation of net ecosystem exchange into assimilation and respiration using a light response curve approach: critical issues and global evaluation. Biogeosciences, 16, 187–208.Google Scholar
Lassoie, J.P., Dougherty, P.M., Reich, P.B., et al. (1983). Ecophysiological investigations of understory eastern redcedar in Central Missouri. Ecology, 64, 1355–1366.CrossRefGoogle Scholar
Latham, R.E. (1992). Co-occurring tree species change rank in seedling performance with resources varied experimentally. Ecology, 73, 2129–2144.CrossRefGoogle Scholar
Latorre, C., Quade, J. and McIntosh, W.C. (1997). The expansion of C4 grasses and global change in the late Miocene: stable isotope evidence from the Americas. Earth Planetary Science Letters, 146, 83–96.CrossRefGoogle Scholar
Latowski, D., Kruk, J. and Strzalka, K. (2005). Inhibition of zeaxanthin epoxidase activity by cadmium ions in higher plants. Journal of Inorganic Biochemistry, 99, 2081–2087.CrossRefGoogle ScholarPubMed
Lauer, M.J. and Boyer, J.S. (1992). Internal CO2 measured directly in leaves. Abscisic acid and low leaf water potential cause opposing effects. Plant Physiology, 98, 1310–1316.CrossRefGoogle Scholar
Lauer, M.J., Pallardy, S.G., Blevins, D.G., et al. (1989). Whole leaf carbon exchange characteristics of phosphate deficient soybeans (Glycine max L.). Plant Physiology, 91, 848–854.CrossRefGoogle Scholar
Lauteri, M., Scartazza, A., Guido, C., et al. (1997). Genetic variation in photosynthetic capacity, carbon isotope discrimination and mesophyll conductance in provenances of Castanea sativa adapted to different environments. Functional Ecology, 11, 675–683.CrossRefGoogle Scholar
Lavergne, J. and Junge, W. (1992). Proton release during the redox cycle of the water oxidase. Photosynthesis Research, 38, 279–296.CrossRefGoogle Scholar
Law, B.E., Falge, E., Gu, L., et al. (2002). Environmental controls over carbon dioxide and water vapor exchange of terrestrial vegetation. Agricultural and Forest Meteorology, 113, 97–120.CrossRefGoogle Scholar
Lawlor, D. (2001). Photosynthesis. BIOS Scientific Publishers, Oxford, UK, pp. 386.
Lawlor, D.W. (1995). Photosynthesis, productivity and environment. Journal of Experimental Botany, 46, 1449–1461.CrossRefGoogle Scholar
Lawlor, D.W. (2002). Carbon and nitrogen assimilation in relation to yield: mechanisms are the key to understanding production systems. Journal of Experimental Botany, 53, 773–787.CrossRefGoogle ScholarPubMed
Lawlor, D.W. and Cornic, G. (2002). Photosynthetic carbon assimilation and associated metabolism in relation to water deficits in higher plants. Plant, Cell and Environment, 25, 275–294.CrossRefGoogle ScholarPubMed
Lawlor, D.W. and Tezara, W. (2009). Causes of decreased photosynthetic rate and metabolic capacity in water-deficient leaf cells: a critical evaluation of mechanisms and integration of processes. Annals of Botany, 103, 561–579.CrossRefGoogle Scholar
Lawson, T., Lefebvre, S., Baker, N.R., et al. (2008). Reductions in mesophyll and guard cell photosynthesis impact on the control of stomatal responses to light and CO2. Journal of Experimental Botany, 59, 3609–3619.CrossRefGoogle ScholarPubMed
Lawson, T., Oxborough, K., Morison, J.I.L., et al. (2002). Responses of photosynthetic electron transport in stomatal guard cells and mesophyll cells in intact leaves to light, CO2 and humidity. Plant Physiology, 128, 52–62.CrossRefGoogle ScholarPubMed
Lawson, T., Oxborough, K., Morison, J.I.L., et al. (2003). The responses of guard and mesophyll cell photosynthesis to CO2, O2, light and water stress in a range of species are similar. Journal of Experimental Botany, 54, 1743–1752.CrossRefGoogle Scholar
Lawyer, A.L., Cornwell, K.L., Larsen, P.O., et al. (1981). Effects of carbon dioxide and oxygen on the regulation of photosynthetic carbon metabolism by ammonia in spinach mesophyll cells. Plant Physiology, 68, 1231–1236.CrossRefGoogle ScholarPubMed
Le Roux, X., Grand, S., Dreyer, E., et al. (1999). Parameterization and testing of a biochemically based photosynthesis model for walnut (Juglans regia) trees and seedlings. Tree Physiology, 19, 481–492.CrossRefGoogle ScholarPubMed
Leadley, P. and Drake, B. (1993). Open top chambers for exposing plant canopies to elevated CO2 concentration and for measuring net gas exchange. CO2 and Biosphere, 104, 3–15.CrossRefGoogle Scholar
Leakey, A.D.B., Ainsworth, E.A., Bernacchi, C.J., et al. (2009b). Elevated CO2 effects on plant carbon, nitrogen, and water relations: six important lessons from FACE. Journal of Experimental Botany, 60, 2859–2876.CrossRefGoogle ScholarPubMed
Leakey, A.D.B., Bernacchi, C.J., Dohleman, F.G., et al. (2004). Will photosynthesis of maize (Zea mays) in the U.S. Corn Belt increase in future [CO2] rich atmospheres? An analysis of diurnal courses of CO2 uptake under Free-Air Concentration Enrichment (FACE). Global Change Biology, 10, 951–962.CrossRefGoogle Scholar
Leakey, A.D.B., Uribelarrea, M., Ainsworth, E.A., et al. (2006). Photosynthesis, productivity, and yield of maize are not affected by open-air elevation of CO2 concentration in the absence of drought. Plant Physiology, 140, 779–790.CrossRefGoogle Scholar
Leakey, A.D.B., Xu, F., Gillespie, K.M., et al.. (2009a). The genomic basis for stimulated respiratory carbon loss to the atmosphere by plants growing under elevated [CO2]. Proceedings of the National Academy of Sciences, USA, 106, 3597–3602.CrossRefGoogle Scholar
Lebaube, S., Le Goff, N., Ottorini, J.M., et al. (2000). Carbon balance and tree growth in a Fagus sylvatica stand. Annals of Forest Science, 57, 49–61.CrossRefGoogle Scholar
LeCain, D.R., Morgan, J.A., Mosier, A.R., et al. (2003). Soil and plant water relations, not photosynthetic pathway, primarily influence photosynthetic responses in a semi-arid ecosystem under elevated CO2. Annals of Botany, 92, 41–52.CrossRefGoogle ScholarPubMed
LeCain, D.R., Morgan, J.A., Schuman, G.E., et al. (2000). Carbon exchange of grazed and ungrazed pastures of a mixed grass prairie. Journal of Range Management, 53, 199–206.CrossRefGoogle Scholar
LeCain, D.R., Morgan, J.A., Schuman, G.E., et al. (2002). Carbon exchange and species composition of grazed pastures and exclosures in the shortgrass steppe of Colorado. Agriculture, Ecosystems and Environment, 93, 421–435.CrossRefGoogle Scholar
Ledford, H.K. and Niyogi, K.K. (2005). Singlet oxygen and photo-oxidative stress management in plants and algae. Plant, Cell Environment, 28, 1037–1045.CrossRefGoogle Scholar
Lee, D.W. (1987). The spectral distribution of radiation in two neotropical rainforests. Biotropica, 19, 161–166.CrossRefGoogle Scholar
Lee, D.W. and Collins, T.M. (2001). Phylogenetic and ontogenetic influences on the distribution of anthocyanins and betacyanins in leaves of tropical plants. International Journal of Plant Sciences, 162, 1141–1153.CrossRefGoogle Scholar
Lee, D.W., Lowry, J.B. and Stone, B.C. (1979). Abaxial anthocyanin layer in leaves of tropical rainforest plants: enhancer of light capture in deep shade. Biotropica, 11, 70–77.CrossRefGoogle Scholar
Lee, D.W., O’Keefe, J., Holbrook, N.M., et al. (2003). Pigment dynamics and autumn leaf senescence in a New England deciduous forest, eastern USA. Ecological Research, 18, 677–694.CrossRefGoogle Scholar
Lee, H.S.J., Lüttge, U., Medina, E., et al. (1989). Ecophysiology of xerophytic and halophytic vegetation of a coastal alluvial plain in northern Venezuela. III. Bromelia humilis Jacq. a terrestrial CAM bromeliad. New Phytologist, 111, 253–271.CrossRefGoogle Scholar
Lee, H.Y., Hong, Y.N. and Chow, W.S. (2001a). Photoinactivation of photosystem II complexes and photoprotection by non-functional neighbours in Capsicum annuum L. leaves. Planta, 212, 332–342.CrossRefGoogle ScholarPubMed
Lee, J.H. and Schöffl, F. (1996). An Hsp70 antisense gene affects the expression of HSP70/HSC70, the regulation of HSF, and the acquisition of thermotolerance in transgenic Arabidopsis thaliana. Molecular and General Genetics, 252, 11–19.Google ScholarPubMed
Lee, K.C., Cunningham, B.A., Paulsen, G.M., et al. (1976). Effects of cadmium on respiration rate and activities of several enzymes in soybean seedlings. Physiologia Plantarum, 36, 4–6.CrossRefGoogle Scholar
Lee, K.P., Kim, C., Landgraf, F., et al. (2007). EXECUTER1- and EXECUTER2- dependent transfer of stress-related signals from the plastid to the nucleus of Arabidopsis thaliana. Proceedings of the National Academy of Sciences USA, 104, 10270–10275.CrossRefGoogle ScholarPubMed
Lee, R.H., Wang, C.H., Huang, L.T., et al. (2001b). Leaf senescence in rice plants: cloning and characterization of senescence up-regulated genes. Journal of Experimental Botany, 52, 1117–1121.CrossRefGoogle ScholarPubMed
Lee, T.D., Tjoelker, M.G., Ellsworth, D.S., et al. (2001c). Leaf gas exchange responses of 13 prairie grassland species to elevated CO2 and increased nitrogen supply. New Phytologist, 150, 405–418.CrossRefGoogle Scholar
Leegood, R.C. (1990). Enzymes of the Calvin cycle. In: Enzymes of Primary Metabolism, Methods in Plant Biochemistry 3 (eds Lea, P.J. and Harborne, J.B.), Academic Press, London. pp. 15–37.Google Scholar
Leegood, R.C. (1993). Carbon metabolism. In: Photosynthesis and Production in a Changing Environment. A Field Laboratory Manual (eds Hall, D.O., Scurlock, J.M.O., Bolhar-Nordenkampf, J.R., Leegood, R.C. and Long, S.P.), Chapman and Hall, London, UK.Google Scholar
Leegood, R.C. (2002). C4 photosynthesis: principles of CO2 concentration and prospects for its introduction into C3 plants. Journal of Experimental Botany, 53, 581–590.CrossRefGoogle Scholar
Leegood, R.C. (2008). C4 photosynthesis: minor or major adjustments to a C3 theme? In: Charting New Pathways to C4 in Rice (eds Sheehy, J.E., Mitchel, P.L. and Hardy, B.) World Scientific Pub, Hackensack, USA.Google Scholar
Leegood, R.C. and Edwards, G.E. (1996). Carbon metabolism and photorespiration: temperature dependence in relation to other environmental factors. In: Advances in Photosynthesis Vol.5: Photosynthesis and the Environment (ed. Baker, N.R.) Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 191–221.Google Scholar
Lefi, E., Medrano, H. and Cifre, J. (2004). Water uptake dynamics, photosynthesis and water use efficiency in field-grown Medicago arborea and Medicago citrina under prolonged Mediterranean drought conditions. Annals of Applied Biology, 144, 299–307.CrossRefGoogle Scholar
Legge, A.H. and Krupa, S.V. (2002). Effects of sulphur dioxide. In: Air Pollution and Plant Life, 2nd edn (eds Bell, J.N.B. and Treshow, M.), John Wiley and Sons, Chichester, USA, pp. 135–162.Google Scholar
Legge, A.H., Jäger, H.J. and Krupa, S.V. (1998). Sulfur dioxide. In: Recognition of Air Pollution Injury to Vegetation: A Pictorial Atlas, 2nd ed (ed. Flagler, R.B.), Air and Waste Management Association, Pittsburgh, USA, pp. 3–42.Google Scholar
Lehto, K., Tikkanen, M., Hiriart, J.B., et al. (2003). Depletion of the photosystem II core complex in mature tobacco leaves infected by the Flavum strain of Tobacco mosaic virus. Molecular Plant-Microbe Interactions, 16, 1135–1144.CrossRefGoogle ScholarPubMed
Leigh, R.A. and Wyn Jones, R.G. (1984). A hypothesis relating critical potassium concentrations for growth to the distribution and functions of this ion in the plant cell. New Phytologist, 97, 1–13.CrossRefGoogle Scholar
Leinonen, I. and Jones, H.G. (2004). Combining thermal and visible imagery for estimating canopy temperature and identifying plant stress. Journal of Experimental Botany, 55, 1423–1431.CrossRefGoogle ScholarPubMed
Leinonen, I., Grant, O.M., Tagliavia, C.P.P., et al. (2006). Estimating stomatal conductance with thermal imagery. Plant, Cell and Environment, 29, 1508–1518.CrossRefGoogle ScholarPubMed
Leipner, J., Basilides, A., Stamp, P., et al. (2000). Hardly increased oxidative stress after exposure to low temperature in chilling-acclimated and non-acclimated maize leaves. Plant Biology, 2, 243–251.CrossRefGoogle Scholar
Leith, J.H. and Reynolds, J.F. (1987). The nonrectangular hyperbola as a photosynthetic light response model: geometrical interpretation and estimation of the parameter. Photosynthetica, 21, 363–366.Google Scholar
Lemaire, S.D., Michelet, L., Zaffagnini, M., et al. (2007). Thioredoxins in chloroplasts. Current Genetics, 51, 343–365.CrossRefGoogle ScholarPubMed
Lemaître, T., Urbanczyk-Wochniak, E., Flesch, V., et al. (2007). NAD-dependent isocitrate dehydrogenase mutants of Arabidopsis suggest the enzyme is not limiting for nitrogen assimilation. Plant Physiology, 144, 1546–1558.CrossRefGoogle Scholar
Lenk, S. and Buschmann, C. (2006). Distribution of UV-shielding of the epidermis of sun and shade leaves of the beech (Fagus sylvatica L.) as monitored by multi-colour fluorescence imaging. Journal of Plant Physiology, 163, 1273–1283.CrossRefGoogle Scholar
Lenk, S., Chaerle, L., Pfündel, E.E., et al. (2007). Multispectral fluorescence and reflectance imaging at the leaf level and its possible applications. Journal of Experimental Botany, 58, 807–814.CrossRefGoogle ScholarPubMed
León, A.M., Palma, J.M., Corpas, F J., et al. (2002). Antioxidative enzymes in cultivars of pepper plants with different sensitivity to cadmium. Plant Physiology and Biochemistry, 40, 813–820.CrossRefGoogle Scholar
Leonardos, E.D., Savitch, L.V., Huner, N.P.A., et al. (2003). Daily photosynthetic and C-export patterns in winter wheat leaves during cold stress and acclimation. Physiologia Plantarum, 117, 521–531.CrossRefGoogle ScholarPubMed
Lerdau, M. (2007). A positive feedback with negative consequences. Science, 316, 212–213.CrossRefGoogle ScholarPubMed
Lerdau, M. and Slobodkin, K. (2002). Trace gas emissions and species-dependent ecosystem services. Trends in Ecology and Evolution, 17, 309–312.CrossRefGoogle Scholar
Leroux, X. and Mordelet, P. (1995). Leaf and canopy CO2 assimilation in a West-African humid savanna during the early growing season. Journal of Tropical Ecology, 11, 529–545.CrossRefGoogle Scholar
Leshem, Y., Melamed-Book, , Cagnac, O., Ronen, G., et al. (2006). Suppression of Arabidopsis vesicle-SNARE expression inhibited fusion of H2O2-containing vesicles with tonoplast and increased salt tolerance. Proceedings of the National Academy of Sciences USA, 103, 18008–18013.CrossRefGoogle ScholarPubMed
Letts, M.G., Phelan, C.A., Johnson, D.R.E., et al. (2008). Seasonal photosynthetic gas exchange and leaf reflectance characteristics of male and female cottonwoods in a riparian woodland. Tree Physiology, 28, 1037–1048.CrossRefGoogle Scholar
Leuning, R. (1990). Modelling stomatal behaviour and photosynthesis of Eucalyptus grandis. Australian Journal of Plant Physiology, 17, 159–175.CrossRefGoogle Scholar
Leuning, R. (1995). Critical-appraisal of a combined stomatal-photosynthesis model for C3 plants. Plant, Cell and Environment, 18, 339–355.CrossRefGoogle Scholar
Leuning, R. (2002). Temperature dependence of two parameters in a photosynthesis model. Plant, Cell and Environment, 25, 1205–1210.CrossRefGoogle Scholar
Leuning, R. and Judd, M.J. (1996). The relative merits of open- and closed-path analysers for measurements of eddy fluxes. Global Change Biology, 2, 241–254.CrossRefGoogle Scholar
Leuning, R., Cleugh, H.A., Zegelin, S.J., et al. (2005). Carbon and water fluxes over a temperate Eucalyptus forest and a tropical wet/dry savanna in Australia: measurements and comparison with MODIS remote sensing estimates. Agricultural and Forest Meteorology, 129, 151–173CrossRefGoogle Scholar
Leuning, R., Dunin, F. and Wang, Y. (1998). A two-leaf model for canopy conductance, photosynthesis and partitioning of available energy. II. Comparison with measurements. Agricultural and Forest Meteorology, 91, 113–125.CrossRefGoogle Scholar
Leuning, R., Kelliher, F.M., De Pury, D.G.G., et al. (1995). Leaf nitrogen photosynthesis, conductance and transpiration-scaling from leaves to canopies. Plant, Cell and Environment, 18, 1183–1200.CrossRefGoogle Scholar
Leverenz, J.W. (1988). The effects of illumination sequence, CO2 concentration, temperature and acclimation on the convexity of the photosynthetic light response curve. Physiologia Plantarum, 74, 332–341.CrossRefGoogle Scholar
Leverenz, J.W. (1995). Shade shoot structure of conifers and the photosynthetic response to light at two CO2 partial pressures. Functional Ecology, 9, 413–421.CrossRefGoogle Scholar
Leverenz, J.W. and Hinckley, T.M. (1990). Shoot structure, leaf-area index and productivity of evergreen conifer stands. Tree Physiology, 6, 135–149.CrossRefGoogle ScholarPubMed
Leverenz, J.W., Bruhn, D. and Saxe, H. (1999). Responses of two provenances of Fagus sylvatica seedlings to a combination of four temperature and two CO2 treatments during their first growing season: gas exchange of leaves and roots. New Phytologist, 144, 437–454.CrossRefGoogle Scholar
Levi, A., Ovnat, L., Paterson, A.H., et al. (2009). Photosynthesis of cotton near-isogenic lines introgressed with QTLs for productivity and drought related traits. Plant Science, 177, 88–96.CrossRefGoogle Scholar
Levin, D.A. (1973). The role of trichomes in plant defense. The Quarterly Review of Biology, 48, 3–15.CrossRefGoogle Scholar
Levine, A. (1999). Oxidative stress as a regulator of environmental responses in plants. In: Plant Responses to Environmental Stresses: From Phytohormones to Genome Reorganization (ed. Lerner, H.R.), CRC Press, Florida, USA, pp. 248–266.Google Scholar
Levitt, J. (1972). Responses of Plants to Environmental Stresses, Academic Press, New York, USA.
Levitt, J. (1980). Responses of Plants to Environmental Stresses, Adademic Press, New York, USA.
Lewis, C.E., Noctor, G., Causton, D., et al. (2000). Regulation of assimilate partitioning in leaves. Australian Journal of Plant Physiology, 27, 507–519.Google Scholar
Lewis, D.A. and Nobel, P.S. (1977). Thermal energy exchange model and water loss of a barrel cactus, Ferocactus acanthodes. Plant Physiology, 60, 609–616.CrossRefGoogle ScholarPubMed
Lewis, J.D., Griffin, K.L., Thomas, R.B., et al. (1994). Phosphorus supply affects the photosynthetic capacity of loblolly pine grown in elevated carbon-dioxide. Tree Physiology, 14, 1229–1244.CrossRefGoogle ScholarPubMed
Lewis, S.L., Lopez-Gonzalez, G., Sonké, B., et al. (2009). Increasing carbon storage in intact African tropical forests. Nature, 457, 1003–1006.CrossRefGoogle ScholarPubMed
Leyman, B., Geelen, D., Quintero, F.J., et al. (1999). A tobacco syntaxin with a role in hormonal control of guard cell ion channels. Science, 283, 537–540.CrossRefGoogle ScholarPubMed
Li, C., Liu, S. and Berninger, F. (2004). Picea seedlings show apparent acclimation to drought with increasing altitude in the eastern Himalaya. Trees, 18, 277–283.CrossRefGoogle Scholar
Li, C., Potuschak, T., Colón-Carmona, A., et al. (2005). Arabidopsis TCP20 links regulation of growth and cell division control pathways. Proceedings of the National Academy of Sciences USA, 102, 12978–12983.CrossRefGoogle ScholarPubMed
Li, S., Goodwin, S. and Pezeshki, S.R. (2007). Photosynthetic gene expression in black willow under various soil moisture regimes. Biologia Plantarum, 51, 593–596.CrossRefGoogle Scholar
Li, X., Gilmore, A., Caffarri, S., et al. (2004). Regulation of photosynthetic light harvesting involves intrathylakoid lumen pH sensing by the PsbS protein. Journal of Biological Chemistry, 279, 22866–22874.CrossRefGoogle ScholarPubMed
Li, X.P., Björkman, O., Shih, C., et al. (2000). A pigment-binding protein essential for regulation of photosynthetic light harvesting. Nature, 403, 391–395.CrossRefGoogle ScholarPubMed
Li, Y., Gao, Y., Xu, X., et al. (2009). Light-saturated photosynthetic rate in high-nitrogen rice (Oryza sativa L.) leaves is related to chloroplastic CO2 concentration. Journal of Experimental Botany, 6, 2351–2360.CrossRefGoogle Scholar
Li, Z., Zhang, S., Hu, H., et al. (2008). Photosynthetic performance along a light gradient as related to leaf characteristics of a naturally occurring Cypripedium flavum. Journal of Plant Research, 121, 559–569.CrossRefGoogle ScholarPubMed
Liang, J.S. and Zhang, J.H. (1999) The relations of stomatal closure and reopening to xylem ABA concentration and leaf water potential during soil drying and rewatering. Plant Growth Regulation, 29, 77–86.CrossRefGoogle Scholar
Liberloo, M., Tulva, I., Raïm, O., et al. (2007). Photosynthetic stimulation under long-term CO2 enrichment and fertilization is sustained across a closed Populus canopy profile (EUROFACE). New Phytologist, 173, 537–549.CrossRefGoogle Scholar
Lichtenthaler, H.K., Hak, R. and Rinderle, U. (1990). The chlorophyll fluorescence ratio F690/F730 in leaves of different chlorophyll content. Photosynthesis Research, 25, 295–298.CrossRefGoogle ScholarPubMed
Lichtenthaler, H.K. (1987). Chlorophylls and carotenoids: pigments of photosynthetic biomembranes. Methods in Enzymology, 148, 350–382.CrossRefGoogle Scholar
Lichtenthaler, H.K. (1999). The 1-deoxy D-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 47–65.CrossRefGoogle ScholarPubMed
Lichtenthaler, H.K. and Babani, F. (2000). Detection of photosynthetic activity and water stress by imaging the red chlorophyll fluorescence. Plant Physiology and Biochemistry, 38, 889–895.CrossRefGoogle Scholar
Lichtenthaler, H.K. and Miehé, J.A. (1997). Fluorescence imaging as a diagnostic tool for plant stress. Trends in Plant Sciences, 2, 316–320.CrossRefGoogle Scholar
Lichtenthaler, H.K., Buschmann, C., Rinderle, U., et al. (1986). Application of chlorophyll fluorescence in ecophysiology. Radiation and Environmental Biophysics, 25, 297–308.CrossRefGoogle ScholarPubMed
Lichtenthaler, H.K., Langsdorf, G., Lenk, S., et al. (2005). Chlorophyll fluorescence imaging of photosynthetic activity with the flash-lamp fluorescence imaging system. Photosynthetica, 43, 355–369.CrossRefGoogle Scholar
Lichtenthaler, H.K., Rohmer, M. and Schwender, J. (1997). Two independent biochemical pathways for isopentenyl diphosphate and isoprenoid biosynthesis in higher plants. Physiologia Plantarum, 101, 643–652.CrossRefGoogle Scholar
Lichtenthaler, H.K., Wenzel, O., Buschmann, C., et al. (1998). Plant stress detection by reflectance and fluorescence. Annals of the New York Academy of Sciences, 851, 271–285.CrossRefGoogle Scholar
Li-Cor Inc. (2001). Interfacing custom chambers to the LI-6400 sensor head. LI-6400 portable photosynthesis system: application note 3. Li-Cor, Inc, Lincoln, Nebraska, USA.
Li-Cor Inc. (2008). Using the 6400–17 whole plant Arabidopsis chamber. LI-6400 portable photosynthesis system: application note 4. Li-Cor, Inc, Lincoln, Nebraska, USA.
Lidon, F.C., Barreiro, M.G., Ramalho, J.C., et al. (1999). Effects of aluminum toxicity on nutrient accumulation in maize shoots: implications on photosynthesis. Journal of Plant Nutrition, 22, 397–416.CrossRefGoogle Scholar
Liebig, M., Scarano, F.R., Mattos de, E.A., et al. (2001). Ecophysiological and floristic implications of sex expression in the dioecious neotropical CAM tree Clusia hilariana Schltdl. Trees, 15, 278–288.CrossRefGoogle Scholar
Lieth, H. (1975). Modelling the primary productivity of the world. In: Primary Productivity of the Biosphere (ed. Lieth, H. and Whittaker, R.H.). Springer-Verlag, Berlin-Heidelberg-New York, pp. 237–263.
Lieth, H. and Whittaker, R.H. (1975). Primary productivity of the biosphere. In: Ecological Studies 14. Springer-Verlag, Berlin, Germany, pp. 339.CrossRef
Lim, P.O., Kim, H.J. and Nam, H.G. (2007). Leaf senescence. Annual Review of Plant Biology, 58, 115–136.CrossRefGoogle ScholarPubMed
Lim, P.O., Woo, H.R., Nam, H.G. (2003). Molecular genetics of leaf senescence in Arabidopsis. Trends Plant Sci, 8, 272–278.CrossRefGoogle ScholarPubMed
Limousin, J.M., Misson, L., Lavoir, A.V., et al. (2010). Do photosynthetic limitations of evergreen Quercus ilex leaves change with long-term increased drought severity?Plant, Cell and Environment, 33, 863–875.Google ScholarPubMed
Lin, C.T. (2000). Plant blue-light receptors. Trends in Plant Science, 5, 337–342.CrossRefGoogle ScholarPubMed
Lin, G. and Ehleringer, J.R. (1997). Carbon isotope fractionation does not occur during dark respiration in C3 and C4 plants. Plant Physiology, 114, 391–394.CrossRefGoogle Scholar
Lin, G. and Sternberg, L.D.L. (1992a). Effect of growth form, salinity, nutrient and sulfide on photosynthesis, carbon isotope discrimination and growth of red mangrove (Rhizophora mangle L.). Australian Journal of Plant Physiology, 19, 509–517.CrossRefGoogle Scholar
Lin, G. and Sternberg, L.D.L. (1992b). Differences in morphology, carbon isotope ratios, and photosynthesis between scrub and fringe mangroves in Florida, USA. Aquatic Botany, 42, 303–313.CrossRefGoogle Scholar
Lin, M., Turpin, D.H. and Plaxton, W.C. (1989). Pyruvate kinase isozymes from the green alga Selenastrum minutum. Kinetic and regulatory properties. Archives of Biochemistry and Biophysics, 269, 228–238.CrossRefGoogle ScholarPubMed
Lindenthal, M., Steiner, U., Dehne, H.W., et al. (2005). Effect of downy mildew development on transpiration of cucumber leaves visualized by digital infrared thermography. Phytopathology, 95, 233–240.CrossRefGoogle ScholarPubMed
Lindner, R.C., Kirpatrick, H.C. and Weeks, T.E. (1959). Some factors affecting the susceptibility of cucumber cotyledons to infection by tobacco mosaic virus. Phytopatology, 49, 78–88.Google Scholar
Lindquist, J.L., Arkebauer, T.J., Walters, D.T., et al. (2005). Maize radiation use efficiency under optimal growth conditions. Agronomy Journal, 97, 72–78.CrossRefGoogle Scholar
Lindquist, S. and Craig, E.A. (1988). The Heat-Shock Proteins. Annual Review of Genetics, 22, 631–677.CrossRefGoogle ScholarPubMed
Lindroth, A., Grelle, A. and Moren, A.S. (1998). Long-term measurements of boreal forest carbon balance reveal large temperature sensitivity. Global Change Biology, 4, 443–450.CrossRefGoogle Scholar
Littlejohn, R.O. and Ku, M.S.B. (1984). Characterization of early morning crassulacean acid metabolism in Opuntia erinacea var. Columbiana (Griffiths) L. Benson. Plant Physiology, 74, 1050–1054.CrossRefGoogle ScholarPubMed
Liu, C.C., Liu, Y.G., Guo, K., et al. (2010). Influence of drought on the response of six woody karst species subjected to successive cycles of drought and rewatering. Physiologia Plantarum, 139, 39–54.CrossRefGoogle ScholarPubMed
Liu, D., Li, T.Q., Yang, X.E., et al. (2008). Effect of Pb on leaf antioxidant enzyme activities and ultrastructure of the two ecotypes of Sedum alfredii Hance. Russian Journal of Plant Physiology, 55, 68–76.CrossRefGoogle Scholar
Liu, F. and Stützel, H. (2004). Biomass partitioning, specific leaf area, and water use efficiency of vegetable amaranth (Amaranthus spp.) in response to drought stress. Scientia Horticulturae, 102, 15–27.CrossRefGoogle Scholar
Liu, J., Yeo, H.C., Doniger, S.J., et al. (1997). Assay of aldehydes from lipid peroxidation: gas chromatography mass spectrometry compared to thiobarbituric acid. Analytical Biochemistry, 245, 161–166.CrossRefGoogle ScholarPubMed
Liu, L., Zhang, Y., Wang, J, et al. (2005). Detection solar-induced chlorophyll fluorescence from field radiance spectra based on the Fraunhofer line principle. IEEE transaction on Geoscience and Remote Sensing, 43, 827–832.Google Scholar
Liu, L.X., Xu, S.M. and Woo, K.C. (2003). Influence of leaf angle on photosynthesis and the xanthophyll cycle in the tropical tree species Acacia crassicarpa. Tree Physiology, 23, 1255–1261.CrossRefGoogle ScholarPubMed
Liu, M.Z. and Osborne, C.P. (2008). Leaf cold acclimation and freezing injury in C_3 and C_4 grasses of the Mongolian plateau. Journal of Experimental Botany, 59, 4161–4170.CrossRefGoogle Scholar
Liu, N., Peng, C.L., Lin, Z.F., et al. (2006). Changes in photosystem II activity and leaf reflectance features of several subtropical woody plants under simulated SO2 treatment. Journal of Integrative Plant Biology, 48, 1274–1286.CrossRefGoogle Scholar
Ljubešic, N. and Britvec, M. (2006). Tropospheric ozone-induced structural changes in leaf mesophyll cell walls in grapevine plants. Biologia, Bratislava, 61, 85–90.CrossRefGoogle Scholar
Lloret, F., Siscart, D. and Dalmases, C. (2004). Canopy recovery after drought dieback in holm-oak Mediterranean forests of Catalonia (NE Spain). Global Change Biology, 10, 2092–2099.CrossRefGoogle Scholar
Lloyd, J. (1991). Modelling stomatal responses to environment in Macadamia integrifolia. Australian Journal of Plant Physiology, 18, 649–660.CrossRefGoogle Scholar
Lloyd, J. and Farquhar, G.D. (1994). 13C discrimination during CO2 assimilation by the terrestrial biosphere. Oecologia, 99, 201–215.CrossRefGoogle ScholarPubMed
Lloyd, J. and Taylor, J. A. (1994). On the temperature dependence of soil respiration. Functional Ecology, 8, 315–323.CrossRefGoogle Scholar
Lloyd, J., Bird, M., Vellen, L., et al. (2008). Contributions of woody and herbaceous vegetation to tropical savanna ecosystem productivity: a quasi-global estimate. Tree Physiology, 28, 451–468.CrossRefGoogle ScholarPubMed
Lloyd, J., Krujit, B., Hollinger, D.Y., et al.(1996). Vegetation effects on the isotopic composition of CO2 at local and regional scales: theoretical aspects and a comparison between rain forest in Amazonia and a boreal forest in Siberia. Australian Journal of Plant Physiology, 23, 371–392.CrossRefGoogle Scholar
Lloyd, J., Syvertsen, J.P., Kriedemann, P.E., et al. (1992). Low conductances for CO2 diffusion from stomata to the sites of carboxylation in leaves of woody species. Plant, Cell and Environment, 15, 873–899.CrossRefGoogle Scholar
Lloyd, J., Wong, S.C., Styles, J.M., et al. (1995). Measuring and modelling whole-tree gas exchange. Australian Journal of Plant Physiology, 22, 987–1000.CrossRefGoogle Scholar
Lobel, R., Mamane, Y., Gepstein, S., et al. (2001). Particulate Matter Effect on Petunia “Blue Spark” Plants. Abstract of the Eurasap workshop on air pollution and the natural environment: 25–27 April 2001 Sofia, Bulgaria.
Loescher, H.W., Law, B.E., Mahrt, L., et al. (2006). Uncertainties in, and interpretation of, carbon flux estimates using the eddy covariance techniqueJournal of Geophysical Research, 111, 1–19.CrossRefGoogle Scholar
Loescher, H.W., Oberbauer, S.F., Gholz, H.L., et al. (2003). Environmental controls on net ecosystem-level carbon exchange and productivity in a Central American tropical wet forest. Global Change Biology, 9, 396–412.CrossRefGoogle Scholar
Loescher, H.W., Ocheltree, T.W., Tanner, B., et al. (2005). Comparison of temperature and wind statistics in contrasting environments among different sonic anemometer-thermometers. Agricultural and Forest Meteorology, 133, 119–139.CrossRefGoogle Scholar
Logan, B.A., Demmig-Adams, B., Rosenstiel, T.N., et al. (1999). Effect of nitrogen limitation on foliar antioxidants in relationship to other metabolic characteristics. Planta, 209, 213–220.CrossRefGoogle ScholarPubMed
Lohammer, T., Larsson, S., Linder, S., et al. (1980). FAST- Simulation models of gaseous exchange in Scots Pine. Ecological Bulletin, 32, 505–523.Google Scholar
Lohaus, G., Heldt, H.W. and Osmond, C.B. (2000). Infection with phloem limited Abutilon mosaic virus causes localized carbohydrate accumulation in leaves of Abutilon striatum: relationships to symptom development and effects on chlorophyll fluorescence quenching during photosynthetic induction. Plant Biology, 2, 161–167.CrossRefGoogle Scholar
Loik, M.E. and Holl, K.D. (1999). Photosynthetic responses to light for rainforest seedlings planted in abandoned pasture, Costa Rica. Restoration Ecology, 7, 382–391.CrossRefGoogle Scholar
Lomax, B.H., Woodward, F.I., Leitch, I.J., et al. (2009). Genome size as a predictor of guard cell length in Arabidopsis thaliana is independent of environmental conditions. The New Phytologist, 181, 311–314.CrossRefGoogle ScholarPubMed
Long, S.P. (1991). Modification of the response of photosynthetic productivity to rising temperature by atmospheric CO2 concentrations: has its importance been underestimated. Plant, Cell and Environment, 14, 729–739.CrossRefGoogle Scholar
Long, S.P. (1999). Environmental responses. In: The Biology of C4 Photosynthesis (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, USA, pp. 215–249.Google Scholar
Long, S.P. and Bernacchi, C.J. (2003). Gas exchange measurements, what can they tell us about the underlying limitations to photosynthesis? Procedures and sources of error. Journal of Experimental Botany, 54, 2393–2401.CrossRefGoogle ScholarPubMed
Long, S.P. and Drake, B.G. (1992). Photosynthetic CO2 assimilation and rising atmospheric CO2 concentrations. In: Crop Photosynthesis: Spatial and Temporal Determinants (eds Baker, N.R. and Thomas, H.), Elsevier Science Publishers, Amsterdam, Netherlands, pp. 69–103.Google Scholar
Long, S.P. and Hallgren, J.E. (1993). Measurement of CO2 assimilation by plants in the field and the laboratory. In: Photosynthesis and Production in a Changing Environment. A Fi eld and Laboratory Manual (eds Hall, D.O., Scurlock, J.M.O., Bolhar-Norden Kampf, H.R., Leegood, R.C. and Long, S.P.), pp. 129–167. Chapman and Hall, London.
Long, S.P. and Naidu, S.L. (2002). Effects of oxidants at the biochemical, cell and physiological levels, with particular reference to ozone. In: Air Pollution and Plant Life (eds Bell, J.N.B. and Treshow, M.) 2nd edn. John Wiley and Sons, Chichester, USA, pp. 69–88.Google Scholar
Long, S.P., Ainsworth, E.A., Leakey, A.D.B., et al. (2005). Global food insecurity. Treatment of major food crops with elevated carbon dioxide or ozone under large-scale fully open-air conditions suggests recent models may have overestimated future yields. Philosophical Transactions of the Royal Society B, 360, 2011–2020.CrossRefGoogle ScholarPubMed
Long, S.P., Ainsworth, E.A., Leakey, A.D.B., et al. (2006b). Food for thought: lower-than-expected crop yield stimulation with rising CO2 concentration. Science, 312, 1918–1921.CrossRefGoogle Scholar
Long, S.P., Ainsworth, E.A., Leakey, A.D.B., et al. (2007). Crop models, CO2, and climate change – Response. Science, 315, 460–460.Google Scholar
Long, S.P., Ainsworth, E.A., Rogers, A., et al. (2004). Rising atmospheric carbon dioxide: plants FACE the future. Annual Review of Plant Biology, 55, 591–628.CrossRefGoogle ScholarPubMed
Long, S.P., Farage, P.K. and García, R.L. (1996). Measurement of leaf and canopy photosynthetic CO2 exchange in the field. Journal of Experimental Botany, 47, 1629–1642.CrossRefGoogle Scholar
Long, S.P., Farage, P.K., Bolhar-Nordenkampf, H.R., et al. (1989). Separating the contribution of the upper and lower mesophyll to photosynthesis in Zea mays L. leaves. Planta, 177, 207–216.CrossRefGoogle ScholarPubMed
Long, S.P., Humphries, S. and Falkowski, P.G. (1994). Photoinhibition of photosynthesis in nature. Annual Review of Plant Physiology, 45, 633–662.CrossRefGoogle Scholar
Long, S.P., Postl, W.F. and Bolhár-Nordenkampf, H.R. (1993). Quantum yields for uptake of carbon dioxide in C3 vascular plants of contrasting habitats and taxonomic groupings. Planta, 189, 226–234.CrossRefGoogle Scholar
Long, S.P., Zhu, X.I.N.G., Naidu, S.L., et al. (2006a). Can improvement in photosynthesis increase crop yields?Plant, Cell and Environment, 29, 315–330.CrossRefGoogle ScholarPubMed
Loomis, R.S. and Amthor, J.S. (1999). Yield potential, plant assimilatory capacity, and metabolic efficiencies. Crop Science, 39, 1584–1596.CrossRefGoogle Scholar
López, C., Soto, M., Restrepo, S., Piégu, B., et al. (2005). Gene expression profile in response to Xanthomonas axonopodis pv. manihotis infection in cassava using a cDNA microarray. Plant Molecular Biology, 57, 393–410.CrossRefGoogle ScholarPubMed
Lopez-Hoffman, L., Anten, N.P.R., Martinez-Ramos, M., et al. (2007). Salinity and light interactively affect neotropical mangrove seedlings at the leaf and whole plant levels. Oecologia, 150, 545–556.CrossRefGoogle ScholarPubMed
Lopushinsky, W. and Kaufmann, M.R. (1984). Effects of cold soil on water relations and spring growth of Douglas-fir seedlings. Forest Science, 30, 628–634.Google Scholar
Lorenzini, G., Guidi, L., Nali, C., et al. (1997). Photosynthetic response of tomato plants to vascular wilt diseases. Plant Science, 124, 143–152.CrossRefGoogle Scholar
Loreto, F. and Fares, S. (2007). Is ozone flux inside leaves only a damage indicator? Clues from volatile isoprenoid studies. Plant Physiology, 143, 1096–1100.CrossRefGoogle Scholar
Loreto, F. and Sharkey, T.D. (1993). On the relationship between the isoprenene emission and photosynthetic metabolites under different environmental conditions. Planta, 189, 420–424.CrossRefGoogle ScholarPubMed
Loreto, F. and Velikova, V. (2001). Isoprene produced by leaves protects the photosynthetic apparatus against ozone damage, quenches ozone products, and reduces lipid peroxidation of cellular membranes. Plant Physiology, 127, 1781–1787.CrossRefGoogle ScholarPubMed
Loreto, F., Baker, N.R. and Ort, D.O. (2004b). Environmental constraints, chloroplast to leaf. In: Photosynthetic Adaptation, Chloroplast to Landscape (eds Smith, W.K., Vogelmann, T.C. and Critchley, C.), Springer, New York, USA, pp. 231–261.Google Scholar
Loreto, F., Centritto, M. and Chartzoulakis, K. (2003). Photosynthetic limitations in olive cultivars with different sensitivity to salt stress. Plant, Cell and Environment, 26, 595–601.CrossRefGoogle Scholar
Loreto, F., Delfine, S. and Di Marco, G. (1999). Estimation of photorespiratory carbon dioxide recycling during photosynthesis. Australian Journal of Plant Physiology, 26, 733–736.Google Scholar
Loreto, F., Di Marco, G., Tricoli, D., et al. (1994). Measurements of mesophyll conductance, photosynthetic electron transport and alternative electron sinks of field grown wheat leaves. Photosynthesis Research, 41, 397–403.CrossRefGoogle ScholarPubMed
Loreto, F., Harley, P.C., di Marco, G., et al. (1992). Estimation of mesophyll conductance to CO2 flux by three different methods. Plant Physiology, 98, 1437–1443.CrossRefGoogle Scholar
Loreto, F., Mannozzi, M., Maris, C., et al. (2001b). Ozone quenching properties of isoprene and its antioxidant role in leaves. Plant Physiology, 126, 993–1000.CrossRefGoogle ScholarPubMed
Loreto, F., Pinelli, P., Manes, F., et al. (2004a). Impact of ozone on monoterpene emissions and evidence for an isoprene-like antioxidant action of monoterpenes emitted by Quercus ilex leaves. Tree Physiology, 24, 361–367.CrossRefGoogle ScholarPubMed
Loreto, F., Tsonev, T. and Centritto, M. (2009). The impact of blue light on leaf mesophyll conductance. Journal of Experimental Botany, 60, 2283–2290.CrossRefGoogle ScholarPubMed
Loreto, F., Velikova, V. and Di Marco, G. (2001a). Respiration in the light measured by 12CO2 emission in 13CO2 atmosphere in maize leaves. Australian Journal of Plant Physiology, 28, 1103–1108.Google Scholar
Loriaux, S.D., Burns, R.A., Welles, J.M., et al. (2006). Determination of maximal chlorophyll fluorescence using a multiphase single flash of sub-saturating intensity. Poster. August, 2006. American Society of Plant Biologists Annual Meetings, Boston, MA. See: .Google Scholar
Lorimer, G.H., Badger, M.R. and Andrews, T.J. (1977). D-ribulose 1,5-bisphosphate carboxylase-oxygenase. Improved methods for activation and assay of catalytic activities. Analytical Biochemistry, 78, 66–75.CrossRefGoogle ScholarPubMed
Lösch, R., Jensen, C.R. and Andersen, M.N. (1992). Diurnal courses and factorial dependencies of leaf conductance and transpiration of differently potassium fertilized and watered field grown barley plants. Plant and Soil, 140, 205–224.CrossRefGoogle Scholar
Louis, J., Cerovic, Z.G. and Moya, I. (2006). Quantitative study of fluorescent excitation and emission spectra of bean leaves. Journal of Photochemistry and Photobiology B Biology, 85, 65–71.CrossRefGoogle Scholar
Louis, J., Ounis, A., Ducruet, J.M., et al. (2005). Remote sensing of sunlight-induced chlorophyll fluorescence and reflectance on Scots pine in the boreal forest during spring recovery. Remote Sensing of Environment, 96, 37–48.CrossRefGoogle Scholar
Loustau, D., Ben Brahim, M., Gaudillere, J.P., et al. (1999). Photosynthetic responses to phosphorus nutrition in two-year-old maritime pine seedlings. Tree Physiology, 19, 707–715.CrossRefGoogle ScholarPubMed
Loveless, A.R. (1962). Further evidence to support a nutritional interpretation of sclerophylly. Annals of Botany, 26, 551–561.CrossRefGoogle Scholar
Lovelock, C.E., Clough, B.F. and Woodrow, I.E. (1992). Distribution and accumulation of ultraviolet-radiation-absorbing compounds in leaves of tropical mangroves. Planta, 188, 143–154.CrossRefGoogle ScholarPubMed
Lovelock, C.E., Jebb, M. and Osmond, C.B. (1994). Photoinhibition and recovery in tropical plant species: response to disturbance. Oecologia, 97, 297–307.CrossRefGoogle ScholarPubMed
Lovelock, C.E., Osmond, C.B. and Seppelt, R.D. (1995). Photoinhibition in the Antarctic moss Grimmia antarctici Card. when exposed to cycles of freezing and thawing. Plant, Cell and Environment, 18, 1395–1402.CrossRefGoogle Scholar
Loveys, B.R., Schurwater, I., Pons, T.L., et al. (2001). The relative importance of photosynthesis and respiration in determining plant growth rate. In: Proceedings of the 12th International Congress on Photosynthesis, Brisbane Convention and Exhibition Centre, Queensland, Australia, August 18–23, 2001, CSIRO Publishing, Canberra, Australia, pp. S35–010.Google Scholar
Low, M., Haberle, K.H., Warren, C.R., et al. (2007). O3 flux-related responsiveness of photosynthesis, respiration, and stomatal conductance of adult Fagus sylvatica to experimentally enhanced free-air O3 exposure. Plant Biology, 9, 197–206.CrossRefGoogle ScholarPubMed
Löw, M., Herbinger, K., Nunn, A.J., et al. (2006). Extraordinary drought of 2003 overrules ozone impact on adult beech trees (Fagus sylvatica). Trees: Structure and Function, 20, 539–548.CrossRefGoogle Scholar
Lu, C., Lu, Q., Zhang, J., et al. (2001). Characterization of photosynthetic pigment composition, photosystem II photochemistry and thermal energy dissipation during leaf senescence of wheat plants grown in the field. Journal of Experimental Botany, 52, 1805–1810.CrossRefGoogle ScholarPubMed
Lu, X. and Zhuang, Q. (2010). Evaluating evapotranspiration and water-use efficiency of terrestrial ecosystems in the conterminous United States using MODIS and AmeriFlux data. Remote Sensing of the Environment, 114, 1924–1939.CrossRefGoogle Scholar
Lucht, W., Schaphoff, S., Erbrecht, T., et al. (2006). Terrestrial vegetation redistribution and carbon balance under climate change. Carbon Balance and Management, 6, 1–6.Google Scholar
Lüdeker, W., Dahn, H.G. and Günter, K.P. (1996). Detection of fungal infection of plants by laserinduced fluorescence: an attempt to use remote sensing. Journal of Plant Physiology, 148, 579–585.CrossRefGoogle Scholar
Ludwig, L.J. and Canvin, D.T. (1971). The rate of photorespiration during photosynthesis and the relationship of the substrate of light respiration to the products of photosynthesis in sunflower leaves. Plant Physiology, 48, 712–719.CrossRefGoogle ScholarPubMed
Lunde, C., Jensen, P.E., Haldrup, A., et al. (2000). The PSI-H subunit of photosystem I is essential for state transitions in plant photosynthesis. Nature, 408, 613–615.Google ScholarPubMed
Lundell, R., Saarinen, T., Åström, H., et al. (2008) The boreal dwarf shrub Vaccinium vitis-idaea retains its capacity for photosynthesis through the winter. Botany, 86, 491–500.CrossRefGoogle Scholar
Lundmark, M., Cavaco, A.M., Trevanion, S., et al. (2006). Carbon partitioning and export in transgenic Arabidopsis thaliana with altered capacity for sucrose synthesis grown at low temperature: a role for metabolite transporters. Plant Cell Environment, 29, 1703–1714.CrossRefGoogle ScholarPubMed
Lundmark, T., Bergh, J., Strand, M., et al. (1998). Seasonal variation of maximum photochemical efficiency in boreal Norway spruce stands. Trees-Struct Funct, 13, 63–67.CrossRefGoogle Scholar
Lunn, J.E., Feil, R., Hendriks, J.H.M., et al. (2006). Sugar-induced increases in trehalose 6-phosphate are correlated with redox activation of ADP-glucose pyrophosphorlyase and higher rates of starch synthesis in Arabidopsis thaliana. Biochemical Journal, 397, 139–148.CrossRefGoogle Scholar
Lunt, D.J., Ross, I.Hopley, P.J., et al. (2007). Modelling late Oligocene C4 grasses and climate. Palaeogeography, Palaeoclimatology, Palaecoecology, 251, 239–253.CrossRefGoogle Scholar
Luo, H., Oechel, W.C., Hastings, S.J., et al. (2007). Mature semiarid chaparral ecosystems can be a significant sink for atmospheric carbon dioxide. Global Change Biology, 13, 386–396.CrossRefGoogle Scholar
Luo, J., Zang, R. and Li, C. (2006). Physiological and morphological variations of Picea asperata populations originating from different altitudes in the mountains of southwestern China. Forest Ecology and Management, 221, 285–290.CrossRefGoogle Scholar
Luo, R., Wei, H., Ye, L., et al.. (2009). Photosynthetic metabolism of C3 plants shows highly cooperative regulation under changing environments: a systems biological analysis. Proceedings of the National Academy of Sciences USA, 106, 847–852.CrossRefGoogle ScholarPubMed
Luo, S., Su, B. and Currie, W.S. (2004). Progressive nitrogen limitation of ecosystem responses to rising atmospheric carbon dioxide. BioScience, 54, 731–739.CrossRefGoogle Scholar
Luo, Y., Gerten, D., Le Maire, G., et al. (2008). Modeled interactive effects of precipitation, temperature, and [CO2] on ecosystem carbon and water dynamics in different climatic zones. Global Change Biology, 14, 1986–1999.CrossRefGoogle Scholar
Luo, Y., Hui, D. and Zhang, D. (2006). Elevated CO2 stimulates net accumulations of carbon and nitrogen in land ecosystems: a meta-analysis. Ecology, 87, 53–63.CrossRefGoogle ScholarPubMed
Luo, Y., Su, B., Currie, W., et al. (2004). Progressive nitrogen limitation of ecosystem responses to rising atmospheric carbon dioxide. Bioscience, 54, 731–739.CrossRefGoogle Scholar
Luoma, S. (1997). Geographical pattern in photosynthetic light response of Pinus sylvestris in Europe. Functional Ecology, 11, 273–281.CrossRefGoogle Scholar
Lusk, C.H. (1996). Gradient analysis and disturbance history of temperate rain forests of the coast range summit plateau, Valdivia, Chile. Revista Chilena de Historia Natural, 69, 401–411.Google Scholar
Lusk, C.H. and Smith, B. (1998). Life history differeces and tree species coexistence in an old-growth New Zealand rain forest. Ecology, 79, 795–806.CrossRefGoogle Scholar
Lusk, C.H. and Warton, D.I. (2007). Global meta-analysis shows that relationships of leaf mass per area with species shade tolerance depend on leaf habit and ontogeny. The New Phytologist, 176, 764–774.CrossRefGoogle ScholarPubMed
Lusk, C.H., Wright, I. and Reich, P.B. (2003). Photosynthetic differences contribute to competitive advantage of evergreen angiosperm trees over evergreen conifers in productive habitats. New Phytologist, 160, 329–336.CrossRefGoogle Scholar
Lüttge, U. (1986). Nocturnal water storage in plants having crassulacean acid metabolism. Planta, 168, 287–289.Google ScholarPubMed
Lüttge, U. (1987a). Malate relations of plants and crassulacean acid metabolism: protons, carbon dioxide and water: a review. Giornale Botanico Italiano, 121, 217–227.CrossRefGoogle Scholar
Lüttge, U. (1987b). Carbon dioxide and water demand: crassulacean acid metabolism (CAM), a versatile ecological adaptation exemplifying the need for integration in ecophysiological work. New Phytologist, 106, 593–629.CrossRefGoogle Scholar
Lüttge, U. (1989). Vascular epiphytes. Setting the scene. In: Vascular Plants as Epiphytes. Evolution and Ecophysiology. Ecological Studies vol. 76 (ed. Lüttge, U.), Springer-Verlag, Berlin – Heidelberg – New York, pp. 1–14.CrossRefGoogle Scholar
Lüttge, U. (1993). The role of crassulacean acid metabolism (CAM) in the adaptation of plants to salinity. New Phytologist, 125, 59–71.CrossRefGoogle Scholar
Lüttge, U. (1997). Physiological Ecology of Tropical Plants. Springer-Verlag, Berlin-Heidelberg, Germany.CrossRefGoogle Scholar
Lüttge, U. (2000). The tonoplast functioning as the master switch for circadian regulation of crassulacean acid metabolism. Planta, 211, 761–769.Google ScholarPubMed
Lüttge, U. (2002a). CO2-concentrating: consequences in crassulacean acid metabolism. Journal of Experimental Botany, 53, 2131–2142.CrossRefGoogle Scholar
Lüttge, U. (2002b). Performance of plants with C4-carboxylation modes of photosynthesis under salinity. In: Salinity: Environment – Plants – Molecules (eds Läuchli, A. and Lüttge, U.) Kluwer Academic Publishers, Dordrecht, Boston, London, pp. 113–135.Google Scholar
Lüttge, U. (2003). Photosynthesis: CAM plants. In: Encyclopedia of Applied Plant Sciences (eds Thomas, B., Murphy, D. and Murphy, B.), Academic Press, Oxford, UK, pp. 688–705.Google Scholar
Lüttge, U. (2004). Ecophysiology of crassulacean acid metabolism (CAM). Annals of Botany, 93, 629–652.CrossRefGoogle Scholar
Lüttge, U. (2005). Genotypes – phenotypes – ecotypes: relations to crassulacean acid metabolism. Nova Acta Leopoldina NF, 92/342, 177–193.Google Scholar
Lüttge, U. (2006). Photosynthetic flexibility and ecophysiological plasticity: questions and lessons from Clusia, the only CAM tree, in the neotropics. New Phytologist, 171, 7–25.CrossRefGoogle ScholarPubMed
Lüttge, U. (2008a). Physiological Ecology of Tropical Plants. 2nd edn. Springer, Berlin – Heidelberg – New York.Google Scholar
Lüttge, U. (2008b). Stem CAM in arborescent succulents. Trees, 22, 139–148.CrossRefGoogle Scholar
Lüttge, U. and Beck, F. (1992). Endogenous rhythms and chaos in crassulacean acid metabolism. Planta, 188, 28–38.CrossRefGoogle ScholarPubMed
Lüttge, U. ed. (2007): Clusia. A Woody Neotropical Genus of Remarkable Plasticity and Diversity. Ecological Studies, Vol. 194. Springer-Verlag: Berlin-Heideiberg-New York.
Lüttge, U., Medina, E., Cram, W.J., et al. (1989). Ecophysiology of xerophytic and halophytic vegetation of a coastal alluvial plain in northern Venezuela. II. Cactaceae. New Phytologist, 111, 245–251.CrossRefGoogle Scholar
Lütz, C. (1996). Avoidance of photoinhibition and examples of photodestruction in high alpine Eriophorum. Journal of Plant Physiology, 148, 120–128.CrossRefGoogle Scholar
Luyssaert, S., Inglima, I., Jung, M., et al. (2007). CO2 balance of boreal, temperate, and tropical forests derived from a global database. Global Change Biology, 13, 2509–2537.CrossRefGoogle Scholar
Luyssaert, S., Reichstein, M., Schulze, E.D., et al. (2009). Toward a consistency cross-check of eddy covariance flux-based and biometric estimates of ecosystem carbon balance. Global Biogeochemical Cycles, 23, GB3009.CrossRefGoogle Scholar
Lv, S., Zhang, K., Gao, Q., et al. (2008). Overexpression of an H+-PPase gene from Thellungiella halophila in cotton enhances salt tolerance and improves growth and photosynthetic performance. Plant and Cell Physiology, 49, 1150–1164.CrossRefGoogle ScholarPubMed
Lyons, J.M. (1973). Chilling injury in plants. Annual Review of Plant Physiology, 24, 445–466.CrossRefGoogle Scholar
Ma, S., Baldocchi, D.D., Xu, L., et al. (2007). Inter-annual variability in carbón dioxide exchange of and oak/grass savanna and open grassland in California. Agricultural and Forest Metereology, 147, 151–171.Google Scholar
Maas, J.M. (1995). Conversion of tropical dry forest to pasture and agriculture. In: Seasonal Dry Tropical Forests (eds Bullock, S.H., Mooney, H.A. and Medina, E.), Cambridge University Press, Cambridge, UK, pp. 399–422.Google Scholar
Macedo, T.B., Weaver, D.K. and Peterson, R.K.D. (2007). Photosynthesis in wheat at the grain filling stage is altered by the wheat stem sawfly (Hymenoptera: Cephidae) injury and reduced water availability. Journal of Entomological Science, 42, 228–238.CrossRefGoogle Scholar
MacFadden, B.J. (2000). Cenozoic mammalian herbivores from the Americas: Reconstructing ancient diets and terrestrial communities. Annual Review of Ecology and Systematics, 31, 33–59.CrossRefGoogle Scholar
MacFadden, B.J. and Cerling, T.E. (1996). Mammalian herbivore communities, ancient feeding ecology, and carbon isotopes: A 10 million-year sequence from the Neogene of Florida. Journal of Vertebrate Paleontology, 16, 103–115.CrossRefGoogle Scholar
Mächler, F. and Nösberger, J. (1977). Effect of light intensity and temperature on apparent photosynthesis of altitudinal ecotypes of Trifolium repens L. Oecologia, 31, 73–78.CrossRefGoogle ScholarPubMed
Mächler, F. and Nösberger, J. (1978). The adaptation to temperature of photorespiration and of the photosynthetic carbon metabolism of altitudinal ecotypes of Trifolium repens L. Oecologia, 35, 267–276.CrossRefGoogle ScholarPubMed
Mächler, F., Nösberger, J. and Erismann, K.H. (1977). Photosynthetic (14)CO2 fixation products in altitudinal exotypes of Trifolium repens L. with different temperature requirements. Oecologia, 31, 79–84.CrossRefGoogle Scholar
Macinnis-Ng, C., McClenahan, K. and Eamus, D. (2004). Convergence in hydraulic architecture, water relations and primary productivity amongst habitats and across seasons in Sydney. Functional Plant Biology, 31, 429–439.CrossRefGoogle Scholar
MacKenzie, T.D.B., Krol, M., Huner, N.P.A., et al. (2002). Seasonal changes in chlorophyll fluorescence quenching and the induction and capacity of the photoprotective xanthophyll cycle in Lobaria pulmonaria. Canadian Journal of Botany Revue Canadienne De Botanique, 80, 255–261.Google Scholar
Macrobbie, E.A.C. (1998). Signal transduction and ion channels in guard cells. Philosophical Transactions of the Royal Society B, 353, 1475–1488.CrossRefGoogle ScholarPubMed
Magnani, F., Mencuccini, M., Borghetti, M., et al. (2007). The human footprint in the carbon cycle of temperate and boreal forests. Nature, 447, 848–850.CrossRefGoogle ScholarPubMed
Magnani, F. and Borghetti, M. (1995). Interpretation of seasonal changes of xylem embolism and plant hydraulic resistance in Fagus sylvatica. Plant Cell and Environment, 186, 689–696.CrossRefGoogle Scholar
Magnussen, S., Smith, V.G. and Yeatman, C.W. (1986). Foliage and canopy characteristics in relation to aboveground dry matter increment of seven jack pine provenances. Canadian Journal of Forest Research, 16, 464–470.CrossRefGoogle Scholar
Magrin, G., Gay García, C., Cruz Choque, D., et al. (2007). Latin America. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (eds Parry, O.F.C.M.L., Palutikof, J.P., van der Linden, P.J. and Hanson, C.E.), Cambridge University Press, Cambridge, UK, pp. 581–615.Google Scholar
Mahalingam, R., Jambunathan, N., Gunjan, S.K., et al. (2006). Analysis of oxidative signalling induced by ozone in Arabidopsis thaliana. Plant, Cell and Environment, 29, 1357–1371.CrossRefGoogle ScholarPubMed
Mahan, J.R. and Mauget, S.A. (2005). Antioxidant metabolism in cotton seedlings exposed to temperature stress in the field. Crop Science, 45, 2337–2345.CrossRefGoogle Scholar
Maherali, H., Sherrard, M.E., Clifford, M.H., et al. (2008). Leaf hydraulic conductivity and photosynthesis are genetically correlated in an annual grass. The New Phytologist, 180, 240–247.CrossRefGoogle Scholar
Mäkelä, A., Hari, P., Berninger, F., et al. (2004). Acclimation of photosynthetic capacity in Scots pine to the annual cycle of temperature. Tree Physiology, 24, 369–376.CrossRefGoogle Scholar
Makino, A. and Osmond, C.B. (1991). Effects of nitrogen nutrition on nitrogen partitioning between chloroplast and mitochondria in pea and wheat. Plant Physiology, 96, 355–362.CrossRefGoogle ScholarPubMed
Makino, A. and Sage, R.F. (2007). Temperature response of photosynthesis in transgenic rice transformed with ‘sense’ or ‘antisense’ rbcS. Plant Cell Physiology, 48, 1472–1483.CrossRefGoogle ScholarPubMed
Makino, A., Miyake, C. and Yokota, A. (2002). Physiological functions of the water-water cycle (Mehler reaction) and the cyclic electron flow around PSI in rice leaves. Plant and Cell Physiology, 43, 1017–1026.CrossRefGoogle ScholarPubMed
Makino, A., Nakano, H., Mae, T., et al. (2000). Photosynthesis, plant growth and N allocation in transgenic rice plants with decreased Rubisco under CO2 enrichment. Journal of Experimental Botany, 51, 383–389.CrossRefGoogle Scholar
Makino, A., Sakashita, H., Hidema, J., et al. (1992). Distinctive responses of ribulose-1,5-bisphosphate carboxylase and carbonic anhydrase in wheat leaves to nitrogen nutrition and their possible relationship to CO2 transfer resistance. Plant Physiology, 100, 1737–1743.CrossRefGoogle Scholar
Makino, A., Sato, T., Nakano, H., et al. (1997). Leaf photosynthesis, plant growth and nitrogen allocation in rice under different irradiances. Planta, 203, 390–398.CrossRefGoogle Scholar
Makino, A., Tadahiko, M. and Ohira, K. (1985). Enzymic Properties of ribulose-1,5-bisphosphate carboxylase/oxygenase purified from rice leaves. Plant Physiology, 79, 57–61.CrossRefGoogle ScholarPubMed
Maksymowych, R. (1973). Analysis of Leaf Development. Cambridge University Press, Cambridge, UK.
Malenovsky, Z., Mishra, K.B., Zemek, F., et al. (2009). Scientific and technical challenges in remote sensing of plant canopy reflectance and fluorescence. Journal of Experimental Botany, 60, 2987–3004.CrossRefGoogle ScholarPubMed
Malhi, Y., Nobre, A.D., Grace, J., et al. (1998). Carbon dioxide transfer over a Central Amazonian rain forest. Journal of Geophysical Research-Atmospheres, 103, 31593–31612.CrossRefGoogle Scholar
Malkin, S. and Cannaani, O. (1994). The use and characteristics of the photoacoustic method in the study of photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology, 45, 493–526.CrossRefGoogle Scholar
Malkin, S., Bilger, W. and Schreiber, U. (1994). The relationship between millisecond luminescence and fluorescence in tobacco leaves during the induction period. Photosynthesis Research, 39, 57–66.CrossRefGoogle ScholarPubMed
Man, R. and Lieffers, V.J. (1997). Seasonal photosynthetic responses to light and temperature in white spruce (Picea glauca) seedlings planted under an aspen (Populus tremuloides) canopy and in the open. Tree physiology, 17, 437–444.CrossRefGoogle ScholarPubMed
Manca, G. (2003). Analisi dei flussi di carbonio di una cronosequenza di cerro (Quercus cerris L.) dell’Italia centrale attraverso la tecnica della correlazione turbolenta. PhD Thesis. University of Tuscia, Viterbo, Italy.
Manetas, Y. (2004a). Photosynthesizing in the rain: beneficial effects of twig wetting on corticular photosynthesis trough changes in the periderm optical properties. Flora, 199, 334–341.CrossRefGoogle Scholar
Manetas, Y. (2004b). Probing corticular photosynthesis throughin vivo chlorophyll fluorescence measurements: evidence that high internal CO2 levels suppress electron flow and increase the risk of photoinhibition. Physiologia Plantarum, 120, 509–517.CrossRefGoogle Scholar
Manetas, Y. and Pfanz, H. (2005). Spatial heterogeneity of light penetration through periderm and lenticels and concomitant patchy acclimation of corticular photosynthesis. Trees, 19, 409–414.CrossRefGoogle Scholar
Mansfield, J.W. (2005). Biophoton distress flares signal the onset of the hypersensitive reaction. Trends in Plant Sciences, 10, 307–309.CrossRefGoogle ScholarPubMed
Mansfield, T.A. (1999). SO2 pollution: a bygone problem or a continuing hazard? In: Physiological Plant Ecology (ed. Press, M.C.), Blackwell Science, Oxford, UK, pp. 219–240.Google Scholar
Mansfield, T.A. (2002). Nitrogen oxides: old problems and new challenges. In: Air Pollution and Plant Life (ed. Bell, J.N.B. and Treshow, M.) 2nd edn, Wiley, John and Sons, Chichester, USA, pp. 119–134.Google Scholar
Mansfield, T.A. and Pearson, M. (1996). Disturbances in stomatal behaviour in plants exposed to air pollution. In: Plant Response to Air Pollution (eds Yunus, M. and Iqbal, M.), John Wiley and Sons, Chichester, USA, pp. 179–194.Google Scholar
Manter, D.K. and Kavanagh, K.L. (2003). Stomatal regulation in Douglas fir following a fungal-mediated chronic reduction in leaf area. Trees, 17, 485–491.CrossRefGoogle Scholar
Manter, D.K. and Kerrigan, J. (2004). A/Ci curve analysis across a range of woody plant species: influence of regression analysis parameters and mesophyll conductance. Journal of Experimental Botany, 55, 2581–2588.CrossRefGoogle Scholar
Manter, D.K., Kelsey, R.G. and Karchesy, J.J. (2007). Antimicrobial activity of extractable conifer heartwood compounds toward Phytophthora ramorum. Journal of Chemical Ecology, 33, 2133–2147.CrossRefGoogle ScholarPubMed
Manuel, N., Cornic, G., Aubert, S., et al. (1999). Protection against photoinhibition in the alpine plant Geum montanum. Oecologia, 119, 149–158.CrossRefGoogle ScholarPubMed
Marcelis, L.F.M. and Baan Hofman-Eijer, L.R.B. (1995). The contribution of fruit photosynthesis to the carbon requirement of cucumber fruits as affected by irradiance, temperature and ontogeny. Physiologia Plantarum, 93, 476–483.CrossRefGoogle Scholar
Marchi, S., Tognetti, R., Minnocci, A., et al. (2008). Variation in mesophyll anatomy and photosynthetic capacity during leaf development in a deciduous mesophyte fruit tree (Prunus persica) and an evergreen sclerophyllous Mediterranean shrub (Olea europaea). Trees, 22, 559–571.CrossRefGoogle Scholar
Marcolla, B., Pitacco, A. and Cescatti, A. (2003). Canopy architecture and turbulence structure in a coniferous forest. Boundary Layer Meteorology, 108, 39–59.CrossRefGoogle Scholar
Marín-Navarro, J., Manuell, A.L., Wu, J., et al. (2007). Chloroplast translation regulation. Photosynthesis Research, 94, 359–374.CrossRefGoogle ScholarPubMed
Markgraf, T. and Berry, J. (1990). Measurement of photochemical and non-photochemical quenching: correction for turnover of PS2 during steady state photosynthesis. In: Current Research in Photosynthesis, Vol IV (ed. Baltscheffsky, M.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 279–282.Google Scholar
Marks, S. and Clay, K. (1996). Physiological responses of Festuca arundinacea to fungal endophyte infection. New Phytologist, 133, 727–733.CrossRefGoogle Scholar
Maroco, J.P., Rodrigues, M.L., Lopes, C., et al. (2002). Limitations to leaf photosynthesis in field-grown grapevine under drought – metabolic and modelling approaches. Functional Plant Biology, 29, 451–459.CrossRefGoogle Scholar
Marques da Silva, J. and Arrabaça, M.C. (2004). Photosynthesis in the water-stressed C4 grass Setaria sphacelata is mainly limited by stomata with both rapidly and slowly imposed water deficits. Physiologia Plantarum, 121, 409–420.CrossRefGoogle Scholar
Marquez, E.J., Rada, F. and Farinas, M.R. (2006). Freezing tolerance in grasses along an altitudinal gradient in the Venezuelan Andes. Oecologia, 150, 393–397.CrossRefGoogle ScholarPubMed
Marschner, H. (1995). Mineral Nutrition of Higher Plants. Academic Press, London, UK.
Marschner, H. and Cakmak, I. (1989). High light intensity enhances chlorosis and necrosis in leaves of zinc-, potassium- and magnesium-deficient bean (Phaseolus vulgaris) plants. Journal of Plant Physiology, 134, 308–315.CrossRefGoogle Scholar
Marshall, F.M. (2002). Effects of air pollutants in developing countries. In: Air Pollution and Plant Life, 2nd edn (eds Bell, J.N.B. and Treshow, M.), John Wiley and Sons, Chichester, USA.Google Scholar
Mariscal, M.J., Orgaz, F. and Villalobos, F.J. (2000). Radiation-use efficiency and dry matter partitioning of a young olive (Olea europea) orchard. Tree Physiology, 20, 65–72.CrossRefGoogle Scholar
Martin, B. and Ruiz-Torres, N.A. (1992). Effects of water-deficit stress on photosynthesis, its components and component limitations, and on water-use efficiency in wheat (Triticum aestivum L). Plant Physiology, 100, 733–739.CrossRefGoogle Scholar
Martin, B. and Thorstenson, Y. R. (1988). Stable carbon isotope composition (δ13C), water use efficiency, and biomass productivity of Lycopersicon esculentum, Lycopersicon pennellii, and the F1 hybrid. Plant Physiology, 88, 213–217.CrossRefGoogle Scholar
Martin, B., Nienhuis, J., King, G., et al. (1989). Restriction fragment length polymorphisms associated with water-use efficiency in tomato. Science, 243, 1725–1728.CrossRefGoogle ScholarPubMed
Martin, C.E. (1994). Physiological ecology of the Bromeliaceae. Botanical Review, 60, 1–82.CrossRefGoogle Scholar
Martin, R.E., Asner, G.P. and Sack, L. (2007). Genetic variation in leaf pigment, optical and photosynthetic function among diverse phenotypes of Metrosideros polymorpha grown in a common garden. Oecologia, 151, 287–400.CrossRefGoogle Scholar
Martínez, J.P., Ledent, J.F., Bajji, M., et al. (2003). Effect of water stress on growth, Na+ and K+ accumulation and water use efficiency in relation to osmotic adjustment in two populations of Atriplex halimus L. Plant Growth Regulation, 41, 63–73.CrossRefGoogle Scholar
Martino-Catt, S. and Ort, D.R. (1992). Low-temperature interrupts circadian regulation of transcriptional activity in chilling-sensitive plants. Proc Natl Acad Sci USA, 89, 3731–3735.CrossRefGoogle ScholarPubMed
Maruyama, A. and Kuwagata, T. (2008). Diurnal and seasonal variation in bulk stomatal conductance of the rice canopy and its dependence on developmental stage. Agricultural and Forest Meteorology, 148, 1161–1173.CrossRefGoogle Scholar
Masclaux-Daubresse, C., Purdy, S., Lemaitre, T., et al. (2007). Genetic variation suggests interaction between cold acclimation and metabolic regulation of leaf senescence. Plant Physiology, 143, 434–446.CrossRefGoogle ScholarPubMed
Masle, J., Gilmore, S.R. and Farquhar, G.D. (2005). The ERECTA gene regulates plant transpiration efficiency in Arabidopsis. Nature, 436, 866–870.CrossRefGoogle ScholarPubMed
Masle, J., Hudson, G.S. and Badger, M.R. (1993). Effects of ambient CO2 concentration on growth and nitrogen use in tobacco (Nicotiana tabacum) plants transformed with an antisense gene to the small subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase. Plant Physiology, 103, 1075–1088.CrossRefGoogle ScholarPubMed
Mason, E.A. and Marrero, T.R. (1970). The diffusion of atoms and molecules. Advances in At Molecular Physics, 6, 155–232.CrossRefGoogle Scholar
Masoni, A., Ercoli, L. and Mariotti, M. (1996). Spectral properties of leaves deficient in iron, sulfur, magnesium and manganese. Agronomy Journal, 88, 937–943.CrossRefGoogle Scholar
Massad, R.S., Tuzet, A. and Bethenod, O. (2007). The effect of temperature on C4-type leaf photosynthesis parameters. Plant, Cell and Environment, 30, 1191–1204.CrossRefGoogle Scholar
Massman, W.J., Lee, X. (2002). Eddy covariance flux corrections and uncertainties in long-term studies of carbon and energy exchanges, Agricultural and Forest Meteorology, 113, 121–144.CrossRefGoogle Scholar
Massonnet, C., Costes, E., Rambal, S., et al. (2007). Stomatal regulation of photosynthesis in apple leaves: evidence for different water-use strategies between two cultivars. Annals of Botany, 100: 1347–1356.CrossRefGoogle ScholarPubMed
Massonnet, C., Regnard, J.L., Lauri, P.E., et al. (2008). Contributions of foliage distribution and leaf functions to light interception, transpiration and photosynthetic capacities in two apple cultivars at branch and tree scales. Tree Physiology, 28, 665–678.CrossRefGoogle ScholarPubMed
Mast, A.R. and Givnish, T.J. (2002). Historical biogeography and the origin of stomatal distributions in Banksia and Dryandra (Proteaceae) based on their cpDNA phylogeny. American Journal of Botany, 89, 1311–1323.CrossRefGoogle ScholarPubMed
Masterson, J. (1994). Stomatal size in fossil plants: evidence for polyploidy in majority of angiosperms. Science, 264, 421–424.CrossRefGoogle ScholarPubMed
Matile, P., Hörtensteiner, S. and Thomas, H. (1999). Chlorophyll degradation. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 67–95.CrossRefGoogle ScholarPubMed
Matouš, K., Benediktyová, Z., Berger, S., et al. (2006). Case study of combinatorial imaging: what protocol and what chlorophyll fluorescence image to use when visualizing infection of Arabidopsis thaliana by Pseudomonas syringae. Photosynthesis Research, 90, 243–253.CrossRefGoogle ScholarPubMed
Matson, P.A., Parton, W.J., Power, A.G., et al. (2007). Agricultural intensification and ecosystem properties. Science, 277, 504–509.CrossRefGoogle Scholar
Matsubara, S., Gilmore, A.M. and Osmond, C.B. (2001). Diurnal and acclimatory responses of violaxanthin and lutein epoxide in the Australian mistletoe Amyema miquelii. Autralian Journal of Plant Physiology, 28, 793–800.Google Scholar
Matsubara, S., Gilmore, A.M., Ball, M.C., et al. (2002). Sustained downregulation of photosystem II in mistletoes during winter depression of photosynthesis. Functional Plant Biology, 29, 1157–1169.CrossRefGoogle Scholar
Matsubara, S., Krause, G.H., Aranda, J., et al. (2009). Sun-shade patterns of leaf carotenoid composition in 86 species of neotropical forest plants. Functional Plant Biology, 36, 20–36.CrossRefGoogle Scholar
Matsubara, S., Krause, G.H., Seltmann, M., et al. (2008). Lutein epoxide cycle, light harvesting and photoprotection in species of the tropical tree genus Inga. Plant Cell and Environment, 31, 548–561.CrossRefGoogle ScholarPubMed
Matsubara, S., Morosinotto, T., Osmond, C.B., et al. (2007). Short- and long-term operation of the lutein-epoxide cycle in light-harvesting antenna complexes. Plant Physiology, 144, 926–941.CrossRefGoogle Scholar
Matsueda, H. and Inoue, H. (1996). Measurements of atmospheric CO2 and CH4 using a commercial airliner from 1993 to 1994. Atmospheric Environment, 30, 1647–1655.CrossRefGoogle Scholar
Matsumoto, K., Ohta, T. and Tanaka, T. (2005). Dependence of stomatal conductance on leaf chlorophyll concentration and meteorological variables. Agricultural and Forest Meteorology, 132, 44–57.CrossRefGoogle Scholar
Matsumoto, K., Ohta, T., Nakai, T., et al. (2008). Responses of surface conductance to forest environments in the Far East. Agricultural and Forest Meteorology, 148, 1926–1940.CrossRefGoogle Scholar
Matsumura, T., Tabayashi, N., Kamagata, Y., et al. (2002). Wheat catalase expressed in transgenic rice can improve tolerance against low temperature stress. Physiologia Plantarum, 116, 317–327.CrossRefGoogle Scholar
Matteucci, G. (1998). Bilancio del carbonio in una faggeta dell’Italia Centro-Meridionale: determinanti ecofisiologici, integrazione a livello di copertura e simulazione dell’impatto dei cambiamenti ambientali. PhD Thesis, University of Padua, Italy.
Matteucci, G., Valentini, R., Scarascia, Mugnosa G., et al. (1995). Struttura e funzionalita’ di una comunita’ vegetale Mediterranea a Quercus IlexL. dell’Italia Centrale. I. Bilancio del carbonio: variazioni stagionali e fattori limitanti. Studi Trent. Sci. Nat., Acta biologica, 69, 1992, 127–141.Google Scholar
Matthews, R.E.F. and Sarkar, S. (1976). A light-induced structural change in chloroplasts of chinese cabbage cells infected with turnip yellow mosaic virus. Journal of General Virology, 33, 435–446.CrossRefGoogle Scholar
Matthews, S. and Donoghue, M.J. (1999). The root of angiosperm phylogeny inferred from duplicate phytochrome genes. Science, 286, 947–950.CrossRefGoogle Scholar
Mattoo, A.K., Pick, U., Hoffman-Falk, H., et al. (1981). The rapidly metabolized 32,000-dalton polypeptide of the chloroplast is the “proteinaceous shield” regulating photosystem II electron transport and mediating diuron herbicide sensitivity. Proceedings of the National Academy of Sciences of the United States of America, 78, 1572–1576.CrossRefGoogle ScholarPubMed
Maurel, C., Verdoucq, L., Luu, D.T., et al. (2008). Plant aquaporins: membrane channels with multiple integrated functions. Annual Review of Plant Biology, 59, 595–624.CrossRefGoogle ScholarPubMed
Maximov, N.A. (1931). The physiological significance of the xeromorphic structure of plants. Journal of Ecology, 19, 273–282.CrossRefGoogle Scholar
Maximov, N.A. (1929). The Plants in Relation to Water. Allen and Unwin, London, UK.
Maxwell, D.P., Laudenbach, D.E. and Huner, N.P.A. (1995). Redox regulation of light-harvesting complex II and cab mRNA abundance in Dunaliella salina. Plant Physiology, 109, 787–795.CrossRefGoogle ScholarPubMed
Maxwell, D.P., Wong, Y. and McIntosh, L. (1999). The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells. Proceedings of the National Academy of Sciences, 96, 8271–8276.CrossRefGoogle ScholarPubMed
Maxwell, K. and Johnson, G.N. (2000). Chlorophyll fluorescence – a practical guide. Journal of Experimental Botany, 51, 659–668.CrossRefGoogle ScholarPubMed
Maxwell, K., von Caemmerer, S. and Evans, J.R. (1997). Is a low internal conductance to CO2 diffusion a consequence of succulence in plants with crassulacean acid metabolism?Australian Journal of Plant Physiology, 24, 777–786.CrossRefGoogle Scholar
May, J.D. and Killingbeck, K.T. (1992). Effects of preventing nutrient resorption on plant fitness and foliar nutrient dynamics. Ecology, 73, 1868–1878.CrossRefGoogle Scholar
Mayr, S., Schwienbacher, F. and Bauer, H. (2003). Winter at the Alpine timberline. Why does embolism occur in Norway spruce but not in stone pine?Plant Physiology, 131, 780–792.CrossRefGoogle Scholar
Mazzafera, P., Kubo, R.K. and Inomotos, M.M. (2004). Carbon fixation and partitioning in coffee seedlings infested with Pratylenchus coffeae. European Journal of Plant Pathology, 110, 861–865.CrossRefGoogle Scholar
Mazzella, M.A. and Casal, J.J. (2001). Interactive signalling by phytochromes and cryptochromes generates de-etiolation homeostasis in Arabidopsis thaliana. Plant Cell and Environment, 24, 155–161.CrossRefGoogle Scholar
McAinsh, M.R., Evans, N.H., Montgomery, L.T., et al. (2002). Calcium signalling in stomatal responses to pollutants. New Phytologist, 153, 441–447.CrossRefGoogle Scholar
McCallum, C.M., Comai, L., Greene, E.A., et al. (2000a). Targeted screening for induced mutations. Nature Biotechnology, 18, 455–457.CrossRefGoogle ScholarPubMed
McCallum, C.M., Comai, L., Greene, E.A., et al. (2000b). Targeting induced local lesions in genomes (TILLING) for plant functional genomics. Plant Physiology, 123, 439–442.CrossRefGoogle ScholarPubMed
McCarron, J.K. and Knapp, A.K. (2001). C3 woody plant expansion in a C4 grassland: are grasses and shrubs functionally distinct?American Journal of Botany, 88, 1818–1823.CrossRefGoogle Scholar
McCarthy, H.R., Pataki, D.E. and Jenerette, G.D. (2011). Plant water use efficiency as a metric of urban ecosystem services. Ecological Applications, 21, 3115–3127.CrossRef
McCarthy, I., Romero-Puertas, M.C., Palma, J.M., et al. (2001). Cadmium induces senescence symptoms in leaf peroxisomes of pea plants. Plant, Cell and Environment, 24, 1065–1073.CrossRefGoogle Scholar
McCool, M.M. (1935). Effect of light intensity on the manganese content of plants. Contributions from Boyce Thompson Institute, 7, 427–437.Google Scholar
McCree, K.J. (1986). Measuring the whole-plant daily carbon balance. Photosynthetica, 20, 82–93.Google Scholar
McCree, K.J. (1988). Sensitivity of sorghum grain yield to ontogenetic changes in respiration coefficients. Crop Science, 28, 114–120.CrossRefGoogle Scholar
McDermitt, D.K. (1990). Sources of error in the estimation of stomatal conductance and transpiration from porometer data. HortScience, 25, 1538–1547.Google Scholar
McDonald, E.P., Agrell, J. and Lindroth, R.L. (1999). CO2 and light effects on deciduous trees: growth, foliar chemistry, and insect performance. Oecologia, 119, 389–399.Google ScholarPubMed
McDowell, N.G. (2011). Mechanisms linking drought, hydraulics, carbon metabolism, and vegetation mortality. Plant Physiology, 155, 1051–1059.CrossRefGoogle ScholarPubMed
McElrone, A.J. and Forseth, I.N. (2004). Photosynthetic responses of a temperate liana to Xylella fastidiosa infection and water stress. Journal of Phytopathology, 152, 9–20.CrossRefGoogle Scholar
McElrone, A.J., Sherald, J.L. and Forseth, I.N. (2003). Interactive effects of water stress and xylem-limited bacterial infection on the water relations of a host vine. Journal of Experimental Botany, 54, 419–430.CrossRefGoogle ScholarPubMed
McElwain, J.C., Beerling, D.J., Woodward, F.I. (1999). Fossil plants and global warming at the Triassic-Jurassic boundary. Science, 285, 1386–1390.CrossRefGoogle ScholarPubMed
McElwain, J.C. and Chaloner, W.G. (1995). Stomatal density and index of fossil plants track atmospheric carbon dioxide in the Paleozoic. Annals of Botany, 76, 389–395.CrossRefGoogle Scholar
McGonigle, B. and Nelson, T. (1995). C4 isoform of NADP-malate dehydrogenase: cDNA cloning and expression in leaves of C4, C2, and C2-C4 intermediate species of Flaveria. Plant Physiology, 108, 1119–1126.CrossRef
McKee, I.F., Bullimore, J.F. and Long, S.P. (1997). Will elevated CO2 concentrations protect the yield of wheat from O3 damage?Plant, Cell and Environment, 20, 77–84.CrossRefGoogle Scholar
McKee, I.F., Farage, P.K. and Long, S.P. (1995). The interactive effects of elevated CO2 and O3 concentration on photosynthesis in spring wheat. Photosynthesis Research, 45, 111–119.CrossRefGoogle ScholarPubMed
McKenzie, R., Conner, B. and Bodeker, G. (1999). Increased summertime UV radiation in New Zealand in response to ozone loss. Science, 285, 1709–1711.CrossRefGoogle ScholarPubMed
McKinney, C.R., McCrea, J.M., Epstein, S., et al. (1950). Improvements in mass spectrometers for the measurement of small differences in isotope abundance ratios. Review of Scientific Instruments, 21, 724–730.CrossRefGoogle ScholarPubMed
McKown, A.D., Moncalvo, J.M. and Dengler, N.G. (2005). Phylogeny of Flaveria (Asteraceae) and inference of C-4 photosynthesis evolution. American Journal of Botany, 92, 1911–1928.CrossRefGoogle ScholarPubMed
McLeod, A.R. and Long, S.P. (1999). Free-air carbon dioxide enrichment (FACE) in Global Change Research. A review. Advances in Ecological Research, 28, 1–15.CrossRefGoogle Scholar
McMullen, C.R., Gardner, W.S. and Myers, G.A. (1978). Aberrant plastids in balley leaf tissue infected with balley stripe mosaic virus. Phytopathology, 68, 317–325.CrossRefGoogle Scholar
McNaughton, S.J., Milchunas, D. and Frank, D.A. (1996). How can net primary productivity be measured in grazing ecosystems?Ecology, 77, 974–977.CrossRefGoogle Scholar
McNevin, D.B., Badger, M.R., Whitney, S.M., et al. (2007). Differences in carbon isotope discrimination of three variant Rubiscos reflect differences in their catalytic mechanisms. Journal of Biological Chemistry, 282, 36068–36076.CrossRefGoogle Scholar
Medhurst, J., Parsby, J., Linder, S., et al. (2006). A whole-tree chamber system for examining tree-level physiological responses of field-grown trees to environmental variation and climate change. Plant, Cell and Environment, 29, 1853–1869.CrossRefGoogle ScholarPubMed
Medina, E. (1982). Temperature and humidity effects on dark CO2 fixation by Kalanchoë pinnata. Zeitschrift für Pflanzenphysiologie, 107, 251–258.CrossRefGoogle Scholar
Medina, E. (1987). Ecophysiological aspects of CAM plants in the tropics. Revista De Biologia Tropical, 35, 55–70.Google Scholar
Medina, E. and Delgado, M. (1976). Photosynthesis and night CO2 fixation in Echeveria columbiana Poellnitz. Photosynthetica, 10, 155–163.Google Scholar
Medina, E. and Francisco, M. (1994). Photosynthesis and water relations of savanna tree species differing in leaf phenology. Tree Physiology, 14, 1367–1381.CrossRefGoogle ScholarPubMed
Medina, E., Cram, W.J., Lee, H.S.J., et al. (1989). Ecophysiology of xerophytic and halophytic vegetation on an alluvial plain in northern Venezuela. I. Site description and plant communities. New Phytologist, 111, 233–243.CrossRefGoogle Scholar
Medlyn, B.E. (2004). A MAESTRO retrospective. In: Forests at the Land-atmosphere Interface (eds Mencuccini, M., Grace, J.C., Moncrieff, J. and McNaughton, K.), CAB International, Wallingford, UK, pp. 105–121.CrossRef
Medlyn, B.E., Badeck, F.W., De Pury, D.G.G., et al. (1999). Effects of elevated [CO2] on photosynthesis in European forest species: a meta-analysis of model parameters. Plant, Cell and Environment, 22, 1475–1495.CrossRefGoogle Scholar
Medlyn, B.E., Barton, C.V.M., Broadmeadow, M.S.J., et al. (2001). Stomatal conductance of European forest species after long-term exposure to elevated [CO2]: a synthesis of experimental data. New Phytologist, 149, 247–264.CrossRefGoogle 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.CrossRefGoogle Scholar
Medlyn, B.E., Dreyer, E., Ellsworth, D., et al. (2002a). Temperature response of parameters of a biochemically based model of photosynthesis. II. A review of experimental data. Plant, Cell and Environment, 25, 1167–1179.CrossRefGoogle Scholar
Medlyn, B.E., Loustau, D. and Delzon, S. (2002b). Temperature response of parameters of a biochemically based model of photosynthesis. I. Seasonal changes in mature maritime pine (Pinus pinaster Ait). Plant, Cell and Environment, 25, 1155–1165.CrossRefGoogle Scholar
Medrano, H., Keys, A.J., Lawlor, D.W., et al. (1995). Improving plant production by selection for survival at low CO2 concentrations. Journal of Experimental Botany, 46, 1389–1396.CrossRefGoogle Scholar
Medrano, H., Bota, J., Abadía, A., et al. (2002a). Effects of drought on light-energy dissipation mechanisms in high-light-acclimated, field-grown grapevines. Functional Plant Biology, 29, 1197–1207.CrossRefGoogle Scholar
Medrano, H., Escalona, J.M., Bota, J., et al. (2002b). Regulation of photosynthesis of C3 plants in response to progressive drought: stomatal conductance as a reference parameter. Annals of Botany, 89, 895–905.CrossRefGoogle ScholarPubMed
Medrano, H., Flexas, J. and Galmés, J. (2009). Variability in water use efficiency at the leaf level among Mediterranean plants with different growth forms. Plant and Soil, 317, 17–29.CrossRefGoogle Scholar
Medrano, H. and Primo-Millo, E. (1985). Selection of Nicotiana tabacum haploids of high photosynthetic effeciency. Plant Physiology, 79, 505–508.CrossRefGoogle Scholar
Meehl, G.A., Stocker, T.F., Collins, W.D., et al. (2007). Global climate projections. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (eds Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B.Tignor, M., and Miller, H.L.), Cambridge University Press, Cambridge, UK and New York, USA.Google Scholar
Meidner, H. (1975). Water supply, evaporation, and vapour diffusion in leaves. Journal of Experimental Botany, 26, 666–672.CrossRefGoogle Scholar
Meinzer, F.C. and Zhu, J. (1998). Nitrogen stress reduces the efficiency of the C4 CO2 concentrating system, and therefore quantum yield, in Saccharum (sugarcane) species. Journal of Experimental Botany, 49, 1227–1234.Google Scholar
Meinzer, F.C.., Wisdom, C. S., Gonzalez-Coloma, A., et al. (1990). Effects of leaf resin on stomata1 behaviour and gas exchange of Larrea tridentata (DC.) Cov. Functional Ecology, 4, 579–84.CrossRefGoogle Scholar
Meir, P., Kruijt, B., Broadmeadow, M., et al. (2002). Acclimation of photosynthetic capacity to irradiance in tree canopies in relation to leaf nitrogen concentration and leaf mass per unit area. Plant Cell Enviroment, 25, 343–357.CrossRefGoogle Scholar
Meister, M., Agostino, A. and Hatch, M.D. (1996). The roles of malate and aspartate in C4 photosynthetic metabolism in Flaveria bidentis (L.). Planta, 199, 262–269.CrossRefGoogle Scholar
Melakeberhan, H., Ferris, H. and Dias, J.M. (1990). Physiological response of resistant and susceptible Vitis vinifera to Meloidogyne incognita. Journal of Nematology, 22, 224–230.Google ScholarPubMed
Melillo, J.M., McGuire, A.D., Kicklighter, D.W., et al. (1993). Global climate change and terrestrial net primary production. Nature, 263, 234–239.CrossRefGoogle Scholar
Melis, A. (1999). Photosystem-II damage and repair cycle in chloroplasts: what modulates the rate of photodamage in vivo?Trends in Plant Science, 4, 130–135.CrossRefGoogle Scholar
Mellerowicz, E.J., Coleman, W.K., Riding, R.T., et al. (1992). Periodicity of cambial activity in Abies balsamea. I. Effects of temperature and photoperiod on cambial dormancy and frost hardiness. Physiologia Plantarum, 85, 515–525.CrossRefGoogle Scholar
Melotto, M., Underwood, W., Koczan, J., et al. (2006). Plant stomata function in innate immunity against bacterial invasion. Cell, 126, 969–980.CrossRefGoogle ScholarPubMed
Mencuccini, M., Mambelli, S. and Comstock, J. (2000). Stomatal responsiveness to leaf water status in common bean (Phaseolus vulgaris L.) is a function of time of day. Plant Cell Environment, 23, 1109–1118.CrossRefGoogle Scholar
Mendez, M., Jones, G.D. and Manetas, Y. (1999). Enhanced UV-B radiation under field conditions increases anthocyanin and reduces the risk of photoinhibition but does not affects growth in the carnivorous plant Pinguicula vulgaris. New Phytologist, 144, 1–8.CrossRefGoogle Scholar
Mendham, N.J., Rao, M.S.S. and Buzza, G.C. (1991). The apetalous flower character as a component of a high yielding ideotype. Proceedings of the 8th International Rapeseed Congress, Saskatoon, Canada, 2, 596–600.Google Scholar
Merlot, S., Leonhardt, N., Fenzi, F. et al. (2007). Constitutive activation of a plasma membrane H+-ATPase prevents abscisic acid-mediated stomatal closure. The EMBO Journal, 26, 3216–3226.CrossRefGoogle ScholarPubMed
Meroni, M. and Colombo, R. (2006). Leaf level detection of solar induced chlorophyll fluorescence by means of a subnanometer resolution spectroradiometer. Remote Sens. Environ., 103, 438–448.CrossRefGoogle Scholar
Meroni, M., Rossini, M., Guanter, L., et al. (2009). Remote sensing of solar- induced chlorophyll fluorescence: review of methods and applications. Remote Sensing of the Environment, 113, 2037–2051.CrossRefGoogle Scholar
Meroni, M., Rossini, M., Picchi, V., et al. (2008). Assessing steady state fluorescence and PRI from hyperspectral proximal sensing as early indicators of plant stress: the case of ozone exposure. Sensors 2008, 8, 1740–1754.Google ScholarPubMed
Merzlyak, M.N. and Gitelson, A. (1995). Why and what for the leaves are yellow in autumn? On the interpretation of optical spectra of senescing leaves (Acer platanoides L.). Journal of Plant Physiology, 145, 315–320.CrossRefGoogle Scholar
Mesarch, M.A., Walter-Shea, E.A., Asner, G.P., et al. (1999). A revised measurement methology for conifer needles spectral optical properties: evaluating the influence of gaps between elements. Remote Sensing of Environment, 68, 177–192.CrossRefGoogle Scholar
Messinger, S., Buckley, T.N. and Mott, K.A. (2006). Evidence for involvement of photosynthetic processes in the stomatal response to CO2. Plant Physiology, 140, 771–778.CrossRefGoogle Scholar
Mészáros, R., Zsely, I.G., Szinyei, D., et al. (2009). Sensitivity analysis of an ozone deposition model. Atmospheric Environment, 43, 663–672.CrossRefGoogle Scholar
Methy, M. (2000a). A two-channel hyperspectral radiometer for the assesment of photosynthetic radiation-use efficiency. Journal of Agricole Engineering Research, 75, 107–110.CrossRefGoogle Scholar
Méthy, M. (2000b). Stress-induced effects on Quercus ilex under a Mediterranean climate: contribution of chlorophyll fluorescece signatures. In: Life and the Environment in the Mediterranean (ed. Trabaud, L.), Wit Press, Southampton, USA, pp. 203–228.Google Scholar
Methy, M., Joffre, R. and Rambal, S. (1999). Remote sensing of canopy photosynthetic performances: two complementary ways for assessing the photochemical reflectance index. Photosynthetica, 37, 239–247.CrossRefGoogle Scholar
Meyer, A.J., Brach, T., Marty, L., et al. (2007). Redox-sensitive GFP in Arabidopsis thaliana is a quantitative biosensor for the redox potential of the cellular glutathione redox buffer. Plant Journal, 52, 973–986.CrossRefGoogle ScholarPubMed
Meyer, K.M., Ward, D., Moustakas, A., et al. (2005). Big is not better: small Acacia mellifera shrubs are more vital after fire. African Journal of Ecology, 43, 131–136.CrossRefGoogle Scholar
Meyer, M., Seibt, U. and Griffiths, H. (2008). To concentrate or ventilate? Carbon acquisition, isotope discrimination and physiological ecology of early land life forms. Philosophical Transactions of the Royal Society B, 363, 2767–2778.CrossRefGoogle ScholarPubMed
Meyer, S. and Genty, B. (1998). Mapping intercellular CO2 mole fraction (Ci) in Rosa rubiginosa leaves fed with abscisic acid by using chlorophyll fluorescence imaging. Significance of Ci estimated from leaf gas exchange. Plant Physiology, 116, 947–957.CrossRefGoogle ScholarPubMed
Meyer, S., Cartelat, A., Moya, I., et al. (2003). UV-induced blue-green and far-red fluorescence along wheat leaves: a potential signature of leaf ageing. Journal of Experimental Botany, 54, 757–769.CrossRefGoogle ScholarPubMed
Meyer, S., Cerovic, Z.G., Goulas, Y., et al. (2006). Relationships between optically assessed polyphenols and chlorophyll concentrations, and leaf mass per area ratio in woody plants: a signature of the carbon-nitrogen balance within leaves?Plant, Cell and Environment 29, 1338–1348.CrossRefGoogle Scholar
Meyer, S., Saccardy Adji, K., Rizza, F., et al. (2001). Inhibition of photosynthesis by Colletotrichum lindemuthianum in bean leaves determined by chlorophyll fluorescence imaging. Plant, Cell and Environment, 24, 947–955.CrossRefGoogle Scholar
Miao, Z.W., Xu, M., Lathrop, R.G., et al. (2009). Comparison of the A-C curve fitting methods in determining maximum ribulose 1,5-bisphosphate carboxylase/oxygenase carboxylation rate, potential light saturated electron transport rate and leaf dark respiration. Plant, Cell and Environment, 32, 109–122.CrossRefGoogle Scholar
Micol, J.L. (2009). Leaf development: time to turn over a new leaf?Current Opinion in Plant Biology, 12, 9–16.CrossRefGoogle Scholar
Middleton, E.M., Kim, M.S., Krizek, D.T., et al. (2005). Evaluating UV-B effects and EDU protection in soybean leaves using fluorescence. Photochemistry and Photobiology, 81, 1075–1085.CrossRefGoogle ScholarPubMed
Mikkelsen, T.N. and Heide-Jųrgensen, H.S. (1996). Acceleration of leaf senescence in Fagus sylvatica L. by low levels of tropospheric ozone demonstrated by leaf colour, chlorophyll fluorescence and chloroplast ultrastructure. Trees: Structure and Function, 10, 145–156.Google Scholar
Miles, C.D., Brandle, J.R., Daniel, D.J., et al. (1972). Inhibition of photosystem II in isolated chloroplasts by lead. Plant Physiology, 49, 820–825.CrossRefGoogle ScholarPubMed
Millar, A.H., Heazlewood, J.L., Kristensen, B.K., et al. (2005). The plant mitochondrial proteome. Trends in Plant Science, 10, 36–43.CrossRefGoogle ScholarPubMed
Millar, A.H., Mittova, V., Kiddle, G., et al. (2003). Control of ascorbate synthesis by respiration and its implications for stress responses. Plant Physiology, 133, 443–447.CrossRefGoogle ScholarPubMed
Miller, A., Schlagnhaufer, C., Spalding, M., et al. (2000). Carbohydrate regulation of leaf development: prolongation of leaf senescence in Rubisco antisense mutants of tobacco. Photosynthesis Research, 63, 1–8.CrossRefGoogle Scholar
Miller, A.M., van Iersel, M.W. and Armitage, A.M. (2001). Whole-plant carbon dioxide exchange responses of Angelonia angustifolia to temperature and irradiance. Journal of the American Society for Horticultural Science, 125, 606–610.Google Scholar
Miller, E.W., Albers, A.E., Pralle, A., et al. (2005). Boronate-based fluorescent probes for imaging cellular hydrogen peroxide. Journal of the American Chemical Society, 127, 16652–16659.CrossRefGoogle ScholarPubMed
Miller, G., Suzuki, N., Ciftci-Yilmaz, S., et al. (2010). Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant, Cell and Environment, 33, 453–467.CrossRefGoogle ScholarPubMed
Miller, J.B., Yakir, D., White, J.W.C., et al. (1999). Measurements of 18O/16O in the soil-atmosphere CO2 flux. Global Biogeochemical Cycles, 13, 761–774.CrossRefGoogle Scholar
Miller, J.M., Williams, R.J. and Farquhar, G.D. (2001). Carbon isotope discrimination by a sequence of Eucalyptus species along a subcontinental rainfall gradient in Australia. Functional Ecology, 15, 222–232.CrossRefGoogle Scholar
Miller, P.M., Eddleman, L.E. and Miller, J.M. (1995). Juniperus occidentalis juvenile foliage: advantages and disadvantages for a stress-tolerant, invasive conifer. Canadian Journal of Forest Research, 25, 470–479.CrossRefGoogle Scholar
Miller, R.E., Watling, J.R. and Robinson, S.A. (2009). Functional transition in the floral receptacle of the sacred lotus (Nelumbo nucifera): from thermogenesis to photosynthesis. Functional Plant Biology, 36, 471–480.CrossRefGoogle Scholar
Mills, G., Ball, G., Hayes, F., et al. (2000). Development of a multi-factor model for predicting the critical level of ozone for white clover. Environmental Pollution, 109, 533–542.CrossRefGoogle Scholar
Mills, G.A. and Urey, H.C. (1940). The kinetics of isotopic exchange between carbon dioxide, bicarbonate ion, carbonate ion and water. Journal of the American Chemical Society, 62, 1019–1026.CrossRefGoogle Scholar
Miloslavina, Y., Nilkens, M., Jahns, P., et al. (2007). Reorganization of LHC II complexes in the thylakoid membrane in response to high light adaptation (NPQ). Photosynthesis Research, 91, 256–256.Google Scholar
Minamikawa, T., Toyooka, K., Okamoto, T., et al. (2001). Degradation of ribulose-bisphosphate carboxylase by vacuolar enzymes of senescing French bean leaves: immunocytochemical and ultrastructural observations. Protoplasma, 218, 144–153.CrossRefGoogle ScholarPubMed
Miranda, T. and Ducruet, J.M. (1995). Characterization of the chlorophyll thermoluminescence afterglow in dark-adapted or far-red-illuminated plant leaves. Plant Physiology and Biochemistry, 33, 689–699.Google Scholar
Mishra, R.K. and Singhal, G.S. (1993). Photosynthetic activity and peroxidation of thylakoid lipids during photoinhibition and high temperature treatment of isolated wheat chloroplasts. Journal of Plant Physiology, 141, 286–292.CrossRefGoogle Scholar
Misson, L., Baldocchi, D.D., Black, T.A., et al. (2007.) Partitioning forest carbon fluxes with overstory and understory eddy covariance measurements: a synthesis based on FLUXNET data. Agricultural and Forest Meteorology, 144, 14–31.CrossRefGoogle Scholar
Misson, L., Limousin, J.M., Rodriguez, R., et al. (2010). Leaf physiological responses to extreme droughts in Mediterranean Quercus ilex forest. Plant, Cell and Environment, 33, 1898–1910.CrossRefGoogle ScholarPubMed
Misson, L., Panek, J.A. and Goldstein, A.H. (2004). A comparison of three approaches to modeling leaf gas exchange in annually drought-stressed ponderosa pine forests. Tree Physiology, 24, 529–541.CrossRefGoogle ScholarPubMed
Mitchell, C.A. (1992). Measurement of photosynthetic gas exchange in controlled environments. HortScience, 27, 764–767.Google ScholarPubMed
Mitchell, T.D., Carter, T.R., Jones, P.D., et al. (2004). A comprehensive set of high resolution grids of monthly climate for Europe and the globe: the observed record (1901–2000) and 16 scenarios (2001–2100). Working paper 55 (Tyndall Centre for Climate Change Research, July 2004); available at _papers/wp55.pdf.
Mittler, R. (2002). Oxidative stress, antioxidants and stress tolerance. Trends in Plant Science, 7, 405–410.CrossRefGoogle ScholarPubMed
Mittler, R. (2006). Abiotic stress, the field environment and stress combination. Trends in Plant Science, 11, 15–19.CrossRefGoogle ScholarPubMed
Mittler, R. and Blumwald, E. (2010). Genetic engineering for modern agriculture: challenges and perspectives. Annual Review of Plant Biology, 61, 443–462.CrossRefGoogle Scholar
Mittler, R., Vanderauwera, S., Gollery, M., et al. (2004). Reactive oxygen gene network of plants. Trends in Plant Science, 9, 490–498.CrossRefGoogle Scholar
Miyaji, K., Da Silva, W.S. and De Paulo, T.A. (1997). Longeivity of leaves of a tropical tree, Theombroma cacao, grown under shading, in relation to position within the canopy and time of emergence. New Phytologist, 135, 445–454.CrossRefGoogle Scholar
Miyake, C. and Yokota, A. (2000). Determination of the rate of photoreduction of O2 in the water-water cycle in watermelon leaves and enhancement of the rate by limitation of photosynthesis. Plant Cell Physiology, 41, 335–343.CrossRefGoogle ScholarPubMed
Miyashita, K., Tanakamaru, S., Maitani, T., et al. (2005). Recovery responses of photosynthesis, transpiration, and stomatal conductance in kidney bean following drought stress. Environmental and Experimental Botany, 53, 205–214.CrossRefGoogle Scholar
Miyazawa, S.I. and Terashima, I. (2001). Slow development of leaf photosynthesis in an evergreen broad-leaved tree, Castanopsis sieboldii: relationships between leaf anatomical characteristics and photosynthetic rate. Plant, Cell and Environment, 24, 279–291.CrossRefGoogle Scholar
Miyazawa, S.I., Makino, A. and Terashima, I. (2003). Changes in mesophyll anatomy and sink-source relationships during leaf development in Quercus glauca, an evergreen tree showing delayed leaf greening. Plant Cell Environment, 26, 745–755.CrossRefGoogle Scholar
Miyazawa, S.I., Satomi, S. and Terashima, I. (1998). Slow leaf development of evergreen broad-leaved tree species in Japanese warm temperate forests. Annals of Botany, 82, 859–869.CrossRefGoogle Scholar
Miyazawa, S.I., Yoshimura, S., Shinzaki, Y., et al. (2008). Deactivation of aquaporins decreases internal conductance to CO2 diffusion in tobacco leaves grown under long term drought. Functional Plant Biology, 35, 553–564.CrossRefGoogle Scholar
Miyazawa, Y. and Kikuzawa, K. (2005a). Winter photosynthesis by saplings of evergreen broadleaved trees in a deciduous temperate forest. The New Phytologist, 165, 857–866.CrossRefGoogle Scholar
Miyazawa, Y. and Kikuzawa, K. (2005b). Physiological basis of seasonal trend in leaf photosynthesis of five evergreen broad-leaved species in a temperature deciduous forest. Tree Physiology, 26, 249–256.CrossRefGoogle Scholar
Miyazawa, Y. and Kikuzawa, K. (2006). Photosynthesis and physiological traits of evergreen broadleafed saplings during winter under different light environments in a temperate forest. Canadian Journal of Botany, 85, 60–69.CrossRefGoogle Scholar
Mochizuki, N., Tanaka, R., Tanaka, A., et al. (2008). The steady state level of mg-protoporphyrin IX is not a determinant of plastid-to-nucleus signalling in Arabidopsis. Proceedings of the National Academy of Sciences USA, 105, 15,184–15,189.CrossRefGoogle ScholarPubMed
Moffat, A.M., Papale, D., Reichstein, M., et al. (2007). Comprehensive comparison of gap-filling techniques for eddy covariance net carbon fluxes. Agricultural and Forest Meteorology, 147, 209–232.CrossRefGoogle Scholar
Moghaddam, P.R. and Wilman, D. (1998). Cell wall thickness and cell dimensions in plant parts of eight forage species. Journal of Agricultural Science Cambridge, 131, 59–67.CrossRefGoogle Scholar
Moghaieb, R.E.A., Tanaka, N., Saneoka, H., et al. (2008). Characterization of salt tolerance in ectoine-transformed tobacco plants (Nicotiana tabaccum): photosynthesis, osmotic adjustment, and nitrogen partitioning. Plant, Cell and Environment, 29, 173–182.CrossRefGoogle Scholar
Mohanty, P., Vani, B. and Prakash, J.S.S. (2002). Elevated temperature treatment induced alteration in thylakoid membrane organization and energy distribution between the two photosystems in Pisum sativum. Zeitschrift fur Natuforschung part C, 57, 836–842.Google ScholarPubMed
Moise, N. and Moya, I. (2004a). Correlation between lifetime heterogeneity and kinetics heterogeneity during chlorophyll fluorescence induction in leaves: 1. Mono-frequency phase and modulation analysis reveals a conformational change of a PSII pigment complex during the IP thermal phase. Biochimica et Biophysica Acta, 1657, 33–46.CrossRefGoogle ScholarPubMed
Moise, N. and Moya, I. (2004b). Correlation between lifetime heterogeneity and kinetics heterogeneity during chlorophyll fluorescence induction in leaves: 2. Multi-frequency phase and modulation analysis evidences a loosely connected PSII pigment-protein complex. Biochimica et Biophysica Acta, 1657, 47–60.CrossRefGoogle ScholarPubMed
Mommer, L., Pons, T.L., Wolters-Arts, M., et al. (2005). Submergence-induced morphological, anatomical, and biochemical responses in a terrestrial species affect gas diffusion resistance and photosynthetic performance. Plant Physiology, 139, 497–508.CrossRefGoogle Scholar
Monclus, R., Dreyer, E., Delmotte, F.M., et al. (2005) Productivity, leaf traits and carbon isotope discrimination in 29 Populus deltoides × P. nigra clones. New Phytologist, 167, 53–62.CrossRefGoogle ScholarPubMed
Monclus, R., Dreyer, E., Villar, M., et al. (2006). Impact of drought on productivity and water use efficiency in 29 genotypes of Populus deltoides × Populus nigra. New Phytologist, 169, 765–777.CrossRefGoogle ScholarPubMed
Moncrieff, J.B., Malhi, Y. and Leuning, R. (1996). The propagation of errors in long-term measurements of land atmosphere fluxes of carbon and water. Global Change Biology, 2, 231–240.CrossRefGoogle Scholar
Monje, O. and Bugbee, B. (1998). Adaptation to high CO2 concentration in an optimal environment; radiation capture, canopy quantum yield and carbon use efficiency. Plant Cell and Environment, 21, 315–324.CrossRefGoogle Scholar
Monk, C.D. (1966). An ecological significance of evergreens. Ecology, 47, 504–505.CrossRefGoogle Scholar
Monnet, F., Vaillant, N., Vernay, P., et al. (2001). Relationship between PSII activity, CO2 fixation, and Zn, Mn and Mg contents of Lolium perenne under zinc stress. Journal of Plant Physiology, 158, 1137–1144.CrossRefGoogle Scholar
Monroy, A.F. and Dhindsa, R.S. (1995). Low-temperature signal transduction: induction of cold acclimation-specific genes of alfalfa by calcium at 25 degrees C. Plant Cell, 7, 321–331.Google ScholarPubMed
Monsi, M. and Saeki, T. (1953). Über den lichtfaktor in den pflanzengesellschaften und seine bedeutung für die stoffproduktion. Japanese Journal of Botany, 14, 22–52.Google Scholar
Monsi, M. and Saeki, T. (2005). On the factor light in plant communities and ist importance for matter production. Annals of Botany, 95, 549–567.CrossRefGoogle Scholar
Monson, R.K. (1989a). The relative contributions of reduced photorespiration, and improved water- and nitrogen-use efficiencies, to the advantages of C3-C4 intermediate photosynthesis in Flaveria. Oecologia, 80, 215–221.CrossRefGoogle ScholarPubMed
Monson, R.K. (1989b). On the evolutionary pathways resulting in C4 photosynthesis and crassulacean acid metabolism. Advances in Ecological Research, 19, 57–110.CrossRefGoogle Scholar
Monson, R.K. (1999). The origins of C4 genes and evolutionary pattern in the C4 metabolic phenotype. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, CA, USA, pp. 377–410.CrossRef
Monson, R.K. (2003). Gene duplication, neofunctionalization and the evolution of C4 photosynthesis. International Journal of Plant Sciences, 164 (suppl), S43–S54.CrossRefGoogle Scholar
Monson, R.K. and Rawsthorne, S. (2000). CO2 assimilation in C3-C4 intermediate plants. In: Photosynthesis: Physiology and Metabolism, Advances in Photosynthesis (eds Leegood, R.C., Sharkey, T.D. and von Caemmerer, S.), Kluwer Academic Press, New York, USA, pp. 533–550.Google Scholar
Monson, R.K., Littlejohn, R.O. Jr and Williams, G.J. III (1983). Photosynthetic adaptation to temperature in four species from the Colorado shortgrass steppe: a physiological model for coexistence. Oecologia, 58, 43–51.CrossRefGoogle ScholarPubMed
Monson, R.K., Sparks, J.P., Rosenstiel, T.N., et al. (2005). Climatic influences on net ecosystem CO2 exchange during the transition from wintertime carbon source to springtime carbon sink in a high-elevation, subalpine forest. Oecologia, 146, 130–147.CrossRefGoogle Scholar
Monson, R.K., Turnipseed, A.A., Sparks, J.P. et al. (2002). Carbon sequestration in a high-elevation, subalpine forest. Global Change Biology, 8, 459–478.CrossRefGoogle Scholar
Montagu, K.D. and Woo, K.C. (1999). Recovery of tree photosynthetic capacity from seasonal drought in the wet-dry tropics: the role of phyllode and canopy processes in Acacia auriculiformis. Australian Journal of Plant Physiology, 26, 135–145.CrossRefGoogle Scholar
Montalbini, P. and Lupattelli, M. (1989). Effect of localized and systemic tobacco mosaic-virus infection on some photochemical and enzymatic-activities of isolated tobacco chloroplasts. Physiological and Molecular Plant Pathology, 34, 147–162.CrossRefGoogle Scholar
Montanaro, G., Dichio, B. and Xiloyannis, C. (2007). Response of photosynthetic machinery of field-grown kiwifruit under Mediterranean conditions during drought and re-watering. Photosynthetica, 45, 533–540.CrossRefGoogle Scholar
Monteiro, J.A.F. and Prado, C.H.B.A. (2006). Apparent carboxylation efficiency and relative stomatal and mesophyll limitations of photosynthesis in an evergreen cerrado species during water stress. Photosynthetica, 44, 39–45.CrossRefGoogle Scholar
Monteith, J.L. (1972). Solar radiation and productivity in tropical ecosystems. Journal of Applied Ecology, 9, 747–766.CrossRefGoogle Scholar
Monteith, L.J. (1973). Principles of Environmental Physics. American Elsevier, New York, USA, pp. 243.Google Scholar
Monteith, J.L. (1984). Consistency and convenience in the choice of units for agricultural science. Experimental Agriculture, 20, 105–117.CrossRefGoogle Scholar
Monteith, J.L. and Unsworth, M.H. (1990). Principles of Environmental Physics. Edward Arnold, London, UK.Google Scholar
Montesano, M., Scheller, H.R., Wettstein, R., et al. (2004). Down-regulation of photosystem I by Erwinia carotovora-derived elicitors correlates with H2O2 accumulation in chloroplasts of potato. Molecular Plant Pathology, 5, 115–123.CrossRefGoogle Scholar
Montgomery, R., Goldstein, G. and Givnish, T.J. (2008). Photoprotection of PSII in Hawaiian lobeliads from diverse light environments. Functional Plant Biology, 35, 595–605.CrossRefGoogle Scholar
Montgomery, R.A. (2004a). Effects of understory foliage on patterns of light attentuation near the forest floor. Biotropica, 36, 33–39.Google Scholar
Montgomery, R.A. (2004b). Relative importance of photosynthetic physiology and biomass allocation for tree seedling growth across a broad light gradient. Tree Physiology, 24, 155–167.CrossRefGoogle ScholarPubMed
Montgomery, R.A. and Givnish, T.J. (2008). Adaptive radiation of photosynthetic physiology in the Hawaiian lobeliads: dynamic photosynthetic responses. Oecologia, 155, 455–467.CrossRefGoogle ScholarPubMed
Monti, A., Bezzi, G. and Venturi, G. (2009). Internal conductance under different light conditions along the plant profile of Ethiopian mustard (Brassica carinata A. Brown). Journal of Experimental Botany, 60, 2341–2350.CrossRefGoogle Scholar
Monti, A., Brugnoli, E, Scartazza, A., et al. (2006). The effect of transient and continuous drought on yield, photosynthesis and carbon isotope discrimination in sugar beet (Beta vulgaris L.). Journal of Experimental Botany, 57, 1253–1262.CrossRefGoogle Scholar
Montpied, P., Granier, A. and Dreyer, E. (2009). Seasonal time-course of gradients of photosynthetic capacity and mesophyll conductance to CO2 across a beech (Fagus sylvatica L.) canopy. Journal of Experimental Botany, 60, 2407–2418.CrossRefGoogle ScholarPubMed
Montzka, S.A., Calvert, P., Hall, B.D., et al. (2007). On the global distribution, seasonality, and budget of atmospheric carbonyl sulfide (COS) and some similarities to CO2. Journal of Geophysical Research – Atmospheres 112: Article number D09302.CrossRefGoogle Scholar
Mook, W.G. (1986). 13C in atmospheric CO2. Netherlands Journal Sea Research, 20, 211–223.CrossRefGoogle Scholar
Mook, W.G., Bommerson, J.C., Staverman, W.H. (1974). Carbon isotope fractionation between dissolved carbonate and gaseous carbon dioxide. Earth Planet Science Letters, 22, 169–176.CrossRefGoogle Scholar
Mook, W.G., Koopmans, M., Carter, A.F., Keeling, C.D. (1983). Seasonal, latitudinal and secular variations in the abundance and isotopic ratios of atmospheric carbon dioxide. 1. Results from land stations. Journal of Geophysical Research, 88, 10915–10933.CrossRefGoogle Scholar
Mooney, H.A. (1972). The carbon balance of plants. Annual Review of Ecology and Systematics, 3, 315–346.CrossRefGoogle Scholar
Mooney, H.A. and Billings, W.D. (1961). Comparative physiological ecology of Arctic and Alpine populations of Oxyria digyna. Ecological Monographs, 31, 1–29.CrossRefGoogle Scholar
Mooney, H.A. and Ehleringer, J.R. (1978). The carbon gain benefits of solar tracking in a desert annual. Plant, Cell and Environment, 1, 307–11.CrossRefGoogle Scholar
Mooney, H.A., Björkman, O. and Collatz, G.J. (1978). Photosynthetic acclimation to temperature in the desert shrub Larrea divaricata. I. Carbon dioxide exchange characteristics of intact leaves. Plant Physiology, 61, 406–410.CrossRefGoogle ScholarPubMed
Mooney, H.A., Ehleringer, J.R. and Berry, J.A. (1976). High photosynthetic capacity of a winter desert annual in Death Valley. Science, 194, 322–324.CrossRefGoogle Scholar
Mooney, H.A., Field, C., Gulmon, S. L., et al. (1981). Photosynthetic capacity in relation to leaf positions in desert versus old-field annuals. Oecologia, 50, 109–112.CrossRefGoogle Scholar
Mooney, H.A., Pearcy, R.W. and Ehleringer, J. (1987). Plant physiological ecology today. BioScience, 37, 18–20.CrossRefGoogle Scholar
Moore, B.D. and Seemann, J.R. (1992). Metabolism of 2’-carboxyarabinitol in leaves. Plant Physiolology, 99, 1551–1555.CrossRefGoogle ScholarPubMed
Moore, B.D., Cheng, S.H., Sims, D., et al. (1999). The biochemical and molecular basis for photosynthetic acclimation to elevated atmospheric CO2. Plant Cell and Environment, 22, 567–582.CrossRefGoogle Scholar
Moore, R.C. and Purugganan, M.D. (2005). The evolutionary dynamics of plant duplicate genes. Current Opinion in Plant Biology, 8, 122–128.CrossRefGoogle ScholarPubMed
Morales, F., Abadía, A. and Abadía, J. (1990). Characterization of the xanthophyll cycle and other photosynthetic pigment changes induced by iron deficiency in sugar beet (Beta vulgaris L.). Plant Physiology, 94, 607–613.CrossRefGoogle Scholar
Morales, F., Abadía, A. and Abadía, J. (1991). Chlorophyll fluorescence and photon yield of oxygen evolution in iron-deficient sugar beet (Beta vulgaris L.) leaves. Plant Physiology, 97, 886–893.CrossRefGoogle ScholarPubMed
Morales, F., Abadía, A. and Abadía, J. (1998). Photosynthesis, quenching of chlorophyll fluorescence and thermal energy dissipation in iron-deficient sugar beet leaves. Australian Journal of Plant Physiology, 25, 403–412.CrossRefGoogle Scholar
Morales, F., Abadía, A. and Abadía, J. (2006). Photoinhibition and photoprotection under nutrient defiencies, drought and salinity. In: Photoprotection, Photoinhibition, Gene regulation and Environment (eds Demmig-Adams, B., Adams, W.W. III, and Mattoo, A.K.), Springer, Dordrecht, Netherlands, pp 65–85.Google Scholar
Morales, F., Abadía, A., Abadía, J., et al. (2002). Trichomes and photosynthetic pigment composition changes: responses of Quercus ilex subsp ballota (Desf.) Samp. and Quercus coccifera L. to Mediterranean stress conditions. Trees, Structure and Function, 16, 504–510.CrossRefGoogle Scholar
Morales, F., Belkhodja, R., Abadía, A., et al. (1994). Iron deficiency induced changes in the photosynthetic pigment composition of field-grown pear (Pyrus communis L.) leaves. Plant, Cell and Environment, 17, 1153–1160.CrossRefGoogle Scholar
Morales, F., Belkhodja, R., Abadía, A., et al. (2000). Photosystem II efficiency and mechanisms of energy dissipation in iron-deficient, field-grown pear trees (Pyrus communis L.). Photosynthesis Research, 63, 9–21.CrossRefGoogle Scholar
Morales, F., Belkhodja, R., Goulas, Y., et al. (1999). Remote and near-contact chlorophyll fluorescence during photosynthetic induction in iron-deficient sugar beet leaves. Remote Sensing of the Environment, 69, 170–178.CrossRefGoogle Scholar
Morales, F., Cartelat, A., Álvarez-Fernández, A., et al. (2005). Time-resolved spectral studies of blue-green fluorescence of artichoke (Cynara cardunculus L. Var. Scolymus) leaves: identification of CGA as one of the major fluorophores and age-mediated changes. Journal of Agricultural and Food Chemistry, 53, 9668–9678.CrossRefGoogle Scholar
Morales, F., Cerovic, Z.G. and Moya, I. (1996). Time resolved blue-green fluorescence of sugar beet (Beta vulgaris L.) leaves. Spectroscopic evidence for the presence of ferulic acid as the main fluorophore of the epidermis. Biochimica et Biophysica Acta, 1273, 251–262.CrossRefGoogle Scholar
Morales, F., Moise, N., Quílez, R., et al. (2001). Iron deficiency interrupts energy transfer from a disconnected part of the antenna to the rest of photosystem II. Photosynthesis Research, 70, 207–220.CrossRefGoogle ScholarPubMed
Mordelet, P. (1993) Influence of tree shading on carbon assimilation of grass leaves in Lamto Savanna, Côte d’Ivoire. Acta Oecologica-International Journal of Ecology, 14, 119–127.Google Scholar
Mordelet, P. and Menaut, J.C. (1995). Influence of trees on aboveground production dynamics of grasses in a humid savanna. Journal of Vegetation Science, 6, 223–228.CrossRefGoogle Scholar
Morecroft, M.D. and Woodward, F.I. (1996). Experiments on the Causes of Altitudinal Differences in the Leaf Nutrient Contents, Size and δ13C of Alchemilla alpina. New Phytologist, 134, 471–479.CrossRefGoogle Scholar
Morecroft, M.D., Stokes, V.J. and Morison, J.I.L. (2003). Seasonal changes in the photosynthetic capacity of canopy oak (Quercus robur) leaves: the impact of slow development on annual carbon uptake. Int J Biometeorol, 47, 221–226.CrossRefGoogle ScholarPubMed
Moreira, A.G. (2000). Effects of fire protection on savanna structure in central Brazil. Journal of Biogeography, 27, 1021–1029.CrossRefGoogle Scholar
Moreno, J., Gracia-Murria, M.J. and Marin-Navarro, J. (2008). Redox modulation of Rubisco conformation and activity through its cysteine residues. Journal of Experimental Botany, 59, 1605–1614.CrossRefGoogle ScholarPubMed
Moreno, M.T., Riera, D., Carambula, C., et al. (2006). Estimation of water use efficiency in three cultivars of Dactylis glomerata L. under different soil water contents. 21st General Meeting of the European Grasslands Federation, 2–6 April, Badajoz (Spain).Google Scholar
Moreno-Sotomayor, M., Weiss, A., Paparozzi, E.T., et al. (2002). Stability of leaf anatomy and light response curves of field grown maize as a function of age and nitrogen status. Journal of Plant Physiology, 159, 819–826.CrossRefGoogle Scholar
Morgan, J.A., LeCain, D.R., Mosier, A.R., et al. (2001). Elevated CO2 enhances water relations and productivity and affects gas exchange in C3 and C4 grasses of the Colorado shortgrass steppe. Global Change Biology, 7, 451–466.CrossRefGoogle Scholar
Morgan, J.A., Milchunas, D.G., Lecain, D.R., et al. (2007). Carbon dioxide enrichment alters plant community structure and promotes shrub growth in the shortgrass steppe. Proceedings of the National Academy of Sciences, 104, 14,724–14,729.CrossRefGoogle Scholar
Morgan, J.A., Pataki, D.E., Korner, C., et al. (2004). Water relations in grassland and desert ecosystems exposed to elevated atmospheric CO2. Oecologia, 140, 11–25.CrossRefGoogle ScholarPubMed
Morgan, M.E., Kingston, J.D. and Marino, B.D. (1994). Carbon isotopic evidence for the emergence of C4 plants in the Neogene from Pakistan and Kenya. Nature, 367, 162–165.CrossRefGoogle Scholar
Morgan, P.B., Ainsworth, E.A. and Long, S.P. (2003). How does elevated ozone impact soybean? A meta-analysis of photosynthesis, growth and yield. Plant, Cell and Environment, 26, 1317–1328.CrossRefGoogle Scholar
Morgan, P.B., Bernacchi, C.J., Ort, D.R., et al. (2004). An in vivo analysis of the effect of season-long open-air elevation of ozone to anticipated 2050 levels on photosynthesis in soybean. Plant Physiology, 135, 2348–2357.CrossRefGoogle Scholar
Morgenstern, K., Black, T.A., Humpreys, E.R., et al. (2004). Sensitivity and uncertainty of the carbon balance of a Pacific Northwest Douglas-fir forest during an El Niño/La Niña cycle. Agricultural and Forest Meteorology, 123, 201–219.CrossRefGoogle Scholar
Morgenthal, K., Weckwerth, W. and Steuer, R. (2006). Metabolomic networks in plants: transitions from pattern recognition to biological interpretation. Biosystems, 83, 108–117.CrossRefGoogle ScholarPubMed
Moriana, A., Villalobos, F.J. and Fereres, E. (2002). Stomatal and photosynthetic responses of olive (Olea europaea L.) leaves to water deficits. Plant, Cell and Environment, 25, 395–405.CrossRefGoogle Scholar
Moriondo, M., Orlandini, S., Giuntoli, A., et al. (2005). The effect of downy and powdery mildew on grapevine (Vitis vinifera L.) leaf gas exchange. Journal of Phytopathology, 153, 350–357.CrossRefGoogle Scholar
Morison, J.I.L. (1985). Sensitivity of stomata and water use efficiency to high CO2. Plant, Cell and Environment, 8, 467–474.CrossRefGoogle Scholar
Morison, J.I.L. and Lawson, T. (2007). Does lateral gas diffusion in leaves matter?Plant, Cell and Environment, 30, 1072–1085.CrossRefGoogle ScholarPubMed
Morison, J.I.L., Baker, N.R., Mullineaux, P.M., et al. (2008). Improving water use efficiency in crop production. Philosophical transactions of the Royal Society, 363, 639–658.CrossRefGoogle Scholar
Morison, J.I.L., Gallouët, E., Lawson, T., et al. (2005). Lateral diffusion of CO2 in leaves is not sufficient to support photosynthesis. Plant Physiology, 139, 254–266.CrossRefGoogle Scholar
Morison, J.I.L., Lawson, T. and Cornic, G. (2007). Lateral CO2 diffusion inside dicotyledonous leaves can be substantial: Quantification in different light intensities. Plant Physiology, 145, 680–690.CrossRefGoogle ScholarPubMed
Morita, N. (1935). The increased density of air oxygen relative to water oxygen. Journal of the Chemical Society of Japan, 56, 1291.Google Scholar
Morley, R.J. (2000). The Origin and Evolution of Tropical Rain Forests. John Wiley and Sons, Chichester, UK and New York, USA.
Morley, R.J. and Richards, K. (1993). Gramineae cuticle: a key indicator of late Cenozoic climatic change in the Niger Delta. Review of Palaeobotany and Palynology, 77, 119–127.CrossRefGoogle Scholar
Morrison, M.J., Voldeng, H.D. and Cober, E.R. (1999). Physiological changes from 58 years of genetic improvement of short-season soybean cultivars in Canada. Agronomy Journal, 91, 685–689.CrossRefGoogle Scholar
Moser, W. (1973). Licht, Temperatur und Photosynthese an der Station “Hoher Nebelkogel” (3184m). In: Ökosystemforschung (ed. Ellenberg, H.) Springer, Berlin, Germany, pp. 203–223.
Mott, K.A. (1988). Do stomata respond to CO2 concentrations other than intercellular?Plant Physiology, 86, 200–203.CrossRefGoogle Scholar
Mott, K.A. (1995). Effects of patchy stomatal closure on gas exchange measurements following abscisic acid treatment. Plant, Cell and Environment, 18, 1291–1300.CrossRefGoogle Scholar
Mott, K.A. and Parkhurst, D.F. (1991). Stomatal responses to humidity in air and helox. Plant, Cell and Environment, 14, 509–515.CrossRefGoogle Scholar
Mott, K.A., Gibson, A.C. and O’Leary, J.W. (1982). The adaptive significance of amphistomatic leaves. Plant, Cell and Environment, 5, 455–460.CrossRefGoogle Scholar
Mott, K.A., Sibbernsen, E.D. and Shope, J.C. (2008). The role of the mesophyll in stomatal responses to light and CO2. Plant, Cell Environment, 31, 1299–1306.CrossRefGoogle ScholarPubMed
Moulin, M., McCormac, A.C., Terry, M.J., et al. (2008). Tetrapyrrole profiling in Arabidopsis seedlings reveals that retrograde plastid nuclear signalling is not due to mg-protoporphyrin IX accumulation. Proceedings of the National Academy of Sciences (USA), 105, 15178–15183.CrossRefGoogle Scholar
Moustakas, M. and Ouzounidou, G. (1994). Increased non-photochemical quenching in leaves of aluminum-stressed wheat plants is due to Al3+-induced elemental loss. Plant Physiology and Biochemistry, 32, 527–532.Google Scholar
Moustakas, M., Ouzounidou, G. and Lannoye, R. (1995). Aluminum effects on photosynthesis and elemental uptake in an aluminum-tolerant and non-tolerant wheat cultivar. Journal of Plant Nutrition, 18, 669–683.CrossRefGoogle Scholar
Moustakas, M., Ouzounidou, G., Symeonidis, L., et al. (1997). Field study of the effects of excess copper on wheat photosynthesis and productivity. Soil Science and Plant Nutrition, 43, 531–539.CrossRefGoogle Scholar
Moya, I. (1974). Durée de vie et rendement de fluorescence de la chlorophylle in vivo. Leur relation dans differents modeles d’unites photosynthètique. Biochimica et Biophysica Acta, 368, 214–227.CrossRefGoogle Scholar
Moya, I. and Cerovic, Z.G. (2004). Remote sensing of chlorophyll fluorescence : instrumentation and analysis. In: Chlorophyll Fluorescence: The Signature of Green Plant Photosynthesis (eds Papageorgiou, G.C. and Govindgee, J.), Kluwer, Dordrecht, Netherlands, pp. 429–445.Google Scholar
Moya, I., Camenen, L., Evain, S., et al. (2004). A new instrument for passive remote sensing: 1. Measurements of sunlight induced chlorophyll fluorescence. Remote Sens. Environ., 91, 186–197.CrossRefGoogle Scholar
Moya, I., Camenen, L., Latouche, G., et al. (1999). An instrument for the measurement of sunlight excited plant fluorescence. In: Photosynthesis: Mechanisms and Effects (ed. Garab, G.), Kluwer Academic Press, Dordrecht, Netherlands, pp. 4265–4270.Google Scholar
Moya, I., Cartelat, A., Cerovic, Z.G., et al. (2003). Possible approaches to remote sensing of photosynthetic activity. IEEE International Geoscience and Remote Sensing Symposium (IGARSS 2003) 21–25 July 2003. Toulouse, France.CrossRef
Moya, I., Daumard, F., Moise, N., et al. (2006). First airborne multiwavelength passive chlorophyll fluorescence measurements over La Mancha (Spain) fields. The 2nd International Symposium on Recent Advances in Quantitative Remote Sensing: RAQRS’II 25–29 September 2006. Torrent (Valencia)-Spain, C50 Torrent Laurent.Google Scholar
Moya, I., Goulas, Y., Morales, F., et al. (1995). Remote sensing of time-resolved chlorophyll fluorescence and back-scattering of the laser excitation by vegetation. EARSeL Advances in Remote Sensing, 3, 188–197.Google Scholar
Moya, I., Guyot, G. and Goulas, Y. (1992). Remotely sensed blue and red fluorescence emission for monitoring vegetation. ISPRS Journal of Photogrammetry and Remote Sensing, 47, 205–231.CrossRefGoogle Scholar
Moya, I., Sebban, P. and Haehnel, W. (1986). Lifetime of excited states and quantum yield of chlorophyll a fluorescence in vivo. In: Light Emission by Plants and Bacteria (eds Govindjee, J., Amesz, J. and Fork, D.C.), Academic Press, Orlando, USA, pp. 161–190.Google Scholar
Muchow, R.C. and Sinclair, T.R. (1994). Nitrogen response of leaf photosynthesis and canopy radiation use efficiency in field-grown maize and sorghum. Crop Science, 34, 721–727.CrossRefGoogle Scholar
Muhlenbock, P., Szechynska-Hebda, M., Plaszczyca, M., et al. (2008). Chloroplast signalling and LESION SIMULATING DISEASE1 regulate crosstalk between light acclimation and immunity in Arabidopsis. Plant Cell, 20, 2339–2356.CrossRefGoogle Scholar
Mulkey, S.S., Wright, S.J. and Smith, A.P. (1993). Comparative phsiology and demography of three Neotropical forest shrubs: alternative shad-adaptive character syndromes. Oecologia, 96, 526–536.CrossRefGoogle Scholar
Mullen, J.L., Weinig, C. and Hangarter, R.P. (2006). Shade avoidance and the regulation of leaf inclination in Arabidopsis. Plant Cell Environment, 29, 1099–1106.CrossRefGoogle ScholarPubMed
Mųller, C.M. (1946). Untersuchungen über Laubmenge, Stoffverlust und Stoffproduktion des Waldes.
Müller, D.J., Dencher, N.A., Meier, T., et al. (2001). ATP synthase: constrained stoichiometry of the transmembrane rotor. FEBS Letters, 504, 219–222.CrossRefGoogle ScholarPubMed
Müller, J., Eschenröder, A. and Diepenbrock, W. (2009). Through-flow chamber CO2/H2O canopy gas exchange system – construction, microclimate, errors, and measurements in a barley (Hordeum vulgare L.) field. Agricultural and Forest Meteorology, 149, 214–229.CrossRefGoogle Scholar
Müller, M., Zellnig, G., Tausz, M., et al. (1997). Structural changes and physiological stress response of spruce trees to SO2, O3 and elevated levels of CO2. In: Impact of Global Change on Tree Physiology and Forest Ecosystems (eds Mohren, G.M.J., Kramer, K. and Sabaté, S.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 93–102.Google Scholar
Müller, P., Li, X.P. and Niyogi, K.K. (2001). Non-photochemical quenching. A response to excess light energy. Plant Physiology 125, 1558–1566.CrossRefGoogle ScholarPubMed
Mulligan, D.R. (1989). Leaf phosphorus and nitrogen concentrations and net photosynthesis in Eucalyptus seedlings. Tree Physiology, 5, 149–157.CrossRefGoogle ScholarPubMed
Mullin, L.P., Sillett, S.C., Koch, G.W., et al. (2009). Physiological consequences of height-related morphological variation in Sequoia sempervirens foliage. Tree Physiology, 29, 999–1010.CrossRefGoogle ScholarPubMed
Mullineaux, P. and Karpinski, S. (2002). Signal transduction in response to excess light: getting out of the chloroplast. Current Opinion in Plant Biology, 5, 43–48.CrossRefGoogle ScholarPubMed
Mulroy, T.W. and Rundel, P.W. (1977). Annual plants: adaptations to desert environments. BioScience, 27, 109–114.CrossRefGoogle Scholar
Munné-Bosch, S. (2007). Ageing in perennials. Critical Reviews in Plant Sciences, 26, 123–138.CrossRefGoogle Scholar
Munné-Bosch, S. (2008). Do perennials really senesce?Trends in Plant Science, 13, 216–220.CrossRefGoogle ScholarPubMed
Munné-Bosch, S., Alegre, L. ( 2000). Changes in carotenoids, tocopherols and diterpenes during drought and recovery, and the biological significance of chlorophyll loss in Rosmarinus officinalis plants. Planta, 210, 925–931.CrossRefGoogle ScholarPubMed
Munné-Bosch, S. and Alegre, L. (2002). Plant aging increases oxidative stress in chloroplasts. Planta, 214, 608–615.Google ScholarPubMed
Munné-Bosch, S. and Alegre, L. (2004). Die and let live: leaf senescence contributes to plant survival under drought stress. Functional Plant Biology, 31, 203–216.CrossRefGoogle Scholar
Munné-Bosch, S. and Peñuelas, J. (2003). Photo- and antioxidative protection during summer leaf senescence in Pistacia lentiscus L. grown under Mediterranean field conditions. Annals of Botany, 92, 385–391.CrossRefGoogle ScholarPubMed
Munné-Bosch, S., Jubany Marí, T. and Alegre, L. (2001). Drought-induced senescence is characterized by a loss of antioxidant defences in chloroplasts. Plant Cell Environment, 24, 1319–1327.CrossRefGoogle Scholar
Munné-Bosch, S., Nogues, S. and Alegre, L. (1999). Diurnal variations of photosynthesis and dew absorption by leaves in two evergreen shrubs growing in Mediterranean field conditions. New Phytologist, 144, 109–119.CrossRefGoogle Scholar
Munné-Bosch, S., Peñuelas, J., Asensio, D., et al. (2004). Airborne ethylene may alter antioxidant protection and reduce tolerance of holm oak to heat and drought stress. Plant Physiology, 136, 2937–2947.CrossRefGoogle ScholarPubMed
Munns, R. (2002). Comparative physiology of salt and water stress. Plant, Cell and Environment, 25, 239–250.CrossRefGoogle ScholarPubMed
Munns, R. and Tester, M. (2008). Mechanisms of salinity tolerance. Annual Review of Plant Biology, 59, 651–681.CrossRefGoogle ScholarPubMed
Munns, R., James, R.A. and Läuchli, A. (2006). Approaches to increasing the salt tolerance of wheat and other cereals. Journal of Experimental Botany, 57, 1025–1043.CrossRefGoogle ScholarPubMed
Murata, N. and Los, D.A. (1997). Membrane fluidity and temperature perception. Plant Physiology, 115, 875–879.CrossRefGoogle ScholarPubMed
Murata, N. and Los, D.A. (2006). Histidine kinase Hik33 is an important participant in cold-signal transduction in cyanobacteria. Physiologia Plantarum, 126, 17–27.CrossRefGoogle Scholar
Murata, N., Takahashi, S., Nishiyama, Y., et al. (2007). Photoinhibition of photosystem II under environmental stress. Biochimica et Biophysica Acta – Bioenergetics, 1767, 414–421.CrossRefGoogle ScholarPubMed
Murchie, E.H., Pinto, M. and Horton, P. (2009). Agriculture and the new challenges for photosynthesis research. The New Phytologist, 181, 532–552.CrossRefGoogle ScholarPubMed
Murphy, R. and Smith, J.A.C. (1998). Determination of cell water-relation parameters using the pressure-clamp technique. Plant, Cell and Environment, 21, 637–657.CrossRefGoogle Scholar
Murray, M.B., Smith, R.I., Leith, I.D., et al. (1994). Effects of elevated CO2, nutrition and climatic warming on bud phenology in Sitka spruce (Picea sitchensis) and their impact on the risk of frost damage. Tree Physiology, 14, 691–706.CrossRefGoogle Scholar
Murthy, R., Barron-Gafford, G.A., Dougherty, P.M., et al. (2005). Increased leaf area dominates carbon flux response to elevated CO2 in stands of Populus deltoides (Bartr.) and underlies a switch from canopy light-limited CO2 influx in well-watered treatments to individual leaf, stomatally limited influx under water stress. Global Change Biology, 11, 716–731.CrossRefGoogle Scholar
Mustilli, A.C., Merlot, S., Vavasseur, A., et al. (2002). Arabidopsis OST1 protein kinase mediates the regulation of stomatal aperture by abscisic acid and acts upstream of reactive oxygen species production. The Plant Cell, 14, 3089–3099.CrossRefGoogle ScholarPubMed
Myers, B.A., Duff, G.A., Eamus, D., et al. (1997). Seasonal variation in water relations of trees of differing leaf phenology in a wet-dry tropical savanna near Darwin, northern Australia. Australian Journal of Botany, 45, 225–240.CrossRefGoogle Scholar
Myers, J.A. and Kitajima, K. (2007). Carbohydrate storage enhances seedling shade and stress tolerance in a neotropical forest. Journal of Ecology, 95, 383–395.CrossRefGoogle Scholar
Myneni, R.B., Hoffman, S., Knyazikhin, Y., et al. (2002). Global products of vegetation leaf area and fraction absorbed PAR from one year of MODIS data. Remote Sensing of Environment, 76, 139–155.Google Scholar
Nabity, P.D., Zavala, J.A. and DeLucia, E.H. (2009). Indirect suppression of photosynthesis on individual leaves by arthropod herbivory. Annals of botany, 103, 655–663.CrossRefGoogle ScholarPubMed
Naidu, R.A., Krishnan, M., Nayudu, M.V., et al. (1986). Studies on Peanut green mosaic virus infected peanut (Arachis hypogaea L.) leaves. III. Changes in the polypeptides of photosystem II particles. Physiological and Molecular Plant Pathology, 29, 53–58.CrossRefGoogle Scholar
Naidu, S.L. and DeLucia, E.H. (1997). Acclimation of shade-developed leaves on saplings exposed to late-season canopy gaps. Tree Physiology, 17, 367–376.CrossRefGoogle ScholarPubMed
Nainanayake, A.D. (2004). Impact of drought on coconut (Cocos nucifera L.): screening germplasm for photosynthetic tolerance in the field. Ph.D. thesis, University of Essex, pp. 215.
Nakano, R., Ishida, H., Makino, A., et al. (2006). In vivo fragmentation of the large subunit of ribulose-1,5-bisphosphate carboxylase by reactive oxygen species in an intact leaf of cucumber under chilling-light conditions. Plant and Cell Physiology, 47, 270–276.CrossRefGoogle Scholar
Nalborczyk, E. (1978). Dark carboxylation and its possible effect on the value of δ13C in C3 plants. Acta Physiologiae Plantarum, 1, 53–58.Google Scholar
Nali, C., Guidi, L., Ciompi, S., et al. (1995). Photosynthesis of Medicago sativa L. plants exposed to long-term fumigation with sulphur dioxide. In: Responses of Plants to Air Pollution (eds Lorenzini, G. and Soldatini, G.F.), Pacini Editore, Pisa, Italy, pp. 82–89.Google Scholar
Nambudiri, E.M.V., Tidwell, W.D., Smith, B.N., et al. (1978). A C4 plant from the Pliocene. Nature, 276, 816–817.CrossRefGoogle Scholar
Nardini, A., Gortan, E., Ramani, M., et al. (2008). Heterogeneity of gas exchange rates over the leaf surface in tobacco: an effect of hydraulic architecture?Plant, Cell and Environment, 31, 804–812.CrossRefGoogle ScholarPubMed
Nasyrov, Y.S. (1978). Genetic control of photosynthesis and improving of crop productivity. Ann. Rev. Plant Physiol., 29, 215–37.CrossRefGoogle Scholar
Naumann, J.C., Young, D.R. and Anderson, J.E. (2008). Leaf chlorophyll fluorescence, reflectance, and physiological response to freshwater and saltwater flooding in the evergreen shrub, Myrica cerifera. Environmental and Experimental Botany, 63, 402–409.CrossRefGoogle Scholar
Naumburg, E. and Ellsworth, D.S. (2000). Photosynthesis sunfleck utilization potential of understory saplings growing under elevated CO2 in FACE. Oecologia, 122, 163–174.CrossRefGoogle ScholarPubMed
Naumburg, E., Houseman, D.C., Huxman, T.E., et al. (2003). Photosynthetic responses of Mojave Desert shrubs to free air CO2 enrichment are greatest during wet years. Global Change Biology, 9, 276–285.CrossRefGoogle Scholar
Naumburg, E., Loik, M.E. and Smith, S.D. (2004). Photosynthetic responses of Larrea tridentata to seasonal temperature extremes under elevated CO2. New Phytologist, 162, 323–330.CrossRefGoogle Scholar
Neals, T.F. and Incoll, L.D. (1968). The control of leaf photosynthesis rate by the level of assimilate concentration in the leaf: a review of hypotheses. Botanical Review, 34, 107–125.CrossRefGoogle Scholar
Nebdal, L., Soukupová, J., Whitmarsh, J., et al. (2000a). Postharvest imaging of chlorophyll fluorescence from lemons can be used to predict fruit quality. Photosynthetica, 38, 571–579.Google Scholar
Nedbal, L. and Březina, V. (2002). Complex metabolic oscillations in plants forced by harmonic irradiance. Biophysical Journal, 83, 2180–2189.CrossRefGoogle ScholarPubMed
Nedbal, L. and Whitmarsh, J. (2004). Chlorophyll fluorescence imaging of leaves and fruits. In: Chlorophyll a Fluorescence: A Signature of Photosynthesis. Advances in photosynthesis and respiration, vol. 19. (eds Papageorgiou, C.G. and Govindjee, J.), Springer, Dordrecht, Netherlands, pp. 389–407.Google Scholar
Nedbal, L., Soukupová, J., Kaftan, D., et al. (2000b). Kinetic imaging of chlorophyll fluorescence using modulated light. Photosynthesis Research, 66, 25–34.CrossRefGoogle ScholarPubMed
Neftel, A., Friedli, H., Moor, E., et al. (1994). Historical CO2 record from the Siple Station ice core. In: Trends: A Compendium of Data on Global Change. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn., U.S.A.
Neill, S., Barros, R., Bright, J., et al. (2008). Nitric oxide, stomatal closure, and abiotic stress. Journal of Experimental Botany, 59, 165–176.CrossRefGoogle ScholarPubMed
Neill, S.O., Gould, K.S., Kilmartin, P.A., et al. (2002). Antioxidant activities of red versus green leaves in Elatostema rugosum. Plant Cell Environment, 25, 539–547.CrossRefGoogle Scholar
Nelson, D.E., Repetti, P.P., Adams, T.R., et al. (2007). Plant nuclear factor Y (NF-Y) B subunits confer drought tolerance and lead to improved corn yields on water-limited acres. Proceedings of the National Academy of Sciences USA, 104, 16450–16455.CrossRefGoogle ScholarPubMed
Nelson, E.A. and Sage, R.F. (2008). Functional constraints of CAM leaf anatomy: tight cell packing is associated with increased CAM function across a gradient of CAM expression. Journal of Experimental Botany, 59, 1841–1850.CrossRefGoogle ScholarPubMed
Nelson, E.A. and Sage, R.F. (2005). Functional leaf anatomy of plants with crassulacean acid metabolism. Functional Plant Biology, 32, 409–419.CrossRefGoogle Scholar
Nelson, J.A., Morgan, J.A., LeCain, D.R., et al. (2004). Elevated CO2 increases soil moisture and enhances plant water relations in a long-term field study in semi-arid shortgrass steppe of Colorado. Plant and Soil, 259, 169–179.CrossRefGoogle Scholar
Nelson, N., Sacher, A. and Nelson, H. (2002). The significance of molecular slips in transport systems. National Reviews of Molecular and Cellular Biology, 3, 876–881.CrossRefGoogle ScholarPubMed
Nelson, S.V. (2005). Paleoseasonality inferred from equid teeth and intra-tooth isotopic variability. Palaeogeography, Palaeoclimatology, Palaecoecology, 222, 122–144.CrossRefGoogle Scholar
Nemani, R.R., Keeling, C.D., Hashimoto, H., et al. (2003). Climate-driven increases in global terrestrial net primary production from 1982 to 1999. Science, 300, 1560–1563.CrossRefGoogle ScholarPubMed
Neubauer, C. and Schreiber, U. (1987). The polyphasic rise of chlorophyll fluorescence upon onset of strong continuous illumination. I. Saturation characteristics and partial control by the photosystem II acceptor side. Zeitschrift für Naturforschung, 42c, 1246–1254.Google Scholar
Neufeld, H.S., Meinzer, F.C., Wisdom, C.S., et al. (1988). Canopy architecture of Larrea tridentata (DC.) Cov., a desert shrub: foliage orientation and direct beam radiation interception. Oecologia, 75, 54–60.CrossRefGoogle ScholarPubMed
Neuner, G. and Pramsohler, M. (2006). Freezing and high temperature thresholds of photosystem 2 compared to ice nucleation, frost and heat damage in evergreen subalpine plants. Physiologia Plantarum, 126, 196–204.CrossRefGoogle Scholar
Neuner, G., Braun, V., Buchner, O., et al. (1999). Leaf rosette closure in the alpine rock species Saxifraga paniculata Mill.: significance for survival of drought and heat under high irradiation. Plant Cell and Environment, 22, 1539–1548.CrossRefGoogle Scholar
Newell, E.A., McDonald, E.P., Strain, B.R., et al. (1993). Photosynthetic responses of Miconia species to canopy openings in a lowland tropical rainforest. Oecologia, 94, 49–56.CrossRefGoogle Scholar
Newman, S.M. and Cattolico, R.A. (1990). Ribulose bisphosphate carboxylase in algae: synthesis, enzymology and evolution. Photosynthesis Research, 26, 69–85.CrossRefGoogle ScholarPubMed
Ngugi, M.R., Doley, D., Hunt, M.A., et al. (2004). Physiological responses to water stress in Eucalyptus cloeziana and E. argophioia seedlings. Trees-Structure and Function, 18, 381–389.CrossRefGoogle Scholar
Nichol, C.J., Huemmrich, K.F., Black, T.A., et al. (2000). Remote sensing of photosynthetic-light-use efficiency of boreal forest. Agricultural and Forest Meteorology, 101, 131–142.CrossRefGoogle Scholar
Nichol, C.J., Lloyd, J., Shibistova, O., et al. (2002). Remote sensing of photosynthetic-light-use efficiency of a Siberian boreal forest. Tellus, 54B, 677–687.CrossRefGoogle Scholar
Nichol, C.J., Rascher, U., Matsubara, S., et al. (2006). Assessing photosynthetic efficiency in an experimental mangrove canopy using remote sensing and chlorophyll fluorescence. Trees, 20, 9–15.CrossRefGoogle Scholar
Nickrent, D.L., Parkinson, C.L., Palmer, J.D., et al. (2000). Multigene phylogeny of land plants with special reference to Bryophytes and the earliest land plants. Molecular Biology and Evolution, 17, 1885–1895.CrossRefGoogle ScholarPubMed
Nicolás, E., Torrecillas, A., Dell’Amico, J., et al. (2005). The effect of short-term flooding on the sap flow, gas exchange and hydraulic conductivity of young apricot trees. Trees, 19, 51–57.CrossRefGoogle Scholar
Nicotra, A., Chazdon, R.L. and Montgomery, R.A. (2003). Sexes show contrasting patterns of leaf and crown carbon gain in a dioecious rainforest shrub. American Journal of Botany, 90, 347–355.CrossRefGoogle Scholar
Nicotra, A.B. and Davidson, A. (2010). Adaptive phenotypic plasticity and plant water use. Functional Plant Biology, 37, 117–127.CrossRefGoogle Scholar
Nicotra, A.B., Cosgrove, M.J., Cowling, A., et al. (2008). Leaf shape linked to photosynthetic rates and temperature optima in South African Pelargonium species. Oecologia, 154, 625–635.CrossRefGoogle ScholarPubMed
Niederleitner, S. and Knoppik, D. (1997). Effects of the cherry leaf spot pathogen Blumeriella jaapii on gas exchange before and after expression of symptoms on cherry leaves. Physiological and Molecular Plant Pathology, 51, 145–153.CrossRefGoogle Scholar
Niemann, G.J., van der Kerk, A., Niessen, W.M.A., et al. (1991). Free and cell wall-bound phenolics and other constituents from healthy and fungus-infected carnation (Dianthus caryophyllus L.) stems. Physiological and Molecular Plant Pathology, 38, 417–432.CrossRefGoogle Scholar
Nier, A.O. (1940). A mass spectrometer for routine isotope abundance measurements. Review of Scientific Instruments, 11, 212–216.CrossRefGoogle Scholar
Nier, A.O. and Gulbransen, E.A. (1939). Variations in the relative abundance of the carbon isotopes. Journal of the American Chemical Society, 61, 697–698.CrossRefGoogle Scholar
Nihlgård, B. (1972). Plant biomass, primary production and distribution of chemical elements in a beech and a planted spruce forest in South Sweden. Oikos, 23, 69–81.CrossRefGoogle Scholar
Niinemets, Ü. (1997a). Acclimation to low irradiance in Picea abies: influences of past and present light climate on foliage structure and function. Tree Physiology, 17, 723–732.CrossRefGoogle ScholarPubMed
Niinemets, Ü. (1997b). Distribution patterns of foliar carbon and nitrogen as affected by tree dimensions and relative light conditions in the canopy of Picea abies. Trees, 11, 144–154.Google Scholar
Niinemets, Ü. (1997c). Energy requirement for foliage construction depends on tree size in young Picea abies trees. Trees, 11, 420–431.Google Scholar
Niinemets, Ü. (1998). Growth of young trees of Acer platanoides and Quercus robur along a gap – understory continuum: interrelationships between allometry, biomass partitioning, nitrogen, and shade-tolerance. International Journal of Plant Science, 159, 318–330.CrossRefGoogle Scholar
Niinemets, Ü. (1999). Components of leaf dry mass per area – thickness and density – alter leaf photosynthetic capacity in reverse directions in woody plants. The New Phytologist, 144, 35–47.CrossRefGoogle Scholar
Niinemets, Ü. (2001). Global-scale climatic controls of leaf dry mass per area, density, and thickness in trees and shrubs. Ecology, 82, 453–469.CrossRefGoogle Scholar
Niinemets, Ü. (2002). Stomatal conductance alone does not explain the decline in foliar photosynthetic rates with increasing tree age and size in Picea abies and Pinus sylvestris. Tree Physiology, 22, 515–535.CrossRefGoogle Scholar
Niinemets, Ü. (2004). Adaptive adjustments to light in foliage and whole-plant characteristics depend on relative age in the perennial herb Leontodon hispidus. New Phytologist, 162, 683–696.CrossRefGoogle Scholar
Niinemets, Ü. (2006). The controversy over traits conferring shade-tolerance in trees: ontogenetic changes revisited. Journal of Ecology, 94, 464–470.CrossRefGoogle Scholar
Niinemets, Ü. (2007). Photosynthesis and resource distribution through plant canopies. Plant, Cell and Environment, 30, 1052–1071.CrossRefGoogle ScholarPubMed
Niinemets, Ü. (2009). Light interception in plant stands from leaf to canopy in different plant functional types and in species with varying shade tolerance: a review. Ecological Research, 25, 693–714.CrossRefGoogle Scholar
Niinemets, Ü. and Anten, N.P.R. (2009). Packing photosynthesis machinery: from leaf to canopy. In: Photosynthesis in Silico: Understanding Complexity from Molecules to Ecosystems (eds Laisk, A., Nedbal, L. and Govindjee, J.), Springer-Verlag, Berlin, pp. 363–399.Google Scholar
Niinemets, Ü. and Kull, K. (1994). Leaf weight per area and leaf size of 85 Estonian woody species in relation to shade tolerance and light availability. Forest Ecology and Management, 70, 1–10.CrossRefGoogle Scholar
Niinemets, Ü. and Kull, O. (1995a). Effects of light availability and tree size on the architecture of assimilative surface in the canopy of Picea abies: variation in shoot structure. Tree Physiology, 15, 791–798.CrossRefGoogle Scholar
Niinemets, Ü. and Kull, O. (1995b). Effects of light availability and tree size on the architecture of assimilative surface in the canopy of Picea abies: variation in needle morphology. Tree Physiology, 15, 307–315.CrossRefGoogle ScholarPubMed
Niinemets, Ü. and Kull, O. (1998). Stoichiometry of foliar carbon constituents varies along light gradients in temperate woody canopies: implications for foliage morphological plasticity. Tree Physiology, 18, 467–479.CrossRefGoogle ScholarPubMed
Niinemets, Ü. and Sack, L. (2006). Structural determinants of leaf light-harvesting capacity and photosynthetic potentials. In: Progress in Botany, vol. 67 (eds Esser K., Lüttge, U.E., Beyschlag, W. and Murata, J.), Springer Verlag, Berlin, pp. 385–419.CrossRef
Niinemets, Ü. and Tamm, Ü. (2005). Species differences in timing of leaf fall and foliage chemistry modify nutrient resorption efficiency in deciduous temperate forest stands. Tree Physiology, 25, 1001–1014.CrossRefGoogle ScholarPubMed
Niinemets, Ü. and Tenhunen, J.D. (1997). A model separating leaf structural and physiological effects on carbon gain along light gradients for the shade-tolerant species Acer saccharum. Plant, Cell and Environment, 20, 845–866.CrossRefGoogle Scholar
Niinemets, Ü. and Valladares, F. (2004). Photosynthetic acclimation to simultaneous and interacting environmental stresses along natural light gradients: optimality and constraints. Plant Biology, 6, 254–268.CrossRefGoogle ScholarPubMed
Niinemets, Ü. and Valladares, F. (2006). Tolerance to shade, drought, and waterlogging of temperate northern hemisphere trees and shrubs. Ecological Monographs, 76, 521–547.CrossRefGoogle Scholar
Niinemets, Ü., Cescatti, A., Rodeghiero, M., et al. (2005a). Leaf internal diffusion conductance limits photosynthesis more strongly in older leaves of Mediterranean evergreen broad-leaved species. Plant, Cell and Environment, 28, 1552–1566.CrossRefGoogle Scholar
Niinemets, Ü., Cescatti, A., Rodeghiero, M., et al. (2006a). Complex adjustments of photosynthetic capacity and internal mesophyll conductance to current and previous light availabilities and leaf age in Mediterranean evergreen species Quercus ilex. Plant, Cell and Environment, 29, 1159–1178.CrossRefGoogle Scholar
Niinemets, Ü., Diaz-Espejo, A., Flexas, J., et al. (2009a). Importance of mesophyll-diffusion conductance in estimation of plant photosynthesis in the field. Journal of Experimental Botany, 60, 2271–2282.CrossRefGoogle ScholarPubMed
Niinemets, Ü., Díaz-Espejo, A., Flexas, J., et al. (2009b). Role of mesophyll-diffusion conductance in constraining potential photosynthetic productivity in the field. Journal of Experimental Botany, 60, 2249–2270.CrossRefGoogle ScholarPubMed
Niinemets, Ü., Ellsworth, D.S., Lukjanova, A., et al. (2001). Site fertility and the morphological and photosynthetic acclimation of Pinus sylvestris needles to light. Tree Physiology, 21, 1231–1244.CrossRefGoogle Scholar
Niinemets, Ü., Flexas, J. and Peñuelas, J. (2011). Evergreens favoured by higher responsiveness to increased CO2. Trends in Ecology and Evolution, 26, 136–142.CrossRefGoogle Scholar
Niinemets, Ü., Hauff, K., Bertin, N., et al. (2002). Monoterpene emissions in relation to foliar photosynthetic and structural variables in Mediterranean evergreen Quercus species. New Phytologist, 153, 243–256.CrossRefGoogle Scholar
Niinemets, Ü., Kollist, H., García-Plazaola, J.I., et al. (2003). Do the capacity and kinetics for modification of xanthophyll cycle pool size depend on growth irradiance in temperate trees?Plant, Cell and Environment, 26, 1787–1801.CrossRefGoogle Scholar
Niinemets, Ü. and Kull, K. (2005). Co-limitation of plant primary productivity by nitrogen and phosphorus in a species-rich wooded meadow on calcareous soils. Acta Oecologica, 28, 345–356.CrossRefGoogle Scholar
Niinemets, Ü. and Kull, O. (2001). Sensitivity to photoinhibition of photosynthetic electron transport in a temperate deciduous forest canopy: photosystem II centre openness, non-radiative energy dissipation and excess irradiance under field conditions. Tree Physiology, 21, 899–914.CrossRefGoogle Scholar
Niinemets, Ü., Kull, O. and Tenhunen, J.D. (1998). An analysis of light effects on foliar morphology, physiology, and light interception in temperate deciduous woody species of contrasting shade tolerance. Tree Physiology, 18, 681–696.CrossRefGoogle ScholarPubMed
Niinemets, Ü., Kull, O. and Tenhunen, J.D. (1999a). Variability in leaf morphology and chemical composition as a function of canopy light environment in co-existing trees. International Journal of Plant Sciences, 160, 837–848.CrossRefGoogle Scholar
Niinemets, Ü., Kull, O. and Tenhunen, J.D. (2004a). Within canopy variation in the rate of development of photosynthetic capacity is proportional to integrated quantum flux density in temperate deciduous trees. Plant, Cell and Environment, 27, 293–313.CrossRefGoogle Scholar
Niinemets, Ü., Lukjanova, A., Sparrrow, A.D., et al. (2005c). Light-acclimation of cladode photosynthetic potentials in Casuarina glauca: trade-offs between physiological and structural investments. Functional Plant Biology, 32, 571–582.CrossRefGoogle Scholar
Niinemets, Ü., Oja, V. and Kull, O. (1999b). Shape of leaf photosynthetic electron transport versus temperature response curve is not constant along canopy light gradients in temperate deciduous trees. Plant, Cell and Environment, 22, 1497–1514.CrossRefGoogle Scholar
Niinemets, Ü., Portsmuth, A., Tena, D., et al. (2007). Do we underestimate the importance of leaf size in plant economics? Disproportionate scaling of support costs within the spectrum of leaf physiognomy. Annals of Botany, 100, 283–303.CrossRefGoogle Scholar
Niinemets, Ü., Sõber, A., Kull, O., et al. (1999d). Apparent controls on leaf conductance by soil water availability and via light-acclimation of foliage structural and physiological properties in a mixed deciduous, temperate forest. International Journal of Plant Sciences, 160, 707–721.CrossRefGoogle Scholar
Niinemets, Ü., Sonninen, E. and Tobias, M. (2004c). Canopy gradients in leaf intercellular CO2 mole fractions revisited: interactions between leaf irradiance and water stress need consideration. Plant, Cell and Environment, 27, 569–583.CrossRefGoogle Scholar
Niinemets, Ü., Tenhunen, J.D. and Beyschlag, W. (2004b). Spatial and age-dependent modifications of photosynthetic capacity in four Mediterranean oak species. Functional Plant Biology, 31, 1179–1193.CrossRefGoogle Scholar
Niinemets, U., Tenhunen, J.D., Canta, N.R., et al. (1999c). Interactive effects of nitrogen and phosphorus on the acclimation potential of foliage photosynthetic properties of cork oak, Quercus suber, to elevated atmospheric CO2concentrations. Global Change Biology, 5, 455–470.CrossRefGoogle Scholar
Niinemets, Ü., Tobias, M., Cescatti, A., et al. (2006b). Size-dependent variation in shoot light-harvesting efficiency in shade-intolerant conifers. International Journal of Plant Science, 167, 19–32.CrossRefGoogle Scholar
Niinemets, Ü., Wright, I.J. and Evans, J.R. (2009c). Leaf diffusion conductance in 35 Australian sclerophylls covering a broad range of foliage structural and physiological variation. Journal of Experimental Botany, 60, 2433–2449.CrossRefGoogle ScholarPubMed
Niklas, K.J. (2001). Invariant scaling relationships for interspecific plant biomass production rates and body size. Proceedings of the National Academy of Sciences USA, 98, 2922–2927.CrossRefGoogle ScholarPubMed
Niklas, K.J. and Kutschera, U. (2009a). The evolution of the land plant lifecycle. New Phytologist, 185, 27–41.CrossRefGoogle Scholar
Niklas, K.J. and Kutschera, U. (2009b). The evolutionary development of plant body plans. Functional Plant Biology, 36, 682–695.CrossRefGoogle Scholar
Nikolopoulos, D., Liakopoulos, G., Drossopoulos, I., et al. (2002). The relationship between anatomy and photosynthetic performance of heterobaric leaves. Plant Physiology, 129, 235–243.CrossRefGoogle ScholarPubMed
Nilsen, E.T. (1992). The influence of water stress on leaf and stem photosynthesis in Spartium junceum L. Plant, Cell and Environment, 15, 455–461.CrossRefGoogle Scholar
Nilsen, E.T. and Sharifi, M.R. (1997). Carbon isotopic composition of legumes with photosynthetic stems from Mediterranean and desert habitats. American Journal of Botany, 84, 1707–13.CrossRefGoogle ScholarPubMed
Nilsen, E.T., Meinzer, F.C. and Rundel, P.W. (1989). Stem photosynthesis in Psorothamnus spinosus (smoke tree) in the Sonoran desert of California. Oecologia, 79, 193–197.CrossRefGoogle ScholarPubMed
Nilsen, E.T., Rundel, P.W. and Sharifi, M.R. (1996). Diurnal gas exchange characteristics of two stem photosynthesizing legumes in relation to the climate at two contrasting sites in the California desert. Flora, 191, 105–16.CrossRefGoogle Scholar
Nilson, S.E. and Assmann, S.M. (2007). The control of transpiration. Insights from Arabidopsis. Plant Physiology, 143, 19–27.CrossRefGoogle ScholarPubMed
Nilson, T. (1971). A theoretical analysis of the frequency of gaps in plant stands. Agr Meteorol, 8, 25–38.CrossRefGoogle Scholar
Nilsson, H.E. (1995). Remote sensing and image analysis in plant pathology. Annual Review of Phytopathology, 15, 489–527.CrossRefGoogle Scholar
Ning, L., Edwards, G.E., Strobel, G.A., et al. (1995). Imaging fluorometer to detect pathological and physiological change in plants. Applied Spectroscopy, 49, 1381–1389.CrossRefGoogle Scholar
Nippert, J.B., Fay, P.A. and Knapp, A.K. (2007). Photosynthetic traits in C3 and C4 grassland species in mesocosm and field environments. Environmental and Experimental Botany, 60, 412–420.CrossRefGoogle Scholar
Nisbet, E.G., Grassineau, N.V., Howe, C.J., et al. (2007). The age of Rubisco: the evolution of oxygenic photosynthesis. Geobiology, 5, 311–335.CrossRefGoogle Scholar
Nishida, I. and Murata, N. (1996). Chilling sensitivity in plants and cyanobacteria: the crucial contribution of membrane lipids. Annual Review of Plant Physiology and Plant Molecular Biology, 47, 541–568.CrossRefGoogle ScholarPubMed
Nishio, J.N. (2000). Why are higher plants green? Evolution of the higher plant photosynthetic pigment complement. Plant, Cell and Environment, 23, 539–548.CrossRefGoogle Scholar
Nishio, J.N. and Whitmarsh, J. (1993). Dissipation of the proton electrochemical potential in intact chloroplasts. II. The pH gradient monitored by cytochrome f reduction kinetics. Plant Physiology, 101, 89–96.CrossRefGoogle Scholar
Nishio, J.N., Sun, J. and Vogelmann, T.C. (1993). Carbon fixation gradients across spinach leaves do not follow internal light gradients. Plant Cell, 5, 953–961.CrossRefGoogle Scholar
Nitta, I. and Ohsawa, M. (1997). Leaf dynamics and shoot phenology of eleven warm-temperate evergreen broad-leaved trees near their northern limit in central Japan. Plant Ecology, 130, 71–88.CrossRefGoogle Scholar
Nittylä, T., Messerli, G., Trevisan, M., et al. (2004). A novel maltose transporter is essential for starch degradation in leaves. Science, 303, 87–89.CrossRefGoogle Scholar
Niyogi, K.K. (1999). Photoprotection revisited: genetic and molecular approaches. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 333–359.CrossRefGoogle ScholarPubMed
Niyogi, K.K. (2000). Safety valves for photosynthesis. Current Opinion in Plant Biology, 3, 455–460.CrossRefGoogle ScholarPubMed
Niyogi, K.K., Li, X.P., Rosenberg, V., et al.(2005). Is PsbS the site of non-photochemical quenching in photosynthesis?Journal of Experimental Botany, 56, 375–382.CrossRefGoogle ScholarPubMed
Nobel, P.S. ( 1977). Internal leaf area and cellular CO2 resistance: photosynthetic implications of variations with growth conditions and plant species. Physiologia Plantarum, 40, 137–144.CrossRefGoogle Scholar
Nobel, P.S. (1978). Microhabitat, water relations, and photosynthesis of a desert fern, Notholaena parryi. Oecologia, 31, 293–309.CrossRefGoogle ScholarPubMed
Nobel, P.S. (1980). Water vapor conductance and CO2 uptake for leaves of a C4 desert grass, Hilaria rigida. Ecology, 61, 252–258.CrossRefGoogle Scholar
Nobel, P.S. (1982). Orientation of terminal cladodes of platyopuntias. Botanical Gazette, 143, 219–24.CrossRefGoogle Scholar
Nobel, P.S. (1983). Nutrient levels in cacti – relation to nocturnal acid accumulation and growth. American Journal of Botany, 70, 1244–1253.CrossRefGoogle Scholar
Nobel, P.S. (1988). Environmental biology of Agaves and cacti. Cambridge University Press, Cambridge, UK.
Nobel, P.S. (1991). Achievable productivities of CAM plants; basis for high values compared with C3, and C4. New Phytologist, 119, 183–205.CrossRefGoogle Scholar
Nobel, P.S. (1991b). Physicochemical and environmental plant physiology. Academic Press, San Diego, USA.
Nobel, P.S. (1994). Remarkable Agaves and Cacti. Oxford University Press, New York, USA.Google Scholar
Nobel, P.S. (1996). High productivity of certain agronomic CAM species. In: Winter, K., Smith, J.A.C., eds. Crassulacean Acid Metabolism. Biochemistry, Ecophysiology and Evolution. Ecological studies, Vol. 145. Springer-Verlag, Berlin – Heidelberg – New York, 255–265.Google Scholar
Nobel, P.S. (2005). Physicochemical and Environmental Plant Physiology, 3rd edn. Academic Press, San Diego, USA.Google Scholar
Nobel, P.S. and Berry, W.L. (1985). Element responses of agaves. American Journal of Botany 72, 686–694.CrossRefGoogle Scholar
Nobel, P.S. and Smith, S.D. (1983). High and low temperature tolerances and their relationships to distribution of agaves. Plant, Cell and Environment, 6, 711–719.Google Scholar
Nobel, P.S., Forseth, I.N. and Long, S.P. (1993). Canopy structure and light interception. In: Photosynthesis and Production in a Changing Environment: A Field and Laboratory Manual (eds Hall, D.O., Scurlock, J.M.O., Bolhar-Nordenkampf, H.R., Leegood, R.C. and Long, S.P.), Chapman and Hall, New York, USA, pp. 78–90.Google Scholar
Nobel, P.S., Geller, G.N., Kee, S.C., et al. (1986). Temperatures and thermal tolerances for cacti exposed to high temperatures near the soil surface. Plant, Cell and Environment, 9, 279–287.Google Scholar
Nobel, P.S., Longstreth, D.J., Hartsock, T.L. (1978). Effect of water stress on the temperature optima of net CO2 exchange for two desert species. Physiologia Plantarum, 44, 97–101.CrossRefGoogle Scholar
Nobel, P.S., Zaragoza, L.J. and Smith, W.R. (1975). Relation between mesophyll surface area, photosynthesis rate, and illumination level during development for leaves of Plectranthus parviflorus Henckel. Plant Physiology, 55, 1067–1070.CrossRefGoogle ScholarPubMed
Nock, C.A., Caspersen, J.P. and Thomas, S.C. (2008). Large ontogenetic declines in intra-crown leaf-area index in two temperate deciduous tree species. Ecology, 89, 744–753.CrossRefGoogle ScholarPubMed
Noctor, G. and Foyer, C.H. (1998a). Ascorbate and glutathione: keeping active oxygen under control. Annual Review of Plant Physiology and Plant Molecular Biology, 49, 249–279.CrossRefGoogle ScholarPubMed
Noctor, G. and Foyer, C.H. (1998b). A re-evaluation of the ATP:NADPH budget during C3 photosynthesis. A contribution from nitrate assimilation and its associated respiratory activity?Journal of Experimental Botany, 49, 1895–1908.Google Scholar
Noctor, G., Arisi, A.C.M., Jouanin, L., et al. (1998). Glutathione: biosynthesis, metabolism and relationship to stress tolerance explored in transformed plants. Journal of Experimental Botany, 49, 623–647.Google Scholar
Noctor, G., De Paepe, R. and Foyer, C.H. (2007). Mitochondrial redox biology and homeostasis in plants. Trends Plant Science, 12, 125–134.CrossRefGoogle ScholarPubMed
Noguchi, K. and Yoshida, K. (2008). Interaction between photosynthesis and respiration in illuminated leaves. Mitochondrion, 8, 87–99.CrossRefGoogle ScholarPubMed
Nogueira, A., Martínez, C.A., Ferreira, L.L., et al. (2004). Photosynthesis and water use efficiency in twenty tropical tree species of different succession status in a Brazilian reforestation. Photosynthetica, 42, 351–356.CrossRefGoogle Scholar
Nogués, S., Cotxarrera, L., Alegre, L., et al. (2002). Limitations to photosynthesis in tomato leaves induced by Fusarium wilt. New Phytologist, 154, 461–470.CrossRefGoogle Scholar
Nogués, S., Tcherkez, G., Streb, P., et al. (2006). Respiratory carbon metabolites in the high mountain plant species Ranunculus glacialis. Journal of Experimental Botany, 57, 3837–3845.CrossRefGoogle Scholar
Nomura, M., Higuchi, T., Ishida, Y., et al. (2005). Differential expression pattern of C4 bundle sheath expression genes in rice, a C3 plant. Plant and Cell Physiology, 46, 754–761.CrossRefGoogle ScholarPubMed
Noodén, L.D. and Guiamét, J.I. (1997). Senescence mechanisms. Physiologia Plantarum, 101, 746–53.CrossRefGoogle Scholar
Noormets, A., McDonald, E.P., Kruger, E.L., et al. (2001a). The effect of elevated carbon dioxide and ozone on leaf and branch-level photosynthesis and potential plant-level carbon gain in aspen. Trees, 15, 262–270.CrossRefGoogle Scholar
Noormets, A., Sober, A., Pell, E.J., et al. (2001b). Stomatal and non-stomatal limitation to photosynthesis in two trembling aspen (Populus tremuloides Michx.) clones exposed to elevated CO2 and/or O3. Plant, Cell and Environment, 24, 327–336.CrossRefGoogle Scholar
Norby, R.J., Delucia, E.H., Gielen, B., et al. (2005). Forest response to elevated CO2 is conserved across a broad range of productivity. Proceedings of the National Academy of Sciences of the USA, 102, 18052–18056.CrossRefGoogle ScholarPubMed
Norby, R.J., Sholtis, J.D., Gunderson, C.A., et al. (2003). Leaf dynamics of a deciduous forest canopy: no response to elevated CO2. Oecologia, 136, 574–584.CrossRefGoogle ScholarPubMed
Norby, R.J., Wullschleger, S.D., Gunderson, C.A., et al. (1999). Tree responses to rising CO2 in field experiments: implications for the future forest. Plant, Cell and Environment, 22, 683–714.CrossRefGoogle Scholar
Nordborg, M. (2000). Linkage disequilibrium, gene trees and selfing: ancestral recombination graph with partial self-fertilization. Genetics, 154, 923–939.Google ScholarPubMed
Norman, J.M. and Jarvis, P.G. (1974). Photosynthesis in Sitka spruce (Picea sitchensis (Bong.) Carr.). III. Measurements of canopy structure and interception of radiation. Journal of Applied Ecology, 11, 375–398.CrossRefGoogle Scholar
North, P.R.J. (1996). Three-dimensional forest light interaction model using a Monte Carlo method. IEEE Transactions on Geoscience and Remote Sensing, 34, 946–955.CrossRefGoogle Scholar
Nowak, R.S. and Caldwell, M.M. (1984). A test of compensatory photosynthesis in the field: implications for herbivory tolerance. Oecologia, 61, 311–318.CrossRefGoogle ScholarPubMed
Nowak, R.S., Ellsworth, D.S. and Smith, S.D. (2004). Functional responses of plants to elevated atmospheric CO2 – do photosynthetic and productivity data from FACE experiments support early predictions?New Phytologist, 162, 253–280.CrossRefGoogle Scholar
Noy Meir, I. (1973). Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics, 4, 25–51.CrossRefGoogle Scholar
Nunes-Nesi, A., Araújo, W.L. and Fernie, A.R. (2011). Targeting mitochondrial metabolism and machinery as a means to enhance photosynthesis. Plant Physiology, 155, 101–107.CrossRefGoogle ScholarPubMed
Nunes-Nesi, A., Carrari, F., Gibon, Y., et al. (2007). Deficiency of mitochondrial fumarase activity in tomato plants impairs photosynthesis via an effect on stomatal function. Plant Journal, 50, 1093–1106.CrossRefGoogle ScholarPubMed
Nunes-Nesi, A., Carrari, F., Lytovchenko, A., et al. (2005). Enhanced photosynthetic performance and growth as a consequence of decreasing mitochondrial malate dehydrogenase activity in transgenic tomato plants. Plant Physiology, 137, 611–622.CrossRefGoogle ScholarPubMed
Nunes-Nesi, A., Sulpice, R., Gibon, Y. et al. (2008). The enigmatic contribution of mitochondrial function in photosynthesis. Journal of Experimental Botany, 59, 1675–1684.CrossRefGoogle ScholarPubMed
Ocheltree, T.W. and Loescher, H.W. (2007). Design of the AmeriFlux portable eddy covariance system and uncertainty analysis of carbon measurements. Journal of Atmospheric and Oceanic Technology, 24, 1389–1406.CrossRefGoogle Scholar
Odening, W.R., Strain, B.R. and Oechel, W.C. (1974). The effect of decreasing water potential on net CO, exchange of intact desert shrubs. Ecology, 55, 1086–95.CrossRefGoogle Scholar
Oechel, W.C. and Strain, B.R. (1985). Native species responses to increased atmospheric carbon dioxide concentration. In: Direct Effects of Increasing Carbon Dioxide on Vegetation (eds Strain, B.D. and Cure, J.D.), Department of Energy, Office of Basic Energy Sciences, Carbon Dioxide Research Division, Springfield, VA,Washington DC, USA, pp. 117–154.Google Scholar
Oerke, E.C. and Dehne, H.W. (2004). Safeguarding production – losses in major crops and the role of crop protection. Crop Protection, 23, 275–285.CrossRefGoogle Scholar
Oerke, E.C., Steiner, U., Dehne, H.W., et al. (2006). Thermal imaging of cucumber leaves affected by downy mildew and environmental conditions. Journal of Experimental Botany, 57, 2121–2132.CrossRefGoogle ScholarPubMed
Ogee, J.Cuntz, M., Peylin, P., et al. (2007). Non-steady state, non-uniform transpiration rate and leaf anatomy effects on the progressive stable isotope enrichment of leaf water along monocot leaves. Plant, Cell and Environment, 30, 367–387.CrossRefGoogle ScholarPubMed
Ogle, K. and Reynolds, J.F. (2002). Desert dogma revisited: coupling of stomatal conductance and photosynthesis in the desert shrub, Larrea tridentata. Plant, Cell and Environment, 25, 909–21.CrossRefGoogle Scholar
Ögren, E. (1988). Photoinhibition of photosynthesis in willow leaves under field conditions. Planta, 175, 229–236.CrossRefGoogle ScholarPubMed
Ögren, 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
Ögren, E. and Baker, N.R. (1985). Evaluation of a technique for the measurement of chlorophyll fluorescence from leaves exposed to continuous white light. Plant, Cell, Environment, 8, 539–547.CrossRefGoogle Scholar
Ögren, E. and Evans, J.R. (1993), Photosynthetic light response curves. 1. The influence of CO2 partial pressure and leaf inversion. Planta, 189, 182–190.CrossRefGoogle Scholar
Ogren, W.L. (1984). Photorespiration pathways, regulation, and modification. Annual Review of Plant Physiology, 35, 415–442.CrossRefGoogle Scholar
Ogren, W.L. (2003). Affixing the o to Rubisco: discovering the source of photorespiratory glycolate and its regulation. Photosynthesis Research, 76, 53–63.CrossRefGoogle ScholarPubMed
Ogren, W.L. and Bowes, G. (1971). Ribulose diphosphate carboxylase regulates soybean photorespiration. Nature New Biology, 230, 159–160.CrossRefGoogle ScholarPubMed
Oguchi, R., Douwstra, P., Fujita, T. et al. (2011). Intra-leaf gradients of photoinhibition induced by different colour lights: implications for the dual mechanisms of photoinhibition and for the application of conventional chlorophyll fluorometers. The New Phytologist, 191, 146–159.CrossRefGoogle ScholarPubMed
Oguchi, R., Hikosaka, K. and Hirose, T. (2005). Leaf anatomy as a constraint for photosynthetic acclimation: differential responses in leaf anatomy to increasing growth irradiance among three deciduous trees. Plant, Cell and Environment, 28, 916–927.CrossRefGoogle Scholar
Oguchi, R., Hikosaka, K., Hirose, T. (2003). Does leaf photosynthetic light-acclimation need change in leaf anatomy?Plant Cell Environment, 26, 505–512.CrossRefGoogle Scholar
Oguchi, R., Hikosaka, K., Hiura, T., et al. (2006). Leaf anatomy and light acclimation in woody seedlings after gap formation in a cool-temperate forest. Oecologia, 149, 571–582.CrossRefGoogle Scholar
Oguchi, R., Jia, H., Barber, J., et al. (2008). Recovery of photoinactivated photosystem II in leaves: retardation due to restricted mobility of photosystem II in the thylakoid membrane. Photosynthesis Research, 98, 621–629.CrossRefGoogle ScholarPubMed
Ohki, K. (1976). Effect of zinc nutrition on photosynthesis and carbonic anhydrase activity in cotton. Physiologia Plantarum, 38, 300–304.CrossRefGoogle Scholar
Ohki, K. (1985). Manganese deficiency and toxicity effects on photosynthesis, chlorophyll, and transpiration in wheat. Crop Science, 25, 187–191.CrossRefGoogle Scholar
Ohki, K. (1986). Photosynthesis, chlorophyll, and transpiration responses in aluminum stressed wheat and sorghum. Crop Science, 26, 572–575.CrossRefGoogle Scholar
Ohki, K., Wilson, D.O. and Anderson, O.E. (1981). Manganese deficiency and toxicity sensitivities of soybean cultivar. Agronomy Journal, 72, 713–716.CrossRefGoogle Scholar
Ohsawa, M. and Nitta, I. (1997). Patterning of subtropical/warm-temperate evergreen broad-leaved forests in East Asian mountains with special reference to shoot phenology. Tropics, 6, 317–334.CrossRefGoogle Scholar
Ohsugi, R., Samejima, M., Chonan, N., et al. (1988). δ13C values and the occurrence of suberized lamellae in some Panicum species. Annals of Botany, 62, 53–59.CrossRefGoogle Scholar
Ohsumi, A., Hamasaki, A., Nakagawa, H., et al. (2007). A model explaining genotypic and ontogenetic variation of leaf photosynthetic rate in rice (Oryza sativa) based on leaf nitrogen content and stomatal conductance. Annals of Botany, 99, 265–273.CrossRefGoogle ScholarPubMed
Ohya, T., Yoshida, S. and Kawabata, R. (2002). Biophoton emission due to drought injury in red beans: possibility of early detection of drought injury. Japanese Journal of Applied Physics Part 1-Regular Papers Short Notes and Review Papers, 41, 4766–4771.CrossRefGoogle Scholar
Oker-Blom, P. (1984). Penumbral effects of within-plant and between-plant shading on radiation distribution and leaf photosynthesis: a Monte Carlo simulation. Photosynthetica, 18, 522–528.Google Scholar
O’Leary, M.H. (1984). Measurement of the isotope fractionation associated with diffusion of carbon dioxide in aqueous solution. Journal of Physical Chemistry, 88, 823–825.CrossRefGoogle Scholar
O’Leary, M.H. (1993). Biochemical basis of carbon isotope fractionation. In: Stable Isotopes and Plant Carbon-water Relations (eds Ehleringer, J.R., Hall, A.E. and Farquhar, G.D.), Academic Press, San Diego, California, USA, pp. 19–28.Google Scholar
Oleksyn, J., Tjoelker, M.G., Lorenc-Plucinska, G., et al. (1997). Needle CO2 exchange, structure and defence traits in relation to needle age in Pinus heldreichii Christ – a relict of Tertiary flora. Trees, 12, 82–89.Google Scholar
Oleksyn, J., Zhytkowiak, R., Reich, P.B., et al. (2000). Ontogenetic patterns of leaf CO2 exchange, morphology and chemistry in Betula pendula trees. Trees, 14, 271–281.CrossRefGoogle Scholar
Oliveira, G. and Peñuelas, J. (2002). Comparative protective strategies of Cistus albidus and Quercus ilex facing photoinhibitory winter conditions. Environmental Experimental Botany, 47, 281–289.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, 19336–19341.CrossRefGoogle ScholarPubMed
Olsen, J.E. and Junttila, O. (2002). Far red end-of-day treatment restores wild-type-like plant length in hybrid aspen overexpressing phytochrome A. Physiologia Plantarum, 115, 448–457.CrossRefGoogle ScholarPubMed
Omami, E.N., Hammes, P.S. and Robbertse, P.J. (2006). Differences in salinity tolerance for growth and water-use efficiency in some amaranth (Amaranthus spp.) genotypes. New Zealand Journal of Crop and Horticultural Sciences, 34, 11–22.CrossRefGoogle Scholar
Omasa, K. and Takayama, K. (2003). Simultaneous measurement of stomatal conductance, non-photochemical quenching, and photochemical yield of photosystem II in intact leaves by thermal and chlorophyll fluorescence imaging. Plant Cell Physiology, 44, 1290–1300.CrossRefGoogle ScholarPubMed
Omasa, K., Hashimoto, Y., Kramer, P.J., et al. (1985). Direct observation of reversible and irreversible stomatal response of attached sunflower leaves to SO2. Plant Physiology, 7, 153–158.CrossRefGoogle Scholar
Omasa, K., Shimazaki, K.I., Aiga, I., et al. (1987). Image analysis of chlorophyll fluorescence transients for diagnosing the photosynthetic system of attached leaves. Plant Physiology, 84, 748–752.CrossRefGoogle ScholarPubMed
Onoda, Y., Hikosaka, K. and Hirose, T. (2004). Allocation of nitrogen to cell walls decreases photosynthetic nitrogen-use efficiency. Functional Ecology, 18, 419–425.CrossRefGoogle Scholar
Onoda, Y., Hikosaka, K. and Hirose, T. (2005). Seasonal change in the balance between capacities of RuBP carboxylation and RuBP regeneration affects CO2 response of photosynthesis in Polygonum cuspidatum. Journal of Experimental Botany, 56, 755–763.CrossRefGoogle ScholarPubMed
Oppenheimer, H.R. (1960). Adaptation to drought: xerophytism. In: Plant-water Relationships in arid and Semi-arid Conditions. Vol. 15. Reviews of Research, UNESCO – Arid Zone Research, pp. 105–138.Google Scholar
Öquist, G. and Huner, N.P.A. (2003). Photosynthesis of overwintering evergreen plants. Annual Review of Plant Physiology, 54, 329–355.Google ScholarPubMed
Öquist, G. and Huner, N.P.A. (1993). Cold-hardening induced resistance to photoinhibition in winter rye is dependent upon an increased capacity for photosynthesis. Planta, 189, 150–156.CrossRefGoogle Scholar
Öquist, G., Brunes, L. and Hällgren, J.E. (1982). Photosynthetic efficiency of Betula pendula acclimated to different quantum flux densities. Plant Cell Environment, 5, 9–15.Google Scholar
Öquist, G., Hällgren, J.E. and Brunes, L. (1978). An apparatus for measuring photosynthetic quantum yields and quanta absorption spectra of intact plants. Plant, Cell and Environment, 1, 21–27.CrossRefGoogle Scholar
Oren, R., Phillips, N., Katul, G., et al. (1998). Scaling xylem sap flux and soil water balance and calculating variance: a method for partitioning water flux in forests. Annals of Forest Science, 55, 191–216.CrossRefGoogle Scholar
Oren, R., Waring, R.H., Stafford, S.G., et al. (1987). Twenty four years of ponderosa pine growth in relation to canopy leaf area and understory competition. Forest Science, 33, 538–547.Google Scholar
Orians, G.H. and Milewski, A.V. (2007). Ecology of Australia: the effects of nutrient poor soils and intense fires. Biological Reviews, 82, 393–423.CrossRefGoogle Scholar
Ort, D.R. and Baker, N.R. (2002). A photoprotective role for O2 as an alternative electron sink in photosynthesis?Current Opinion in Plant Biology, 5, 193–198.CrossRefGoogle Scholar
Ortiz-López, A, Ort, D.R. and Boyer, J.S. (1991). Photophosphorylation in attached leaves of Helianthus annuus at low water potentials. Plant Physiology, 96, 1018–1025.CrossRefGoogle ScholarPubMed
Osborne, C.P., Beerling, D.J., Lomax, B.H., et al. (2004a). Biophysical constraints on the origin of leaves inferred from the fossil record. Proceedings of the National Academy of Sciences, 101, 10360–10362.CrossRefGoogle ScholarPubMed
Osborne, C.P. (2008). Atmosphere, ecology and evolution: what drove the Miocene expansion of C4 grasslands?Journal of Ecology, 96, 35–45.Google Scholar
Osborne, C.P. and Beerling, D.J. (2006). Nature’s green revolution: the remarkable evolutionary rise of C4 plants. Philosophical Transactions of the Royal Society, B, 361, 173–194.CrossRefGoogle ScholarPubMed
Osborne, C.P., Chaloner, W.G. and Beerling, D.J. (2004b). Falling atmospheric CO2 – the key to megaphyll leaf origins. In: The Evolution of Plant Physiology (eds. Hemsley, A.R. and Poole, I.), Elsevier Academic Press, London, UK, pp. 197–215.Google Scholar
Osborne, C.P., Mitchell, P.L., Sheehy, J.E., et al. (2000). Modelling the recent historical impacts of atmospheric CO2 climate change on Mediterranean vegetation. Global Change Biology, 6, 445–458.CrossRefGoogle Scholar
Osmond, B., Ananyav, G., Berry, J.A. et al. (2004). Changing the way we think about global change research: scaling up in experimental ecosystem science. Global Change Biology, 10, 393–407.CrossRefGoogle Scholar
Osmond, B., Schwartz, O. and Gunning, B. (1999). Photoinhibitory printing on leaves, visualized by chlorophyll fluorescence imaging and confocal microscopy, is due to diminished fluorescence from grana. Australian Journal of Plant Physiology, 26, 717–724.CrossRefGoogle Scholar
Osmond, C.B. (1979). Crassulacean acid metabolism: a curiosity in context. Annual Reviews in Plant Physiology, 29, 379–414.CrossRefGoogle Scholar
Osmond, C.B. (1983). Interactions between irradiance, nitrogen nutrition, and water stress in the sun-shade responses of Solanum dulcamara. Oecologia, 57, 316–321.CrossRefGoogle ScholarPubMed
Osmond, C.B. (1994). What is photoinhibition? Some insights from comparisons of shade and sun plants. In: Photoinhibition of Photosynthesis. From Molecular Mechanisms to the Field (eds. Baker, N.R. and Bowyer, J.R.), BIOS Scientific, Oxford, UK, pp. 1–24.Google Scholar
Osmond, C.B., Anderson, J.M., Ball, M.C., et al. (1999). Compromising efficiency: the molecular ecology of light-resource utilization in plants. In: Physiological Plant Ecology. The 39th Symposium of the British Ecological Society held at the University of York, 7–9 September 1998 (Press, M.C., Scholes, J.D. and Barker, M.G.), pp. 1–24, Blackwell Science, Oxford, UK.Google Scholar
Osmond, C.B., Badger, M., Maxwell, K., et al. (1997). Too many photons: photorespiration, photoinhibition and photooxidation. Trends in Plant Science, 2, 119–121.CrossRefGoogle Scholar
Osmond, C.B., Björkman, O. and Anderson, D.J. (1980b). Physiological processes in plant ecology. Toward a synthesis with Atriplex. Springer Verlag, Berlin – Heidelberg – New York.
Osmond, C.B., Daley, P.F., Badger, M.R., et al. (1998). Chlorophyll fluorescence quenching during photosynthetic induction in leaves of Abutilon striatum Dicks. infected with Abutilon mosaic virus, observed with a field-portable imaging system. Botanica Acta, 111, 390–397.CrossRefGoogle Scholar
Osmond, C.B., Ludlow, M.M., Davis, R., et al. (1979). Stomatal responses to humidity in Opuntia inermis in relation to control of CO2 and H2O exchange patterns. Oecologia, 41, 65–76.CrossRefGoogle ScholarPubMed
Osmond, C.B., Neales, T.F. and Stange, G. (2008). Curiosity and context revisited: crassulacean acid metabolism in the anthropocene. Journal of Experimental Botany 59, 1489–1502.CrossRefGoogle ScholarPubMed
Osmond, C.B., Winter, K. and Ziegler, H. (1982). Funtional significance of different pathways of CO2 fixation in photosynthesis. In: Encyclopedia of Plant Physiology, Vol 12B (eds. Lange, O.L., Nobel, P.S., Osmond, C.B. and Ziegler, H.) Springer-Verlag, Berlin, Germany, pp. 479–547.Google Scholar
Osmond, C.B., Winter, K. and Powles, S.B. (1980a). Adaptive significance of carbon dioxide recycling during photosynthesis in water stressed plants. In: Adaptation of Plants to Water and High Temperature Stress? (eds Turner, N.C. and Kramer, P.J.), Wiley Interscience, New York, USA, pp. 137–154.Google Scholar
Osteryoung, K.W. and Nunnari, J. (2003). The division of endosymbiotic organelles. Science, 302, 1698–1704.CrossRefGoogle ScholarPubMed
Oswald, O., Martin, T., Dominy, P.J., et al. (2001). Plastid redox state and sugars: interactive regulators of nuclear-encoded photosynthetic gene expression. Proceedings of the Natural Academy of Sciences USA, 98, 2047–2052.CrossRefGoogle ScholarPubMed
Otegui, M.E., Nicolini, M.G., Ruiz, R.A., et al. (1995). Sowing date effects on grain yield components for different maize genotypes. Agronomy Journal, 87, 29–33.CrossRefGoogle Scholar
Ott, T., Clarke, J., Birks, K., et al. (1999). Regulation of the photosynthetic electron transport chain. Planta, 209, 250–258.CrossRefGoogle ScholarPubMed
Ottander, C., Campbell, D. and Öquist, G. (1995). Seasonal changes in photosystem II organization and pigment composition in Pinus sylvestris. Planta, 197, 176–183.CrossRefGoogle Scholar
Ottoni, T.B., Matthias, A.D., Guerra, A.F., et al. (1992). Comparison of three resistance methods for estimating heat flux under stable conditions. Agricultural and Forest Meteorology, 58, 1–18.CrossRefGoogle Scholar
Ounis, A., Cerovic, Z.G., Briantais, J.M., et al. (2001a). Dual-excitation FLIDAR for the estimation of epidermal UV absorption in leaves and canopies. Remote Sensing of Environment, 76, 33–48.CrossRefGoogle Scholar
Ounis, A., Evain, S., Flexas, J., et al. (2001b). Adaptation of a PAM-fluorometer for remote sensing of chlorophyll fluorescence. Photosynthesis Research, 68, 113–120.CrossRefGoogle ScholarPubMed
Ouzounidou, G. (1996). The use of photoacoustic spectroscopy in assessing leaf photosynthesis under copper stress: correlation of energy storage to photosystem II fluorescence parameters and redox change of P700. Plant Science, 113, 229–237.CrossRefGoogle Scholar
Ouzounidou, G., Ilias, I., Tranopoulou, H., et al. (1998). Amelioration of copper toxicity by iron on spinach physiology. Journal of Plant Nutrition, 21(10), 2089–2101.CrossRefGoogle Scholar
Ouzounidou, G., Moustakas, M. and Strasser, R.J. (1997). Sites of action of copper in the photosynthetic apparatus of maize leaves: kinetic analysis of chlorophyll fluorescence, oxygen evolution, absorption changes and thermal dissipation as monitored by photoacoustic signals. Australian Journal of Plant Physiology, 24, 81–90.CrossRefGoogle Scholar
Ouzounidou, G., Symeonidis, L., Babalonas, D., et al. (1994). Comparative responses of a copper-tolerant and a copper-sensitive population of Minuartia hirsuta to copper toxicity. Journal of Plant Physiology, 144, 109–115.CrossRefGoogle Scholar
Ovaska, J., Maenpaa, P., Nurmi, A., et al. (1990). Distribution of chlorophyll-protein complexes during chilling in the light compared with heat-induced modifications. Plant Physiology, 93, 48–54.CrossRefGoogle ScholarPubMed
Ovaska, J.A., Nilsen, J., Wielgolaski, F.E., et al. (2005). Phenology and performance of mountain birch provenances in transplant gardens: latitudinal, altitudinal and oceanity continentality gradients. In: Plant Ecology, Herbivory, and Human Impact in Nordic Mountain Birch Forests (eds Wielgolaski, F.E., Karlsson, P.S., Neuvonen, S. and Thannheiser, D.), Springer Verlag, Berlin, Germany, pp. 99–115.Google Scholar
Overdieck, D. (1993). Effects of atmospheric CO2 enrichment on CO2 exchange rates of beech stands in small model ecosystems. Water Air and Soil Pollution, 70, 259–277.CrossRefGoogle Scholar
Overdieck, D. and Forstreuter, M. (1994). Evapotranspiration of beech stands and transpiration of beech leaves subject to atmospheric CO2 enrichment. Tree Physiology, 14, 997–1003.CrossRefGoogle Scholar
Overmyer, K., Kollist, H., Tuominen, H., et al. (2008). Complex phenotypic profiles leading to ozone sensitivity in Arabidopsis thaliana mutants. Plant, Cell and Environment, 31, 1237–1249.CrossRefGoogle ScholarPubMed
Owen, P.C. (1957). The effect of infection with tobacco etch virus on the rates of respiration and photosynthesis of tobacco leaves. Annals of Applied Biology, 45, 327–331.CrossRefGoogle Scholar
Owensby, C.E., Ham, J.M., Knapp, A.K., et al. (1997). Water vapour fluxes and their impact under elevated CO2 in a C4-tallgrass prairie. Global Change Biology, 3, 189–195.CrossRefGoogle Scholar
Owensby, C.E., Ham, J.M., Knapp, A.K., et al. (1999). Biomass production and species composition change in a tallgrass prairie ecosystem after long-term exposure to elevated atmospheric CO2. Global Change Biology, 5, 497–506.CrossRefGoogle Scholar
Owen-Smith, R.N. (1988). Megaherbavores: The Influence of Very Large Body Size on Ecology. Cambridge University Press, Cambridge, UK.CrossRefGoogle Scholar
Oxborough, K. (2004). Using chlorophyll a fluorescence imaging to monitor photosynthetic performance. In: Chlorophyll a Fluorescence: A Signature of Photosynthesis, Advances in Photosynthesis and Respiration, vol.19 (eds Papageorgiou, C.G. and Govindjee, J.), Springer, Dordrecht, Netherlands, pp. 389–407.Google Scholar
Oxborough, K. and Baker, N.R. (1997). An instrument capable of imaging chlorophyll a fluorescence from intact leaves at very low irradiance and at cellular and sub-cellular levels. Plant, Cell and Environment, 20, 1473–1483.CrossRefGoogle Scholar
Pace, C.N., Shirley, B.A., McNutt, M., et al. (1996). Forces contributing to the conformational stability of proteins. The FASEB Journal, 10, 75–83.CrossRefGoogle ScholarPubMed
Pachepsky, L.B., Reddy, V.R. and Acock, B. (1994). Cotton canopy photosynthesis model for predicting the effect of temperature and elevated carbon dioxide concentration. Biotronics, 23, 35–46.Google Scholar
Pagani, M., Freeman, K.H. and Arthur, M.A. (1999). Late Miocene atmospheric CO2 concentrations and the expansion of C4 grasses. Science, 285, 876–879.CrossRefGoogle Scholar
Pagani, M., Zachos, J.C., Freeman, K.H., et al. (2005). Marked decline in atmospheric carbon dioxide concentrations during the Paleogene. Science, 309, 600–660.CrossRefGoogle ScholarPubMed
Pälike, H., Norris, R.D., Herrle, J.O., et al. (2006). The heartbeat of the Oligocene climate system. Science, 314, 1894–1898.CrossRefGoogle ScholarPubMed
Palmer, J.D. (1985). Comparative organization of chloroplast genomes. Annual Review of Genetics, 19, 325–354.CrossRefGoogle ScholarPubMed
Palmer, J.W. (1992). Effects of varying crop load on photosynthesis, dry matter production and partitioning of Crispin/M.27 apple trees. Tree Physiology, 11, 19–33.CrossRefGoogle ScholarPubMed
Palmer, J.W., Giuliani, R. and Adams, H.M. (1997). Effect of crop load on fruiting and leaf photosynthesis of ‘Braeburn’/M.26 apple trees. Tree Physiology, 17, 741–746.CrossRefGoogle ScholarPubMed
Palmer, T.N. and Räisänen, J. (2002). Quantifying the risk of extreme seasonal precipitation events in a changing climate. Nature, 415, 514–517.CrossRefGoogle Scholar
Palmquist, K. (2000). Carbon economy of lichen. New Phytologist, 148, 11–36.CrossRefGoogle Scholar
Palmqvist, K., Sundblad, L.G., Samuelsson, G., et al. (1986). A correlation between changes in luminescence decay kinetics and the appearance of a CO2-accumulating mechanism in Scenedesmus obliquus. Photosynthesis Research, 10, 113–123.CrossRefGoogle ScholarPubMed
Palmroth, S., Berninger, F., Lloyd, J., et al. (1999). No water conserving behaviour is observed in Scots pine from wet to dry climates. Oecologia, 121, 302–309.CrossRefGoogle Scholar
Panda, S., Mishra, A.K. and Biswal, U.C. (1986). Manganese-induced modification of membrane lipid peroxidation during aging of isolated wheat chloroplasts. Photobiochemistry and Photobiophysics, 13, 53–61.Google Scholar
Paoletti, E. (1998). UV-B and acid rain effects on beech (Fagus sylvatica L.) and holm oak (Quercus ilex L.) leaves. Chemosphere, 36, 835–840.CrossRefGoogle Scholar
Papadopoulos, Y.A., Gordon, R.J., McRae, K.B., et al. (1999). Current and elevated levels of UV-B radiation have few impacts on yields of perennial forage crops. Global Change Biology, 5, 847–856.CrossRefGoogle Scholar
Papageorgiou, G. (1975). Chlorophyll fluorescence: an intrinsic probe of photosynthesis. In: Bioenergetics of Photosynthesis (ed. Govindjee, J.), Academic Press, New York, USA, pp. 319–371.CrossRef
Papageorgiou, G.C. and Govindjee, J. (eds) (2004). Chlorophyll a Fluorescence: A Signature of Photosynthesis. Advances in Photosynthesis and Respiration. Kluwer Academic Publishers, Dordrecht, Netherlands.Google Scholar
Papageorgiou, G.C., Tsimilli-Michael, M. and Stamatakis, K. (2007). The fast and slow kinetics of chlorophyll a fluorescence induction in plants, algae and cyanobacteria: a view point. Photosynthesis Research, 94, 275–290.CrossRefGoogle Scholar
Papale, D. and Valentini, R. (2003). A new assessment of European forests carbon exchanges by eddy fluxes and artificial neural network spacialization, Global Change Biology, 9, 525–535.CrossRefGoogle Scholar
Papale, D., Reichstein, M., Aubinet, M., et al. (2006). Towards a standardized processing of Net Ecosystem Exchange measured with eddy covariance technique: algorithms and uncertainty estimation. Biogeosciences, 3, 1–13.CrossRefGoogle Scholar
Parent, B., Hachez, C., Redondo, E., et al. (2009). Drought and abscisic acid effects on aquaporin content translate into changes in hydraulic conductivity and leaf growth rate: a trans-scale approach. Plant Physiology, 149, 2000–2012.CrossRefGoogle ScholarPubMed
Parida, A.K., Das, A.B. and Mitra, B. (2004). Effects of SALT on growth, ion accumulation, photosynthesis and leaf anatomy of the mangrove, Bruguiera parviflora, 5, 18, 167–174.Google Scholar
Park, R. and Epstein, S. (1960). Carbon isotope fractionation during photosynthesis. Geochimica et Cosmochimica Acta, 21, 110–126.CrossRefGoogle Scholar
Park, S.Y., Yu, J.W., Park, J.S., et al. (2007). The senescence-induced staygreen protein regulates chlorophyll degradation. Plant Cell, 19, 1649–1664.CrossRefGoogle ScholarPubMed
Parkhurst, D.F. (1994). Diffusion of CO2, and other gases inside leaves. New Phytologist, 126, 449–79.CrossRefGoogle Scholar
Parkhurst, D.F. and Mott, K.A. (1990). Intercellular diffusion limits to CO2 uptake in leaves. 1. Studies in air and helox. Plant Physiology, 94, 1024–1032.CrossRefGoogle Scholar
Parkhurst, D.F., Wong, S.C., Farquhar, G.D., et al. (1988). Gradients of intercellular CO2 levels across the leaf mesophyll. Plant Physiology, 86, 1032–1037.CrossRefGoogle Scholar
Pärnik, T.R. and Keerberg, O.F. (2006). Advanced radiogasometric method for the determination of the rates of photorespiratory and respiratory decarboxylations of primary and stored photosynthates under steady state photosynthesis. Physiologia Plantarum, 129, 34–44.CrossRefGoogle Scholar
Pärnik, T.R., Voronin, P.Y., Ivanona, H.N., et al. (2002). Respiratory CO2 fluxes in photosynthesizing leaves of C3 species varying in rates of starch synthesis. Russian Journal of Plant Physiology, 49, 729–735.CrossRefGoogle Scholar
Parra, M.J., Acuña, K., Corcuera, L.J., et al. (2009). Vertical distribution of Hymenophyllaceae species among host tree microhabitats in a temperate rainforest in Southern Chile. Journal of Vegetation Science, 20, 588–595.CrossRefGoogle Scholar
Parry, M.A.J., Andralojc, P.J., Parmar, S., et al. (1997). Regulation of Rubisco by inhibitors in the light. Plant Cell and Environment, 20, 528–534.CrossRefGoogle Scholar
Parry, M.A.J., Flexas, J. and Medrano, H. (2005). Prospects for crop production under drought: research priorities and futures directions. Annals of Applied Biology, 147, 211–226.CrossRefGoogle Scholar
Parry, M.A.J., Keys, A.J. and Gutteridge, S. (1989). Variation in the specificity factor of C-3 higher-plant Rubisco determined by the total consumption of ribulose-P2. Journal of Experimental Botany, 40, 317–320.CrossRefGoogle Scholar
Parry, M.A.J., Keys, A.J., Madgwick, P.J., et al. (2008). Rubisco regulation: a role for inhibitors. Journal of Experimental Botany, 59, 1569–1580.CrossRefGoogle ScholarPubMed
Parry, M.A.J., Madgwick, P.J., Bayon, C., et al. (2009). Mutation discovery for crop improvement. Journal of Experimental Botany, 60, 2817–2825.CrossRefGoogle ScholarPubMed
Parry, M.A.J., Madgwick, P.J., Carvalho, J.F.C., et al. (2007). Prospects for increasing photosynthesis by overcoming the limitations of Rubisco. Journal of Agricultural Science, 145, 31–43.CrossRefGoogle Scholar
Parsons, W.T. and Cuthbertson, E G. (2001). Noxious Weeds of Australia. CSIRO Publishing, Canberra, Australia.Google Scholar
Passey, B.H., Cerling, T.E., Perkins, M.E., et al. (2002). Environmental change in the Great Plains: an isotopic record from fossil horses. Journal of Geology, 110, 123–140.CrossRefGoogle Scholar
Pastenes, C. and Horton, P. (1996). Effect of high temperature on photosynthesis in beans. 1. Oxygen evolution and chlorophyll fluorescence. Plant Physiology, 112, 1245–1251.CrossRefGoogle Scholar
Pastenes, C. and Horton, P. (1999). Resistance of photosynthesis to high temperature in two bean varieties (Phaseolus vulgaris L.). Photosynthesis Research, 62, 197–203.CrossRefGoogle Scholar
Pastenes, C., Pimentel, P. and Lillo, J. (2005). Leaf movements and photoinhibition in relation to water stress in field-grown beans. Journal of Experimental Botany 56, 425–433.CrossRefGoogle ScholarPubMed
Pastori, G.M. and Foyer, C.H. (2002). Common components, networks, and pathways of cross-tolerance to stress. The central role of “redox” and abscisic acid-mediated controls. Plant Physiology, 129, 460–468.CrossRefGoogle ScholarPubMed
Pataki, D.E., Oren, R. and Smith, W.K. (2000). Sap flux of co-occurring species in a western subalpine forest during seasonal soil drought. Ecology, 81, 2557–2566.CrossRefGoogle Scholar
Pate, J. and Arthur, D. (1998). δ13C analysis of phloem sap carbon: novel means of evaluating seasonal water stress and interpreting carbon isotope signatures of foliage and trunk wood of Eucalyptus globulus. Oecologia, 117, 301–311.CrossRefGoogle ScholarPubMed
Patel, M., Siegel, A.J. and Berry, J.O. (2006). Untranslated regions of FbRbcS1 mRNA mediate bundle sheath cell-specific gene expression in leaves of a C4 plant. Journal of Biological Chemistry, 281, 25485–25491.CrossRefGoogle ScholarPubMed
Patiño, S., Tyree, M. and Herre, E.A. (1995). Comparison of hydraulic architecture of woody plants of differing phylogeny and growth form with special reference to free-standing and hemi-epiphytic Ficus species from Panama. New Phytologist, 129, 125–134.CrossRefGoogle Scholar
Patsikka, E., Kairavuo, M., Sersen, F., et al. (2002). Excess copper predisposes photosystem II to photoinhibition in vivo by outcompeting iron and causing decrease in leaf chlorophyll. Plant Physiology, 129, 1359–1367.CrossRefGoogle ScholarPubMed
Paul, M.J. and Driscoll, S.P. (1997). Sugar repression of photosynthesis: the role of carbohydrates in signalling nitrogen deficiency through source:sink imbalance. Plant, Cell and Environment, 20, 110–116.CrossRefGoogle Scholar
Paul, M.J. and Foyer, C.H. (2001). Sink regulation of photosynthesis. Journal of Experimental Botany, 52, 1383–1400.CrossRefGoogle ScholarPubMed
Paula, S. and Pausas, J.G. (2006). Leaf traits and resprouting ability in the Mediterranean basin. Functional Ecology, 20, 941–947.CrossRefGoogle Scholar
Pavel, E.W. and Dejong, T.M. (1993). Seasonal CO2 exchange patterns of developing peach (Prunus persica) fruits in response to temperature, light and CO2 concentration. Physiologia Plantarum, 88, 322–330.CrossRefGoogle Scholar
Pavlovic, A., Masarovičová, E. and Hudák, J. (2007). Carnivorous syndrome in Asian pitcher plants of the genus Nepenthes. Annals of Botany, 100, 527–536.CrossRefGoogle ScholarPubMed
Payton, P., Allen, R.D., Trolinder, N., et al. (1997). Over-expression of chloroplast-targeted Mn superoxide dismutase in cotton (Gossypium hirsutum L., cv. Coker 312) does not alter the reduction of photosynthesis after short exposures to low temperature and high light intensity. Photosynthesis Research, 52, 233–244.CrossRefGoogle Scholar
Pearce, D.W., Millard, S., Bray, D.F., et al. (2006). Stomatal characteristics of riparian poplar species in a semi-arid environment. Tree Physiology, 26, 211–218.CrossRefGoogle Scholar
Pearcy, R.B. and Sims, D.A. (1994). Photosynthetic acclimation to changing light environments: scaling from leaf to the whole plant. In: Exploitation of Environmental Heterogeneity by Plants (eds Caldwell, M.M. and Pearcy, R.W.), Academic Press, San Diego, USA, pp. 145–174.Google Scholar
Pearcy, R.W. (1983). The light environment and growth of C3 and C4 tree species in the understory of a Hawaiian forest. Oecologia, 58, 19–25.CrossRefGoogle ScholarPubMed
Pearcy, R.W. (1988). Photosynthetic utilization of lightflecks by understory plants. Australian Journal of Plant Physiology, 15, 223–238.CrossRefGoogle Scholar
Pearcy, R.W. (1989). Radiation and light measurements. In: Plant Physiological Ecology: Field Methods and Instrumentation, Vol 457 (eds Pearcy, R.W., Ehleringer, J.R., Mooney, H.A. and Rundel, P.W.). Chapman and Hall, New York, USA, pp. 353–359.CrossRefGoogle Scholar
Pearcy, R.W. (1990). Sunflecks and photosynthesis in plant canopies. Annual Review of Plant Physiology and Plant Molecular Biology, 41, 421–453.CrossRefGoogle Scholar
Pearcy, R.W. (2007). Responses of plants to heterogeneous light environments. In: Functional Plant Ecology (eds Pugnaire, F.I. and Valladares, F.), CRC Press, New York, USA, pp. 213–258.Google Scholar
Pearcy, R.W. and Yang, W. (1996). A three-dimensional crown architecture model for assessment of light capture and carbon gain by understory plants. Oecologia, 108, 1–12.CrossRefGoogle ScholarPubMed
Pearcy, R.W., Bjorkman, O., Caldwell, M.M., et al. (1987). Carbon gain by plants in natural environments. BioScience, 37, 21–29.CrossRefGoogle Scholar
Pearcy, R.W., Chazdon, R.L., Gross, L.J., et al. (1994). Photosynthetic utilization of sunflecks, a temporally patchy resource on a time-scale of seconds to minutes. In: Exploitation of Environmental Heterogeneity by Plants (eds Caldwell, M.M. and Pearcy, R.W.), Academic Press, San Diego, USA, pp. 175–208.Google Scholar
Pearcy, R.W., Gross, L.J. and He, D. (1997). An improved dynamic model of photosynthesis for estimation of carbon gain in sunfleck light regimes. Plant, Cell and Environment, 20, 411–424.CrossRefGoogle Scholar
Pearcy, R.W., Harrison, A.T., Mooney, H.A., et al. (1974). Seasonal changes in net photosynthesis of Atriplex hymenelytra shrubs growing in Death Valley, California. Oecologia, 17, 111–121.CrossRefGoogle ScholarPubMed
Pearcy, R.W., Muraoka, H. and Valladares, F. (2005). Crown architecture in sun and shade environments: assessing function and trade-offs with a three-dimensional simulation model. New Phytologist, 166, 791–800.CrossRefGoogle Scholar
Pearcy, R.W., Valladares, F., Wright, S.J., et al. (2004). A functional analysis of the crown architecture of tropical forest Psychotria species: do species vary in light capture efficiency and consequently in carbon gain and growth?Oecologia, 139, 163–177.CrossRefGoogle Scholar
Pearse, I., Heath, K.D. and Cheeseman, J.M. (2005). A partial characterization of peroxidase in Rhizophora mangle. Plant, Cell and Environment, 28, 612–622.CrossRefGoogle Scholar
Pedrós, R., Goulas, Y., Jacquemoud, S., et al. (2009). FluorMODleaf: a new leaf fluorescence emission model based on the PROSPECT model. Remote Sensing of Environment, 114, 155–167.CrossRef
Pedrós, R., Moya, I., Goulas, Y., et al. (2008). Chlorophyll fluorescence emission spectrum inside a leaf. Photochemical and Photobiological Sciences, 7, 498–502.CrossRefGoogle ScholarPubMed
Peet, M.M. and Kramer, P.J. (1980). Effects of decreasing source-sink ratio in soybeans on photosynthesis, photo-respiration, transpiration and yield. Plant, Cell and Environment, 3, 201–206.Google Scholar
Peguero-Pina, F.J., Gil-Pelegrín, E. and Morales, F. (2009). Photosystem II efficiency of the palisade and spongy mesophyll in Quercus coccifera using adaxial/abaxial illumination and excitation light sources with wavelengths varying in penetration into the leaf tissue. Photoynthesis Research, 99, 49–61.CrossRefGoogle Scholar
Peguero-Pina, J.J., Morales, F., Flexas, J., et al. (2008). Photochemistry, remotely sensed physiological reflectance index and de-epoxidation state of the xanthophylls cycle in Quercus coccifera under intense drought. Oecologia, 156, 1–11.CrossRefGoogle ScholarPubMed
Pei, Z.M., Murata, Y., Benning, G., et al. (2000). Calcium channels activated by hydrogen peroxide mediate abscisic acid signalling in guard cells. Nature. 406, 731–734.CrossRefGoogle ScholarPubMed
Peichl, M. and Arain, M.A. (2007). Allometry and partitioning of above- and belowground tree biomass in an age-sequence of white pine forests. Forest Ecology and Management, 253, 68–80.CrossRefGoogle Scholar
Peisker, M. and Apel, H. (2001). Inhibition by light of CO2 evolution from dark respiration: comparison of two gas exchange methods. Photosynthesis Research, 70, 291–298.CrossRefGoogle Scholar
Pell, E.J., Schlagnhaufer, C.D. and Arteca, R.N. (1997). Ozone-induced oxidative stress: mechanisms of action and reaction. Physiologia Plantarum, 100, 264–273.CrossRefGoogle Scholar
Peltier, G. and Cournac, L. (2002). Chlororespiration. Annual Review of Plant Physiology, 53, 523–550.Google ScholarPubMed
Peltier, G. and Thibault, P. (1985). O2 uptake in the light in Chlamydomonas: evidence for persistent mitochondrial respiration. Plant Physiology, 79, 225–230.CrossRefGoogle Scholar
Peltzer, D. and Polle, A. (2001). Diurnal fluctuation of antioxidative systems in leaves of field-grown beech trees (Fagus sylvatica): responses to light and temperature. Physiologia Plantarum, 111, 158–164.CrossRefGoogle Scholar
Penman, H.L. (1948). Natural evaporation from open water bare soil and grass. Proceedings of the Royal Society, 193, 120–145.CrossRefGoogle Scholar
Penning de Vries, F.W.T., Brunsting, A.H.M. and Van Laar, H.H. (1974). Products, requirements and efficiency of biosynthesis: a quantitative approach. Journal of Theoretical Biology, 45, 339–77.CrossRefGoogle ScholarPubMed
Peñuelas, J. and Azcon-Bieto, J. (1992). Changes in leaf δ13C of herbarium plant species during the last 3 centuries of CO2 increase. Plant Cell and the Environment, 15, 485–489.CrossRefGoogle Scholar
Peñuelas, J. and Boada, M. (2003). A global change-induced biome shift in the Motseny mountains (NE Spain). Global Change Biology, 9, 131–140.CrossRefGoogle Scholar
Peñuelas, J. and Filella, I. (1998). Visible and near-infrared reflectance techniques for diagnosing plant physiological status. Trends in Plant Science, 3, 151–156.CrossRefGoogle Scholar
Peñuelas, J. and Llusià, J. (2003). BVOCs: plant defense against climatic warming?Trends in Plant Science, 8, 105–109.CrossRefGoogle Scholar
Peñuelas, J. and Munné-Bosch, S. (2005). Isoprenoids: an evolutionary pool for photoprotection. Trends in Plant Science, 10, 166–169.CrossRefGoogle ScholarPubMed
Peñuelas, J., Baret, F. and Filella, I. (1995a). Semi-empirical indices to assess carotenoids/chlorophyll a ratio from spectral reflectance. Photosynthetica, 31, 221–230.Google Scholar
Peñuelas, J., Filella, I. and Gamon, J.A. (1995c). Assessment of photosynthetic radiation-use efficiency with spectral reflectance. New Phytologist, 131, 291–296.CrossRefGoogle Scholar
Peñuelas, J., Filella, I., Biel, C., et al. (1993a). The reflectance at the 950–970 nm region as an indicator of plant water status. International Journal of Remote Sensing, 14, 1887–1905.CrossRefGoogle Scholar
Peñuelas, J., Filella, I., Lloret, P., et al. (1995b). Reflectance assessment of mite attack on apple trees. International Journal of Remote Sensing, 16, 2727–2733.CrossRefGoogle Scholar
Peñuelas, J., Filella, I., Llusia, J., et al. (1998). Comparative field study of spring and summer leaf gas exchange and photobiology of the Mediterranean trees Quercus ilex and Phillyrea latifolia. Journal of Experimental Botany, 49, 229–238.Google Scholar
Peñuelas, J., Gamon, J.A., Fredeen, A.L., et al. (1994). Reflectance indices associated with physiological changes in nitrogen- and water-limited sunflower leaves. Remote Sensing of Environment, 48, 135–146.CrossRefGoogle Scholar
Peñuelas, J., Gamon, J.A., Griffin, K.L., et al. (1993b). Assessing community type, plant biomass, pigment composition, and photosynthetic efficiency of aquatic vegetation from spectral reflectance. Remote Sensing of the Environment, 46, 1–25.CrossRefGoogle Scholar
Peñuelas, J., Isla, R., Filella, I., et al. (1997b). Visible and near-infrared reflectance assessment of salinity effects on barley. Crop Science, 37, 198–202.CrossRefGoogle Scholar
Peñuelas, J., Llusià, J., Piñol, J., et al. (1997a). Photochemical reflectance index and leaf photosynthetic radition-use-efficiency assessment in Mediterranean trees. International Journal of Remote Sensing, 18, 2863–2868.CrossRefGoogle Scholar
Peñuelas, J., Munné-Bosch, S., Llusià, J., et al. (2004). Leaf reflectance and photo- and antioxidant protection in field-grown summer-stressed Phillyrea angustifolia. Optical signals of oxidative stress?New Phytologist, 162, 115–124.CrossRefGoogle Scholar
Peñuelas, J., Prieto, P., Beier, C., et al. (2007). Response of plant species richness and primary productivity in shrublands along a north-south gradient in Europe to seven years of experimental warming and drought: reductions in primary productivity in the heat and drought year of 2003. Global Change Biology, 13, 2563–2581.CrossRefGoogle Scholar
Pepin, S. and Livingston, N.J. (1997). Rates of stomatal opening in conifer seedlings in relation to air temperature and daily carbon gain. Plant, Cell and Environment, 20, 1462–1472.CrossRefGoogle Scholar
Pereira, J.S. and Chaves, M.M. (1993). Plant water deficit in Mediterranean ecosystems In: Water Deficits: Plant Responses from Cell to Community (eds Smith, J.A.C. and Griffiths, H.), Bios Scientific Publishing, USA, pp. 237–251.Google Scholar
Pereira, J.S., Mateus, J.A., Aires, L.M., et al. (2007). Net ecosystem carbon exchange in three contrasting Mediterranean ecosystems-the effect of drought. Biogeosciences, 4, 791–802.CrossRefGoogle Scholar
Pérez, C., Madero, P., Pequerul, A., et al. (1993). Specificity of manganese in some aspects of soybean (Glycine max L.) physiology. In: Optimization of Plant Nutrition (eds. Fragoso, M.A.C. and van Beusichem, M.L.), Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 503–507.Google Scholar
Pérez-Bueno, M.L., Ciscato, M., vandeVen, M., et al. (2006). Imaging viral infection. Studies on Nicotiana benthamiana plants infected with the pepper mild mottle tobamovirus. Photosynthesis Research, 90, 111–123.CrossRefGoogle ScholarPubMed
Pérez-Bueno, M.L., Rahoutei, J., Sajnani, C., et al. (2004). Proteomic analysis of the oxygen-evolving complex of photosystem II under biotic stress. Studies on Nicotiana benthamiana infected with tobamoviruses. Proteomics, 4, 418–425.CrossRefGoogle Scholar
Perez-Martin, A., Flexas, J., Ribas-Carbó, M., et al. (2009). Interactive effects of soil water deficit and air vapour pressure deficit on mesophyll conductance to CO2 in Vitis vivifera and Olea europaea. Journal of Experimental Botany, 60, 2391–2405.CrossRefGoogle ScholarPubMed
Perez-Peña, J. and Tarara, J. (2004). A portable whole canopy gas exchange system for several mature field-grown grapevines. Vitis, 43, 7–14.Google Scholar
Pérez-Pérez, J.G., Syvertsen, J.P., Botía, P., et al. (2007). Leaf water relations and net gas exchange responses of salinized Carrizo citrange seedlings during drought stress and recovery. Annals of Botany, 100, 335–345.CrossRefGoogle ScholarPubMed
Pérez-Priego, O., Zarco-Tejada, P., Miller, J.R., et al. (2005). Detection of water stress in orchard trees with a high-resolution spectrometer through chlorophyll fluorescence in-filling of the O2-a Band. IEEE Transactions on Geoscience and Remote Sensing, 43, 2, December 2005.
Pérez-Torres, E., Bravo, L.A., Corcuera, L.J., et al. (2007). Is electron transport to oxygen an important mechanism in photoprotection? Contrasting responses from antarctic vascular plants. Physiologia Plantarum, 130, 185–194.CrossRefGoogle Scholar
Pérez-Torres, E., Garcia, A., Dinamarca, J., et al. (2004). The role of photochemical quenching and antioxidants in photoprotection of Deschampsia antartica. Functional Plant Biology, 31, 731–741.CrossRefGoogle Scholar
Perks, M.P., Irvine, J. and Grace, J. (2002). Canopy stomatal conductance and xylem sap abscisic acid (ABA) in mature Scots pine during a gradually imposed drought. Tree Physiology, 22, 877–883.CrossRefGoogle ScholarPubMed
Peschel, S., Beyer, M. and Knoche, M. (2003). Surface characteristics of sweet cherry fruit: stomata-number, distribution, functionality and surface wetting. Scientia Horticulturae, 97, 265–278.CrossRefGoogle Scholar
Peterhansel, C., Niessen, M. and Kebeish, R.M. (2008). Metabolic Engineering towards the enhancement of photosynthesis. Photochemistry and Photobiology, 84, 1317–1323.CrossRefGoogle ScholarPubMed
Peterson, B.J. and Fry, B. (1987). Stable isotopes in ecosystem studies. Annual Review of Ecology and Systematics, 18, 293–320.CrossRefGoogle Scholar
Peterson, R.B. and Aylor, D.E. (1995). Chlorophyll fluorescence induction in leaves of Phaseolus vulgaris infected with bean rust (Uromyces appendiculatus). Plant Physiology, 108, 163–171.CrossRefGoogle Scholar
Peterson, R.B., Oja, V. and Laisk, A. (2001). Chlorophyll fluorescence at 680 and 730 nm and leaf photosynthesis. Photosynthesis Research, 70, 185–196.CrossRefGoogle ScholarPubMed
Peterson, R.K.D. (2001). Photosynthesis, yield loss and injury guilds. In: Biotic Stress and Yield Loss (eds Peterson, R.K.D. and Higley, L.G.), CRC Press, Boca Raton, FL, USA, pp. 83–97.Google Scholar
Peterson, R.K.D., Higley, L.G., Haile, F.J., et al. (1998). Mexican bean beetle (Coleoptera: Coccinelidae) injury affects photosynthesis of Glycine max and Phaseolus vulgaris. Environmental Entomology, 27, 373–381.CrossRefGoogle Scholar
Peterson, R.K.D., Shannon, C.L. and Lenssen, A.W. (2004). Photosynthetic responses of legume species to leaf-mass consumption injury. Environmental Entomology, 33, 450–456.CrossRefGoogle Scholar
Pethybridge, S.J., Haya, F., Eskerb, P., et al. (2008). Visual and radiometric assessments for yield losses caused by ray blight in Pyrethrum. Crop Science, 48, 343–352.CrossRefGoogle Scholar
Petit, A.N., Vaillant, N., Boulay, M., et al. (2006). Alteration of photosynthesis in grapevines affected by Esca. Phytopathology, 96, 1060–1066.CrossRefGoogle ScholarPubMed
Pfannschmidt, T. (2003). Chloroplast redox signals: how photosynthesis controls its own genes. Trends in plant science, 8, 33–41.CrossRefGoogle ScholarPubMed
Pfannschmidt, T., Brautigam, K., Wagner, R., et al. (2009). Potential regulation of gene expression in photosynthetic cells by redox and energy state: approaches towards better understanding. Annals of Botany, 103, 599–607.CrossRefGoogle ScholarPubMed
Pfanz, H., Aschan, G., Langenfeld-Heyser, R., et al. (2002). Ecology and ecophysiology of tree stems: corticular and wood photosynthesis. Naturwissenschaften, 89, 147–162.Google ScholarPubMed
Pfeffer, M. and Peisker, M. (1998). CO2 gas exchange and phosphoenolpyruvate carboxylase activity in leaves of Zea mays L. Photosynthesis Research, 58, 281–291.CrossRefGoogle Scholar
Pfündel, E. (1998). Estimating the contribution of photosystem I to total leaf chlorophyll fluorescence. Photosynthesis Research, 56, 185–195.CrossRefGoogle Scholar
Pfündel, E.E., Ben Ghoslen, N., Meyer, S., et al. (2007). Investigating UV screening in leaves by two different types of portable UV fluorimeters reveals in vivo screening by anthocyanins and carotenoids. Photosynthesis Research, 93, 205–221.CrossRefGoogle Scholar
Phene, C., Baker, D., Lambert, J., et al. (1978). SPAR – Soil-Plant-Atmosphere Research System. Transactions of the ASAE, 21, 924–30.CrossRefGoogle Scholar
Phillips, N.G., Buckley, T.N. and Tissue, D.T. (2008). Capacity of old trees to respond to environmental change. Journal of Integrative Plant Biology, 50, 1355–1364.CrossRefGoogle ScholarPubMed
Piel, C., Frank, E., Le Roux, X., et al. (2002). Effect of local irradiance on CO2 transfer conductance of mesophyll in walnut. Journal of Experimental Botany, 53, 2423–2430.CrossRefGoogle Scholar
Pielke, Sr. R.A., Adegoke, J.O., Chase, T.N., et al. (2007). A new paradigm for assessing the role of agriculture in the climate system and in climate change. Agricultural and Forest Meteorology, 142, 234–254.CrossRefGoogle Scholar
Pierce, J., Tolbert, N.E. and Barker, R. (1980). Interaction of Rubisco with transition state analogues. Biochemistry, 19, 934–962.CrossRefGoogle ScholarPubMed
Pieruschka, R., Chavarría-Krauser, A., Cloos, K., et al. (2008). Photosynthesis can be enhanced by lateral CO2 diffusion inside leaves over distances of several millimetres. New Phytologist, 178, 335–347.CrossRefGoogle Scholar
Pieruschka, R., Huber, G. and Berry, J.A. (2010) Control of transpiration by radiation. Proceedings of the National Academy of Sciences USA, 107, 13372–1337.CrossRefGoogle ScholarPubMed
Pieruschka, R., Schurr, U. and Jahnke, S. (2005). Lateral gas diffusion inside leaves. Journal of Experimental Botany, 56, 857–864.CrossRefGoogle ScholarPubMed
Pieters, A.J. and Núñez, M. (2008). Photosynthesis, water use efficiency, and δ13C in two rice genotypes with contrasting response to water deficit. Photosynthetica, 46, 574–580.CrossRefGoogle Scholar
Pieters, A.J., Paul, M.J. and Lawlor, D.W. (2001). Low sink demand limits photosynthesis under Pi deficiency. Journal of Experimental Botany, 52, 1083–1091.CrossRefGoogle Scholar
Pietrini, F. and Massacci, A. (1998). Leaf anthocyanin content changes in Zea mays L grown at low temperature: Significance fort he relationship between the quantum yield of PSII and the apparent quantum yield of CO2 assimilation. Photosynthesis Research, 58, 213–219.CrossRefGoogle Scholar
Pincebourd, S., Frak, E., Sinoquet, H., et al. (2006). Herbivory mitigation through increased water-use efficiency in a leaf-mining moth-apple tree relationship. Plant Cell and Environment, 29, 2238–2247.CrossRefGoogle Scholar
Pinder, J.E. and Jackson, P.R. (1988). Plant photosynthetic pathways and grazing by phytophagous orthopterans. American Midland Naturalist, 120, 201–211.Google Scholar
Pinder, J.E. and Kroh, G.C. (1987). Insect herbivory and photosynthetic pathways in old-field ecosystems. Ecology, 68, 254–259.CrossRefGoogle Scholar
Pineda, M., Gáspár, L., Morales, F., et al. (2008a). Multicolour fluorescence imaging: a useful tool to visualise systemic viral infections in plants. Photochemistry and Photobiology, 84, 1048–1060.CrossRefGoogle Scholar
Pineda, M., Soukupová, J., Matouš, K., et al. (2008b). Conventional and combinatorial chlorophyll fluorescence imaging of tobamovirus-infected plants. Photosynthetica, 46, 441–451.CrossRefGoogle Scholar
Pinelli, P. and Loreto, F. (2003). 12CO2 emission from different metabolic pathways measured in illuminated and darkened C3 and C4 leaves at low, atmospheric and elevated CO2 concentration. Journal of Experimental Botany, 54, 1761–1769.CrossRefGoogle ScholarPubMed
Pinheiro, C. and Chaves, M.M. (2011). Photosynthesis and drought: can we make metabolic connections from available data?Journal of Experimental Botany, 62, 869–882.CrossRefGoogle ScholarPubMed
Pinheiro, C., António, C., Ortuño, M.F. et al. (2011). Initial water deficit effects on Lupinus albus photosynthetic performance, carbon metabolism, and hormonal balance: metabolic reorganization prior to early stress responses. Journal of Experimental Botany, 62, 4965–4974.CrossRefGoogle Scholar
Pinkard, E.A. and Mohammed, C.L. (2006). Photosynthesis of Eucalyptus globulus with Mycosphaerella leaf disease. New Phytologist, 170, 119–127.CrossRefGoogle ScholarPubMed
Pitzschke, A. and Hirt, H. (2009). Disentangling the complexity of mitogen-activated protein kinases and reactive oxygen species signaling. Plant Physiology, 149, 606–615.CrossRefGoogle ScholarPubMed
Pizon, A. (1902). Anatomie et Physiologie Végétales. Doin Eds, Paris.
Plascyk, J.A. (1975). The MK II Fraunhofer line discriminator (FLD-II) for airborne and orbital remote sensing of solar-stimulated luminescence. Optical Engineering, 14, 339–346.CrossRefGoogle Scholar
Plaxton, W.C. (1996). The organization and regulation of plant glycolysis. Annual Review of Plant Physiology and Plant Molecular Biology, 47, 185–214.CrossRefGoogle ScholarPubMed
Plaziat, J.C., Cavagnetto, C., Koeniguer, J.C., et al. (2001). History and biogeography of the mangrove ecosystem, based on a critical reassessment of the paleontological record. Wetlands Ecology and Management, 9, 161–179.CrossRefGoogle Scholar
Plesnicar, M., Kastori, R., Petrovic, N., et al. (1994). Photosynthesis and chlorophyll fluorescence in sunflower (Helianthus annuus L.) leaves as affected by phosphorus nutrition. Journal of Experimental Botany, 45, 919–924.CrossRefGoogle Scholar
Poni, S., Bernizzoni, F., Civardi, S., et al. (2009). Performance and water-use efficiency (single-leaf vs. whole-canopy) of well-watered and half-stressed split-root Lambrusco grapevines grown in Po Valley (Italy). Agriculture, Ecosystems and Environment, 129, 97–106.CrossRefGoogle Scholar
Poni, S., Magnani, E. and Bernizzoni, F. (2003). Degree of correlation between total light interception and whole-canopy net CO2 exchange rate in two grapevine growth systems. Australian Journal of Grape and Wine Research, 9, 2–11.CrossRefGoogle Scholar
Pons, T.L. and Pearcy, R.W. (1994). Nitrogen reallocation and photosynthetic acclimation in response to partial shading in soybean plants. Physiologia Plantarum, 92, 636–644.CrossRefGoogle Scholar
Pons, T.L. and Welschen, R.A.M. (2002). Overestimation of respiration rates in commercially available clamp-on leaf chambers. Complications with measurement of net photosynthesis. Plant, Cell and Environment, 25, 1367–1372.CrossRefGoogle Scholar
Pons, T.L., Flexas, J., von Caemmerer, S., et al. (2009). Estimating mesophyl conductance to CO2: methodology, potential errors and recommendations. Journal of Experimental Botany, 60, 2217–2234.CrossRefGoogle Scholar
Pontailler, J.Y. (1990). A cheap quantum sensor using a gallium arsenide photodiode. Functional Ecology, 4, 591–596.CrossRefGoogle Scholar
Ponton, S., Flanagan, L.B., Alstad, K.P., et al. (2006). Comparison of ecosystem water-use efficiency among Douglas-fir forest, aspen forest and grassland using eddy covariance and carbon isotope techniques. Global Change Biology, 12, 294–310.CrossRefGoogle Scholar
Poormohammad, Kiani S., Grieu, P. et al. (2007). Genetic variability for physiological traits under drought conditions and differential expression of water stress-associated genes in sunflower (Helianthus annuus L.). Theoretical and Applied Genetics, 114, 193–207.CrossRefGoogle Scholar
Poorter, H. (1993). Interspecific variation in the growth response of plants to an elevated CO2 concentration. Vegetatio, 104/105, 77–97.CrossRefGoogle Scholar
Poorter, H. and Navas, M.L. (2003). Plant growth and competition at elevated CO2: On winners, losers and functional groups. New Phytologist, 157, 175–198.CrossRefGoogle Scholar
Poorter, H., Niinemets, Ü., Poorter, L., et al. (2009). Causes and consequences of variation in leaf mass per area (LMA): a meta-analysis. Tansley review. New Phytologist, 182, 565–588.CrossRefGoogle ScholarPubMed
Poorter, H., Remkes, C. and Lambers, H. (1990). Carbon and nitrogen economy of 24 wild species differing in relative growth rate. Plant Physiology, 94, 621–627.CrossRefGoogle ScholarPubMed
Poorter, L. and Oberbauer, S.F. (1993). Photosynthetic induction responses of 2 rain-forest tree species in relation to light environment. Oecologia, 96, 193–199.CrossRefGoogle ScholarPubMed
Poorter, L. and Werger, M.J.A. (1999). Light environment, sapling architecture and leaf display in six rain forest tree species. American Journal of Botany, 86, 1464–1473.CrossRefGoogle ScholarPubMed
Popper, Z.A. and Fry, S.C. (2003). Primary cell wall composition of bryophytes and charophytes. Annals of Botany, 91, 1–12.CrossRefGoogle ScholarPubMed
Popper, Z.A., Michel, G., Hervé, C. et al. (2011). Evolution and diversity of plant cell walls: from algae to flowering plants. Annual Review of Plant Biology, 62, 567–590.CrossRefGoogle ScholarPubMed
Porcar-Castell, A., Juurola, E., Berninger, F., et al. (2008). Seasonal acclimation of photosystem II in Pinus sylvestris. Studying the effect of light environment through the rate constants of sustained heat dissipation and photochemistry I. Tree Physiology, 28, 1483–1491.CrossRefGoogle Scholar
Porembski, S. and Barthlott, W. eds. (2000). Inselbergs. Ecological Studies, vol. 146, Springer-Verlag, Berlin – Heidelberg-New York.CrossRefGoogle Scholar
Port, M., Tripp, J., Zielinski, D., Weber, C., et al. (2004). Role of Hsp17.4-CII as coregulator and cytoplasmic retention factor of tomato heat stress transcription factor HsfA2 1. Plant Physiology, 135, 1457–1470.CrossRefGoogle Scholar
Portis, A.R., Salvucci, M.E. and Ogren, W.L. (1986). Activation of ribulosebisphosphate carboxylase/oxygenase at physiological CO2 and ribulosebisphosphate concentrations by rubisco activase. Plant Physiology, 82, 967–971.CrossRefGoogle Scholar
Portis, Jr., A.R., Li, C., Wang, D., et al. (2008). Regulation of Rubisco activase and its interaction with Rubisco. Journal of Experimental Botany, 59, 1597–1604.CrossRefGoogle ScholarPubMed
Portsmuth, A., Niinemets, Ü., Truus, L., et al. (2005). Biomass allocation and growth rates in Pinus sylvestris are interactively modified by nitrogen and phosphorus availabilities and by tree size and age. Canadian Journal of Forest Research, 35, 2346–2359.CrossRefGoogle Scholar
Pospisil, P., Skotnica, J. and Naus, J. (1998). Low and high temperature dependence of minimum F-O and maximum F-M chlorophyll fluorescence in vivo. Biochimica et Biophysica Actah, 1363, 95–99.CrossRefGoogle Scholar
Postuka, J.W., Dropkin, V.H. and Nelson, C.J. (1986). Photosynthesis, photorespiration, and respiration of soybean after infection with root nematodes. Photosynthetica, 20, 405–410.Google Scholar
Pou, A., Flexas, J., Alsina, M.M., Bota, J., et al. (2008). Adjustments of water-use efficiency by stomatal regulation during drought and recovery in the drought-adapted Vitis hybrid Richter-110 (V. berlandieri × V. rupestris). Physiologia Plantarum, 134, 313–323.CrossRefGoogle Scholar
Poulson, M.E. and Vogelmann, T.C. (1990). Epidermal focusing and effects upon photosynthetic light-harvesting in leaves of Oxalis. Plant, Cell and Environment, 13, 803–811.CrossRefGoogle Scholar
Poulson, M.E., Torres-Boeger, M.R. and Donahue, R.A. (2006). Response of photosynthesis to high light and drought for Arabidopsis thaliana grown under a UV-B enhanced light regime. Photosynthesis Research, 90, 79–90.CrossRefGoogle Scholar
Pozsar, B.I., Horvath, L., Lehoczky, J., et al. (1969). Effect of the grape chromemosaic and grape fanleaf yellow-mosaic virus infection on the photosynthetical carbon dioxide fixation in vine leaves. Vitis, 8, 206–10.Google Scholar
Prändl, R., Hinderhofer, K., Eggers-Schumacher, G., et al. (1998). HSF3, a new heat shock factor from Arabidopsis thaliana, derepresses the heat shock response and confers thermotolerance when overexpressed in transgenic plants. Molecular and General Genetics, 258, 269–278.Google ScholarPubMed
Prasad, D.D.K. and Prasad, A.R.K. (1987). Altered d-aminolevulinic acid metabolism by lead and mercury in germinating seedlings of bajra (Pennisetum typhoideum). Journal of Plant Physiology, 127, 241–249.CrossRefGoogle Scholar
Prasad, M.N.V. and Strzalka, K. (1999). Impact of heavy metals on photosynthesis. In Heavy Metal Stress in Plants: From Molecules to Ecosystems (eds Prasad, M.N.V. and Hagemeyer, J.), Springer, Berlin, Germany, pp. 117–138.CrossRefGoogle Scholar
Prasad, T.K. (1996). Mechanisms of chilling-induced oxidative stress injury and tolerance in developing maize seedlings: changes in antioxidant system, oxidation of proteins and lipids, and protease activities. Plant Journal, 10, 1017–1026.CrossRefGoogle Scholar
Prendergast, H.D.V. (1989). Geographical distribution of C4 acid decarboxylation types and associated structural variants in native Australian C4 grasses (Poaceae). Australian Journal of Botany, 37, 253–273.CrossRefGoogle Scholar
Prentice, I.C., Farquhar, G.D., Fasham, M.J.R., et al. (2001). The carbon cycle and atmospheric carbon dioxide. In: Climate Change 2001: The Scientific Basis (eds Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M., van der Linden, P.J., Dai, X., Maskell, K. and Johnson, C.A.), Cambridge University Press, Camridge, UK, pp. 183–237.Google Scholar
Priault, P., Tcherkez, G., Cornic, G., et al. (2006). The lack of mitochondrial complex I in a CMSII mutant of Nicotiana sylvestris increases photorespiration through an increased internal resistance. Journal of Experimental Botany, 57, 3195–3207.CrossRefGoogle Scholar
Price, A.H., Cairns, J.E., Horton, P., et al. (2002). Linking drought resistance mechanisms to drought avoidance in upland rice using a QTL approach: progress and new opportunities to integrate stomatal and mesophyll responses. Journal of Experimental Botany, 53, 989–1004.CrossRefGoogle ScholarPubMed
Price, G.D., Von Caemmerer, S., Evans, J.R., et al. (1994). Specific reduction of chloroplast carbonic anhydrase activity by antisense RNA in transgenic tobacco plants has a minor effect on photosynthetic CO2 assimilation. Planta, 193, 331–340.CrossRefGoogle Scholar
Price, P., Waring, G.L., Julkunen-Titto, R., et al. (1989). Carbon-nutrient balance hypothesis in within-species phytochemical variation of Salix lasiolepis. Journal of Chemical Ecology, 15, 1117–1131.CrossRefGoogle ScholarPubMed
Prins, A., Van Heerden, P.D.R., Olmos, E., et al. (2008). Cysteine proteases regulate chloroplast protein content and composition in tobacco leaves: a model for dynamic interactions with ribulose-1, 5-bisphosphate carboxylase-oxygenase (Rubisco) vesicular bodies. Journal of Experimental Botany, 59, 1935–1950.CrossRefGoogle ScholarPubMed
Prins, H.B.A. and de Guia, M.B. (1986). Carbon source of the water soldier (Stratiotes aloides L.). Aquatic Botany, 26, 225–234.CrossRefGoogle Scholar
Prinsley, R.T., Hunt, S., Smith, A.M., et al. (1986). The influence of a decrease in irradiance on photosynthetic carbon assimilation in leaves of Spinacia oleracea L. Planta, 167, 414–420.CrossRefGoogle ScholarPubMed
Prior, L.D., Eamus, D. and Duff, G.A. (1997). Seasonal and diurnal patterns of carbon assimilation, stomatal conductance and leaf water potential in Eucalyptus tetrodonta saplings in a wet–dry savanna in northern Australia. Australian Journal of Botany, 45, 241–258.CrossRefGoogle Scholar
Proctor, M.C.F. (1984). Structure and ecological adaptation. In: The Experimental Biology of Bryophytes, Experimental Botany: An International Series of Monographs (eds Dyer, A.F., Duckett, J.G. and Cronshaw, J.), Academic Press, London, pp. 9–37.Google Scholar
Proctor, M.C.F. (2000). Mosses and Alternative Adaptation to Life on Land. New Phytologist, 148, 1–3.CrossRefGoogle Scholar
Proctor, M.C.F. and Pence, V. (2002). Vegetative tissues: bryophytes, vascular resurrection plants and vegetative propagules. In: Black, M., Pritchard, H.V., eds. Dessication and Survival in Plants: Drying without Dying. Wallingford, UK: CAB International, 207–237.CrossRef
Proietti, P., Famiani, F. and Tombessi, A. (1999). Gas exchange in olive fruit. Photosynthetica, 36, 423–432.CrossRefGoogle Scholar
Prud’homme, B., Gompel, N. and Carroll, S.B. (2007). Emerging principles of regulatory evolution. Proceedings of the National Academy of Sciences, 104, 8605–8612.CrossRefGoogle ScholarPubMed
Pruitt, W.O., Swann, B.D., Held, A.A., et al. (1987). Bowen ratio and Penman: Australia-California tests. In: Irrigation System for the 21st Century (eds James, L.G. and English, M.J), American Society of Civil Engineers, Portland, USA, pp.149–158.Google Scholar
Pryer, K.M., Schneider, H., Smith, A.R., et al. (2001). Horsetails and ferns are a monophyletic group and the closest living relatives to seed plants. Nature, 409, 618–621.CrossRefGoogle ScholarPubMed
Psaras, G.K. and Rhizopoulou, S. (1995). Mesophyll structure during leaf development in Ballota acetabulosa. New Phytologist, 131, 303–309.CrossRefGoogle Scholar
Puckeridge, D.W. (1971). Photosynthesis of wheat under field conditions III. Seasonal trends in carbon dioxide uptake of crop communities. Australian Journal of Agriultural Research, 22, 1–9.CrossRefGoogle Scholar
Pugnaire, F.I., Haase, P., Incoll, L.D., et al. (1996). Response of the tussock grass Stipa tenacissima to watering in a semi-arid environment. Functional Ecology, 10, 265–274.CrossRefGoogle Scholar
Puthiyaveetil, S., Kavanagh, T.A., Cain, P., et al. (2008). The ancestral symbiont sensor kinase CSK links photosynthesis with gene expression in chloroplasts. Proceedings of the National Academy of Sciences (USA), 105, 10061–10066.CrossRefGoogle ScholarPubMed
Pyankov, V.I., Gunin, P.D., Tsoog, S., et al. (2000). C4 plants in the vegetation of Mongolia: their natural occurrence and geographical distribution in relation to climate. Oecologia, 123, 15–31.CrossRefGoogle ScholarPubMed
Pyke, K.A. (1999). Plastid division and development. Plant Cell, 11, 549–556.CrossRefGoogle ScholarPubMed
Pyke, K.A., Rutherford, S.M., Robertson, E.J., et al. (1994). arc6, a fertile Arabidopsis mutant with only two mesophyll cell chloroplasts. Plant Physiology, 106, 1169–1177.CrossRefGoogle ScholarPubMed
Qian, H. and Ricklefs, R.E. (2000). Large-scale processes and the Asian bias in species diversity of temperate plants. Nature, 407, 180–182.Google ScholarPubMed
Quade, J. and Cerling, T.E. (1995). Expansion of C4 grasses in the late Miocene of northern Pakistan: evidence from stable isotopes in paleosols. Palaeogeography, Palaeoclimatology, Palaecoecology, 115, 91–116.CrossRefGoogle Scholar
Quade, J., Cater, J.M.L., Ojha, T.P., et al. (1995). Late Miocene environmental change in Nepal and the northern Indian subcontinent: stable isotopic evidence from paleosols. GSA Bull., 107, 1381–1397.2.3.CO;2>CrossRefGoogle Scholar
Quarrie, S.A. and Jones, H.G. (1977). Effect of abscisic acid and water stress on development and morphology of wheat. Journal of Experimental Botany, 28, 192–203.CrossRefGoogle Scholar
Queiroz-Voltan, R.B. and Paradela-Filho, O. (1999). Caracterização de estruturas anatômicas de citros infectados com Xylella fastidiosa. Laranja, 20, 55–76.Google Scholar
Queitsch, C., Hong, S.W., Vierling, E., et al. (2000). Heat shock protein 101 plays a crucial role in thermotolerance in Arabidopsis. The Plant Cell, 12, 479.CrossRefGoogle Scholar
Queitsch, C., Sangster, T.A. and Lindquist, S. (2002). Hsp90 as a capacitor for genetic variation. Nature, 417, 618–624.CrossRefGoogle Scholar
Querejeta, J.I., Allen, M.F., Alguacil, M.M., et al. (2007). Plant isotopic composition provides insight into mechanisms underlying growth stimulation by AM fungi in a semiarid environment. Functional Plant Biology, 34, 683–691.CrossRefGoogle Scholar
Querejeta, J.I., Barea, J.M., Allen, M.F., et al. (2003). Differential response of δ13C and water use efficiency to arbuscular mycorrhizal infection in two aridland woody plant species. Oecologia, 135, 510–515.CrossRefGoogle ScholarPubMed
Queval, G., Hager, J., Gakière, B., et al. (2008). Why are literature data for H2O2 contents so variable? A discussion of potential difficulties in quantitative assays of leaf extracts. Journal of Experimental Botany, 59, 135–146.CrossRefGoogle Scholar
Queval, G., Issakidis-Bourguet, E., Hoeberichts, F.A., et al. (2007). Conditional oxidative stress responses in the Arabidopsis photorespiratory mutant cat2 demonstrate that redox state is a key modulator of day length-dependent gene expression and define photoperiod as a crucial factor in the regulation of H2O2-induced cell death. Plant Journal, 52, 640–657.CrossRefGoogle Scholar
Quick, W.P. and Stitt, M. (1989). An examination of factors contributing to non-photochemical quenching of chlorophyll fluorescence in barley leaves. Biochimica et Biophysica Acta, 977, 287–296.CrossRefGoogle Scholar
Quinn, P.J. (1988). Effects of temperature on cell membranes. In: Plant Membrane Stability (ed. Society for Experimental Biology), pp. 237–258.
Quinn, P.J., Williams, W.P., Barber, J., et al. (1985). Environmentally induced changes in chloroplast membranes and their effects on photosynthetic function. In: Photosynthetic Mechanisms and the Environment (eds Barber, J. and Baker, N.J.), Elsevier Science Publishers, Dordrecht, Netherlands, pp. 1–47.Google Scholar
Quirino, B.F., Noh, Y.S., Himelblau, E., et al. (2000). Molecular aspects of leaf senescence. Trends Plant Science, 5, 278–282.CrossRefGoogle ScholarPubMed
Rachmilevitch, S., Cousins, A.B. and Bloom, A.J. (2004). Nitrate assimilation in plant shoots depends on photorespiration. Proceedings of the Natural Academy of Sciences USA, 101, 11506–11510.CrossRefGoogle ScholarPubMed
Raftoyannis, Y. and Radoglou, K. (2002). Physiological responses of beech and sessile oak in a natural mixed stand during a dry summer. Annals of Botany, 89, 723–730.CrossRefGoogle Scholar
Raghavendra, A.S. and Padmasree, K. (2003). Beneficial interactions of mitochondrial metabolism with photosynthetic carbon assimilation. Trends Plant Science, 8, 546–553.CrossRefGoogle ScholarPubMed
Rahman, A.F., Cordova, V.D., Gamon, J.A., et al. (2004). Potential of MODIS ocean bands for estimating CO2 flux from terrestrial vegetation: a novel approach. Geophysical Research Letters, 31, L10503.CrossRefGoogle Scholar
Rahman, A.F., Gamon, J.A., Fuentes, D.A., et al. (2001). Modeling spatially distributed ecosystem flux of boreal forest using hyperspectral indices from AVIRIS imagery. Journal of Geophysical Research – Atmospheres, 106, 33579–33591.CrossRefGoogle Scholar
Rahoutei, J., Barón, M., García-Luque, I., et al. (1999). effect of tobamovirus infection on the thermoluminiscence characteristics of chloroplast from infected plants. Z. Naturforsch, 54c, 634–639.Google Scholar
Rahoutei, J., García-Luque, I. and Barón, M. (2000). Inhibition of photosynthesis by viral infection: effect on PSII structure and function. Physiologia Plantarum, 110, 286–292.CrossRefGoogle Scholar
Raines, C.A. (2006). Transgenic approaches to manipulate the environmental responses of the C3 carbon fixation cycle. Plant, Cell and Enviaronment, 29, 331–339.CrossRefGoogle ScholarPubMed
Raison, J.K., Roberts, J.K.M. and Berry, J.A. (1982). Correlations between the thermal stability of chloroplast (thylakoid) membranes and the composition and fluidity of their polar lipids upon acclimation of the higher plant, Nerium oleander, to growth temperature. Biochimica et Biophysica Acta, 688, 218–228.CrossRefGoogle Scholar
Rajabi, A., Ober, E.S. and Griffiths, H. (2009). Genotypic variation for water use efficiency, carbon isotope discrimination, and potential surrogate measures in sugar beet. Field Crops Research, 112, 172–181.CrossRefGoogle Scholar
Rakocevic, M., Sinoquet, H., Christophe, A., et al. (2000). Assessing the geometric structure of a white clover (Trifolium repens L.) canopy using 3-D digitising. Annals of Botany, 86, 519–526.CrossRefGoogle Scholar
Ramalho, J.C., Campos, P.S., Teixeira, M., et al. (1998). Nitrogen dependent changes in antioxidant system and in fatty acid composition of chloroplast membranes from Coffea arabica L. plants submitted to high irradiance. Plant Science, 135, 115–124.CrossRefGoogle Scholar
Rambal, S. (2001). Hierarchy and productivity of Mediterranean-type ecosystems. In: Global Terrestrial Productivity (eds Roy, J., Saugier, B. and Mooney, H.A.), Academic Press, San Diego, USA, pp. 315–344.Google Scholar
Ranieri, A., Castagna, A., Baldan, B., et al. (2001). Iron deficiency differently affects peroxidase isoforms in sunflower. Journal of Experimental Botany, 52, 25–35.CrossRefGoogle ScholarPubMed
Ranjith, S.A., Meinzer, F.S., Perry, M.H., et al. (1995). Partitioning of carboxylase activity in nitrogen stressed sugarcane and its relationship to bundle sheath leakiness to CO2, photosynthesis and carbon isotope discrimination. Australian Journal of Plant Physiology, 22, 903–911.CrossRefGoogle Scholar
Rao, D.N. (1971). A study of the air pollution problem due to coal unloading in Varanasi, India. In: Proceedings of the Second International Clean Air Congress (eds Englund, H.M. and Beery, W.T.), Academic Press, New York, USA, pp. 273–276.Google Scholar
Rao, I.M., Abadía, A. and Terry, N. (1986). Leaf phosphate status and photosynthesis in vivo: changes in light scattering and chlorophyll fluorescence during photosynthetic induction in sugar beet leaves. Plant Science, 44, 133–137.Google Scholar
Rao, I.M. and Terry, N. (1989). Leaf phosphate status, photosynthesis and carbon partitioning in sugar beet. I. Changes in growth, gas exchange and Calvin cycle enzymes. Plant Physiology, 90, 814–819.CrossRefGoogle ScholarPubMed
Rao, I.M. and Terry, N. (1995). Leaf phosphate status, photosynthesis, and carbon partitioning in sugar beet. IV: Changes with time following increased supply of phosphate to low phosphate plants. Plant Physiology, 107, 1313–1321.CrossRefGoogle ScholarPubMed
Rascher, U. and Pieruschka, R. (2008). Spatio-temporal variations of photosynthesis: the potential of optical remote sensing to better understand and scale light use efficiency and stresses of plant ecosystems. Precision Agriculture, 9, 355–366.CrossRefGoogle Scholar
Rascher, U., Agati, G., Alonso, L., et al. (2009). CEFLES2: the remote sensing component to quantify photosynthetic efficiency from the leaf to the region by measuring sun-induced fluorescence in the oxygen absorption bands. Biogeosciences Discussions, 6, 2217–2266.CrossRefGoogle Scholar
Rascher, U., Hütt, M.Th., Siebke, K., et al. (2001). Spatiotemporal variation of metabolism in a plant circadian rhythm. The biological clock as an assembly of coupled individual oscillators. Proceedings of the Natural Academy of Sciences USA, 98, 11, 801–11, 805.CrossRefGoogle Scholar
Rascher, U., Liebig, M. and Lüttge, U. (2000). Evaluation of instant light-response curves of chlorophyll fluorescence parameters obtained with a portable chlorophyll fluorometer on site in the field. Plant, Cell and Environment, 23, 1397–1405.CrossRefGoogle Scholar
Raschke, K. and Resemann, A. (1986). Midday depression of CO2 assimilation in leaves of Arbutus unedo L.: diurnal changes in photosynthetic capacity related to changes in temperature and humidity. Planta, 168, 546–58.CrossRefGoogle ScholarPubMed
Raschke, K. (1956). Über die physikalischen beziehungen zwischen wämeübergangszahl, strahlungsaustausch, temperatur and transpiration eines blattes. Planta, 48, 200–238.CrossRefGoogle Scholar
Rasmusson, A.G. and Escobar, M.A. (2007). Light and diurnal regulation of plant respiratory gene expression. Physiologia Plantarum, 129, 57–67.CrossRefGoogle Scholar
Rastetter, E.B., Agren, G.I. and Shaver, G.R. (1997). Responses of N-limited ecosystems to increased CO2: A balanced-nutrition, coupled-element-cycles model. Ecological Applications, 7, 444–460.Google Scholar
Rasulov, B., Copolovici, L., Laisk, A., et al. (2009). Postillumination isoprene emission: in vivo measurements of dimethylallylpyrophosphate pool size and isoprene synthase kinetics in aspen leaves. Plant Physiology, 149, 1609–1618.CrossRefGoogle Scholar
Rauner, J.L. (1976). Deciduous forests. In: Vegetation and the Atmosphere, Vol 2. Case Studies (ed. Monteith, J.L.), Academic Press, London – New York – San Francisco, pp. 241–264.Google Scholar
Raupach, M.R., Marland, G.Ciais, P., et al. (2007). Global and Regional drivers of accelerating CO2 emissions. Proceedings of the National Academy of Science, 104, 10288–10293.CrossRefGoogle ScholarPubMed
Raupach, M.R., Rayner, P.J., Barrett, D.J., et al. (2005). Model–data synthesis in terrestrial carbon observation: methods, data requirements and data uncertainty specifications. Global Change Biology, 11, 378–397.CrossRefGoogle Scholar
Raven, J.A. (2002b). Evolutionary options. Nature, 415, 375–376.CrossRefGoogle ScholarPubMed
Raven, J.A. (2000). Land plant biochemistry. Philosophical Transactions of the Royal Society of London B, 355, 833–846.CrossRefGoogle ScholarPubMed
Raven, J.A. (2002a). Selection pressures on stomatal evolution. The New Phytologist, 153, 371–386.CrossRefGoogle Scholar
Raven, J.A. and Allen, J.F. (2003). Genomics and chloroplast evolution: what did cyanobacteria do for plants?Genome Biology, 4, 209.CrossRefGoogle ScholarPubMed
Raven, J.A. and Farquhar, G.D. (1990). The influence of N metabolism and organic acid synthesis on the natural abundance of isotopes of carbon in plants. New Phytologist, 116, 505–529.CrossRefGoogle Scholar
Raven, J.A. and Glidewell, S.M. (1981). Processes limiting photosynthetic conductance. In: Processes Limiting Plant Productivity (ed. Johnson, C.B.), Butterworths, London, UK, pp. 109–136.Google Scholar
Raven, J.A. and Spicer, R.A. (1996). The evolution of crassulacean acid metabolism. In: Crassulacean Acid Metabolism. Biochemistry, Ecophysiology and Evolution (eds Winter, K. and Smith, J.A.C.), Springer, New York, USA, pp. 360–385.Google Scholar
Raven, J.A., Cockell, C.S. and De La Roche, C.L. (2008). The evolution of inorganic carbon concentrating mechanisms in photosynthesis. Philosophical Transactions of the Royal Society, B, 363, 2641–2650.CrossRefGoogle ScholarPubMed
Ravenel, J., Peltier, G. and Havaux, M. (1994). The cyclic electron pathways around photosystem I in Chlamydomonas reinhardtii as determined in vivo by photoacoustic measurements of energy storage. Planta, 193, 251–259.CrossRefGoogle Scholar
Rawat, A.S. and Purohit, A.N. (1991). CO2 and water vapour exchange in four alpine herbs at two altitudes and under varying light and temperature conditions. Photosynthesis Research, 28, 99–108.CrossRefGoogle ScholarPubMed
Rawsthorne, S., Hylton, C.M., Smith, A.M., et al. (1988). Photorespiratory metabolism and immunogold localisation of photorespiratory enzymes in leaves of C3 and C3-C4 intermediate species of Moricandia. Planta, 173, 298–308.CrossRefGoogle Scholar
Rawsthorne, S., Morgan, C.L., O’Neill, C.M., et al. (1998). Cellular expression pattern of the glycine decarboxylase P protein in leaves of an intergenic hybrid between the C3-C4 intermediate species Moricandia nitens and the C3 species Brassica napus. Theoretical and Applied Genetics, 96, 922–927.CrossRefGoogle Scholar
Read, J., Sanson, G.D., de Garine-Wichatitsky, M., et al. (2006). Sclerophylly in two contrasting tropical environments: low nutrients and low rainfall. American Journal of Botany, 93, 1601–1614.CrossRefGoogle ScholarPubMed
Rebetzke, G.J., Condon, A.G., Richards, R.A., et al. (2002). Selection for reduced carbon isotope discrimination increases aerial biomass and grain yield of rainfed bread wheat. Crop Science, 42, 739–745.CrossRefGoogle Scholar
Reddy, K.R., Hodges, H.F., Read, J.J., et al. (2001). Soil-Plant-Atmosphere-Research (SPAR) facility: a tool for plant research and modeling. Biotronics, 30, 27–50.Google Scholar
Reich, P.B. (1993). Reconciling apparent discrepancies among studies relating life-span, structure and function of leaves in contrasting plant life forms and climates – the blind men and the elephant retold. Functional Ecology, 7, 721–725.CrossRefGoogle Scholar
Reich, P.B. and Amundson, R.G. (1985). Ambient levels of ozone reduce net photosynthesis in tree and crop species. Science, 230, 566–570.CrossRefGoogle ScholarPubMed
Reich, P.B. and Bolstad, P. (2001). Productivity of evergreen and deciduous temperature forests. In: Terrestrial Global Productivity (eds Mooney, H.A., Saugier, B. and Roy, J.). Academic Press, San Diego, USA, pp. 245–283.Google Scholar
Reich, P.B., Ellsworth, D.S. and Walters, M.B. (1998a). Leaf structure (specific leaf area) modulates photosynthesis-nitrogen relations: evidence from within and across species and functional groups. Functional Ecology, 12, 948–958.CrossRefGoogle Scholar
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., Kloeppel, B.D., Ellsworth, D.S., et al. (1995).Different photosynthesis-nitrogen relations in deciduous hardwood and evergreen coniferous tree species. Oecologia, 104, 24–30.CrossRefGoogle ScholarPubMed
Reich, P.B., Tjoelker, M.G., Machado, J.L., et al. (2006). Universal scaling of respiratory metabolism, size and nitrogen in plants. Nature, 439, 457–461.CrossRefGoogle ScholarPubMed
Reich, P.B., Uhl, C., Walters, M.B., et al. (2004). Leaf demography and phenology in Amazonian rain forest: A census of 40 000 leaves of 23 tree species. Ecological Monographs, 74, 3–23.CrossRefGoogle Scholar
Reich, P.B., Walters, M.B. and Ellsworth, D.S. (1991). Leaf age and season influence the relationships between leaf nitrogen, leaf mass per area and photosynthesis in maple and oak trees. Plant Cell Environment, 14, 251–259.CrossRefGoogle Scholar
Reich, P.B., Walters, M.B. and Ellsworth, D.S. (1997). From tropics to tundra: global convergence in plant functioning. Proceedings of the National Academy of Science of the USA, 94, 13730–13734.CrossRefGoogle ScholarPubMed
Reich, P.B., Walters, M.B. and Ellsworth, D.S. (1992). Leaf life-span in relation to leaf, plant, and stand characteristics among diverse ecosystems. Ecological Monographs, 62, 365–392.CrossRefGoogle Scholar
Reich, P.B., Walters, M.B., Ellsworth, D.S., et al. (1998b). 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
Reich, P.B., Walters, M.B., Tjoelker, M.G., et al. (1998c). Photosynthesis and respiration rates depend on leaf and root morphology and nitrogen concentration in nine boreal tree species differing in relative growth rate. Functional Ecology, 12, 395–405.CrossRefGoogle Scholar
Reich, P.B., Wright, I.J., Cavender-Bares, J., et al. (2003). The evolution of plant functional variation: traits, spectra and strategies. International Journal of Plant Sciences, 164, s143–s164.CrossRefGoogle Scholar
Reich, P. B. and Schoettle, A.W. (1988). Role of phosphorus and nitrogen in photosynthetic and whole plant carbon gain and nutrient use efficiency in eastern white-pine. Oecologia, 77, 25–33.CrossRefGoogle ScholarPubMed
Reichle, D.E. (1981). Dynamic Properties of Forest Ecosystems. IBP 23, Cambridge University Press, Cambridge, UK, pp. 683.
Reichstein, M., Falge, E., Baldocchi, D., et al. (2005). On the separation of net ecosystem exchange into assimilation and ecosystem respiration: review and improved algorithm. Global Change Biology, 11, 1424–1439.CrossRefGoogle Scholar
Reichstein, M., Papale, D., Valentini, R., et al. (2007). Determinants of terrestrial ecosystem carbon balance inferred from European eddy covariance flux sites. Geophysical Research Letters, 34, L01402.CrossRefGoogle Scholar
Reichstein, M., Tenhunen, J.D., Roupsard, O., et al. (2002).Severe drought effects on ecosystem CO2 and H2O fluxes at three Mediterranean evergreen sites: revision of current hypothesis?Global Change Biology, 8, 999–1017.CrossRefGoogle Scholar
Reid, C.D. and Strain, B.R. (1994). Effects of CO2 enrichment on whole-plant carbon budget of seedlings of Fagus grandifolia and Acer saccharum in low irradiance. Oecologia, 98, 31–39.CrossRefGoogle ScholarPubMed
Reinbothe, C., Lebedev, N. and Reinbothe, S. (1999). A protochlorophyllide light-harvesting complex involved in de-etiolation of higher plants. Nature, 397, 80–84.CrossRefGoogle Scholar
Reinero, A. and Beachy, N.R. (1989). Reduced photosystem II activity and accumulation of viral coat protein in chloroplast of leaves infected with tobacco mosaic virus. Plant Physiology, 89, 111–116.CrossRefGoogle ScholarPubMed
Reinert, F., Roberts, A., Wilson, J.M., et al. (1997). Gradation in nutrient composition and photosynthetic pathways across the restinga vegetation of Brazil. Botanica Acta, 110, 135–142.CrossRefGoogle Scholar
Reinert, F., Russo, C.A.M. and Salles, L.O. (2003). The evolution of CAM in the subfamily Pitcairnioideae (Bromeliaceae). Biological Journal of the Linnean Society, 80, 261–268.CrossRefGoogle Scholar
Reinikäinen, J. and Huttunen, S. (1989). The level of injury and needle ultrastructure of acid rain-irrigated pine and spruce seedlings after low temperature treatment. New Phytologist, 112, 29–39.CrossRefGoogle Scholar
Reiter, R., Hötberger, M., Green, R., et al. (2008). Photosynthesis of lichens from lichen-dominated communities in the alpine/nival belt of the Alps – II: laboratory and field measurements of CO2 exchange and water relations. Flora, 203, 34–46.CrossRefGoogle Scholar
Rekika, D., Monneveux, P. and Havaux, M. (1997). The in vivo tolerance of photosynthetic membranes to high and low temperatures in cultivated and wild wheats of the Triticum and Aegilops genera. Journal of Plant Physiology, 150, 734–738.CrossRefGoogle Scholar
Renaut, J., Hausman, J.F. and Wisniewski, M.E. (2006). Proteomics and low-temperature studies: bridging the gap between gene expression and metabolism. Physiologia Plantarum, 126, 97–109.CrossRefGoogle Scholar
Rennenberg, H., Loreto, F., Polle, A., et al. (2006). Physiological responses of forest trees to heat and drought. Plant Biology, 8, 556–571.CrossRefGoogle ScholarPubMed
Renou, J.L., Gerbaud, A., Just, D., et al. (1990). Differing substomatal and chloroplastic concentrations in water-stressed wheat. Planta, 182, 415–419.CrossRefGoogle ScholarPubMed
Repka, V. (2002). Chlorophyll-deficient mutant in oak (Quercus petraea L.) displays an accelerated hypersensitive-like cell death and an enhanced resistance to powdery mildew disease. Photosynthetica, 40, 183–193.CrossRefGoogle Scholar
Repo, T., Leinonen, I., Ryyppö, A. and Finer, L. (2004). The effect of soil temperature on the bud phenology, chlorophyll fluorescence, carbohydrate content and cold hardiness of Norway spruce seedlings. Physiologia Plantarum, 121, 93–100.CrossRefGoogle ScholarPubMed
Resco, V., Ewers, B.E., Sun, W., et al. (2009). Drought-induced hydraulic limitations constrain leaf gas exchange recovery after precipitation pulses in the C3 woody legume, Prosopis velutina. The New Phytologist, 181, 672–682.CrossRefGoogle ScholarPubMed
Restom, T.G. and Nepstad, D.C. (2001). Contribution of vines to the evapotranspiration of a secondary forest in eastern Amazonia. Plant and Soil, 236, 155–163.CrossRefGoogle Scholar
Retallack, G.J. (1997). Earliest Triassic origin of Isoetes and quillwort evolutionary radiation. Journal of Paleontology, 71, 500–521.CrossRefGoogle Scholar
Retallack, G.J. (2001). Cenozoic expansion of grasslands and climatic cooling. Journal of Geology, 109, 407–426.CrossRefGoogle Scholar
Retallack, G.J. (2005). Permian greenhouse crises. In: The Nonmarine Permian (eds Lucas, S.G. and Zeigler, K.E.), New MexicoMuseum of Natural History and Science Bulletin nº 30, pp. 256–269.Google Scholar
Retuerto, R., Fernandez-Lema, B., Rodriguez-Roiloa, S., et al. (2004). Increased photosynthetic performance in holly trees infested by scale. Functional Ecology, 18, 664–669.CrossRefGoogle Scholar
Reumann, S. and Weber, A.P.M. (2006). Plant peroxisomes respire in the light: some gaps of the photorespiratory C2 cycle have become filled, other remain. Biochim Biophys Acta, 1763, 1496–1510.CrossRefGoogle ScholarPubMed
Reverberi, M., Fanelli, C., Zjalic, S., et al. (2005). Relationship among lipoperoxides, jasmonates and indole-3-acetic acid formation in potato tubers after wounding. Free Radical Research, 39, 637–647.CrossRefGoogle Scholar
Reynolds, M., Foulkes, M.J., Slafer, G.A., et al. (2009). Raising yield potential in wheat. Journal of Experimental Botany, 60, 1899–1918.CrossRefGoogle ScholarPubMed
Reynolds, O. (1895). On the dynamical theory of incompressible viscous fluids and the determination of criterion. Philosophical Transactions of Royal Society of London, A174, 935–982.Google Scholar
Rhizopoulou, S. and Psaras, G.K. (2003). Development and Structure of drought-tolerant leaves of the Mediterranean shrub Capparis spinosa L. Annals of Botany, 92, 377–383.CrossRefGoogle ScholarPubMed
Rhoads, D.M. and Subbaiah, C.C. (2007). Mitochondrial retrograde regulation in plants. Mitochondrion, 7, 177–194.CrossRefGoogle ScholarPubMed
Riaño, D., Vaughan, P., Chuvieco, E., et al. (2005). Estimation of fuel moisture content by inversion of radiative transfer models to simulate equivalent water thickness and dry matter content: analysis at leaf and canopy level. IEEE Transactions on Geoscience and Remote Sensing, 43, 819–826.CrossRefGoogle Scholar
Ribas-Carbo, M., Berry, J.A., Yakir, D., et al. (1995). Electron partitioning between the cytochrome and alternative pathways in plant mitochondria. Plant Physiology, 109, 829–837.CrossRefGoogle ScholarPubMed
Ribas-Carbo, M., Giles, L., Flexas, J., et al. (2008). Phytochrome-driven changes in respiratory electron transport partitioning in soybean (Glycine max. L.) cotyledons. Plant Biology, 10, 281–287.CrossRefGoogle ScholarPubMed
Ribas-Carbo, M., Lennon, A.M., Robinson, S.A., et al. (1997). The regulation of the electron partitioning between the cytochrome and alternative pathways in soybean cotyledon and root mitochondria. Plant Physiology, 113, 903–911.CrossRefGoogle ScholarPubMed
Ribas-Carbo, M., Robinson, S.A. and Giles, L. (2005a). The application of the oxygen-isotope technique to assess respiratory pathway partitioning. In: Plant Respiration: From Cell to Ecosystem (eds Lambers, H. and Ribas-Carbo, M.) Springer, USA, pp. 31–42.Google Scholar
Ribas-Carbo, M., Robinson, S.A., Gonzalez-Meler, M.A., et al. (2000). Effects of light on respiration and oxygen isotope fractionation in soybean cotyledons. Plant Cell Environment, 23, 983–989.CrossRefGoogle Scholar
Ribas-Carbo, M., Taylor, N.L., Giles, L., et al. (2005b). Effects of water stress on respiration in soybean (Glycine max. L.) leaves. Plant Physiology, 139, 466–473.CrossRefGoogle Scholar
Ribeiro, R.V., Machado, E.C. and Oliveira, R.F. (2003a). Early photosynthetic responses of sweet orange plants infected with Xylella fastidiosa. Physiological and Molecular Plant Pathology, 62, 167–173.CrossRefGoogle Scholar
Ribeiro, R.V., Machado, E.C. and Oliveira, R.F. (2004). Growth- and leaf-temperature effects on photosynthesis of sweet orange seedlings infected with Xylella fastidiosa. Plant Pathology, 53, 334–340.CrossRefGoogle Scholar
Ribeiro, R.V., Machado, E.C., Oliveira, R.F. et al. (2003b). High temperature effects on the response of photosynthesis to light in sweet orange plants infected with Xylella fastidiosa. Brazilian Journal of Plant Physiology, 15, 89–97.CrossRefGoogle Scholar
Rice, S.A. and Bazzaz, F.A. (1989). Quantification of plasticity of plant traits in response to light intensity: comparing phenotypes at a common weight. Oecologia, 78, 502–507.CrossRefGoogle Scholar
Rich, P.R. (1988). A critical examination of the supposed variable proton stoichiometry of the chloroplast cytochrome Ibf complex. Biochimica et Biophysica Acta, 932, 33–42.CrossRefGoogle Scholar
Richards, A.R. (2008). Genetic opportunities to improve cereal root systems for drylands agriculture. Plant Production Science, 1, 12–16.CrossRefGoogle Scholar
Richards, R.A. (2000). Selectable traits to increase crop photosynthesis and yield of grain crops. Journal Experimental Botany, 51, 447–58.CrossRefGoogle ScholarPubMed
Richardson, A.D., Hollinger, D.Y., Burba, G.G., et al. (2006). A multi-site analysis of random error in tower-based measurements of carbon and energy fluxes, Agricultural and Forest Meteorology, 136, 1–18.CrossRefGoogle Scholar
Richter, A. and Popp, M. (1992). The physiological importance of accumulation of cyclitols in Viscum album L. New Phytologist, 121, 431–438.CrossRefGoogle Scholar
Richter, A., Wanek, W., Werner, R.A., et al. (2009). Preparation of starch and soluble sugars of plant material for analysis of carbon isotope composition: a comparison of methods. Rapid Communications in Mass Spectrometry, 23, 2476–2488.CrossRefGoogle ScholarPubMed
Ricklefs, R.E. (2004). A comprehensive framework for global patterns in biodiversity. Ecology Letters, 7, 1–15.CrossRefGoogle Scholar
Rintamaki, E., Keys, A.J. and Parry, M.A.J. (1988). Comparison of the specific activity of ribulose-1,5-bis-phosphate carboxylase-oxygenase from some C3 and C4 plants. Physiologia Plantarum, 74, 326–331.CrossRefGoogle Scholar
Rintamaki, E., Salonen, M., Suoranta, U.M., et al. (1997). Phosphorylation of light-harvesting complex II and photosystem II core proteins shows different irradiance-dependent regulation in vivo application of phosphothreonine antibodies to analysis of thylakoid phosphoproteins. Journal of Biological Chemistry, 272, 30476–30482.CrossRefGoogle ScholarPubMed
Rios-Estepa, R. and Lange, B.M. (2007). Experimental and mathematical approaches to modeling plant metabolic networks. Phytochemistry, 68, 2351–2374.CrossRefGoogle ScholarPubMed
Ripley, B.S., Gilbert, M.E., Ibrahim, D.G., et al. (2007). Drought constraints on C4 photosynthesis: stomatal and metabolic limitations in C3 and C4 subspecies of Alloteropsis semialata. Journal of Experimental Botany, 58, 1351–1363.CrossRefGoogle ScholarPubMed
Rivero, R.M., Kojima, M., Gepstein, A., et al. (2007). Delayed leaf senescence induces extreme drought tolerance in a flowering plant. Proceedings of the National Academy of Sciences, 104, 49, 19631–19636.CrossRefGoogle Scholar
Rivero, R.M., Shulaev, V. and Blumwald, E. (2009). Cytokinin-dependent photorespiration and the protection of photosynthesis during water deficit. Plant Physiology, 150, 1530–1540.CrossRefGoogle ScholarPubMed
Robakowski, P. (2005). Susceptibility to low-temperature photoinhibition in three conifers differing in successional status. Tree Physiology, 25, 1151–1160.CrossRefGoogle ScholarPubMed
Robert, C., Bancal, M.O., Ney, B., et al. (2005). Wheat leaf photosynthesis loss due to leaf rust, with respect to lesion development and leaf nitrogen status. New Phytologist, 165, 227–241.CrossRefGoogle ScholarPubMed
Roberts, A., Borland, A.M. and Griffiths, H. (1997). Discrimination processes and shifts in carboxylation during the phases of crassulacean acid metabolism. Plant Physiology, 113, 1283–1292.CrossRefGoogle ScholarPubMed
Roberts, D.A. and Corbett, M.K. (1965). Reduced photosynthesis in tobacco plants infected with tobacco ringspot virus. Phytopathology, 55, 370–371.Google Scholar
Roberts, P.L. and Wood, K.R. (1982). Effects of a severe (P6) and a mild (W) strain of cucumber mosaic virus on tobacco leaf chlorophyll, starch and cell ultrastructure. Physiological and Molecular Plant Pathology, 21, 31–37.CrossRefGoogle Scholar
Robinson, J.M. (1988). Does O2 photoreduction occur within chloroplasts in vivo?Physiologia Plantarum, 72, 666–680.CrossRefGoogle Scholar
Robinson, S.A. and Osmond, C.B. (1994). Internal gradients of chlorophyll and carotenoid pigments in relation to photoprotection in thick leaves of plants with Crassulacean acid metabolism. Australian Journal of Plant Physiology, 21, 497–506.CrossRefGoogle Scholar
Robinson, S.P. and Portis, A.R., Jr. (1989). Ribulose-1, 5-bisphosphate carboxylase/oxygenase activase protein prevents the in vitro decline in activity of ribulose- 1,5-bisphosphate carboxylase/oxygenase. Plant Physiology, 90, 968–971.CrossRefGoogle ScholarPubMed
Rochaix, J.D. (2007). Role of thylakoid protein kinases in photosynthetic acclimation. FEBS Letters, 581, 2768–2775.CrossRefGoogle ScholarPubMed
Rochette, P., Desjardins, R.L., Pattey, E., et al. (1996). Instantaneous measurement of radiation and water use efficiencies of a maize crop. Agronomy Journal, 88, 627–635.CrossRefGoogle Scholar
Rockström, J., Lannerstad, M. and Falkenmark, M. (2007). Assessing the water challenge of a new green revolution in developing countries. Proceedings of the National Academy of Sciences USA, 104, 6253–6260.CrossRefGoogle ScholarPubMed
Rodeghiero, M., Niinemets, Ü. and Cescatti, A. (2007). Major diffusion leaks of clamp-on leaf cuvettes still unaccounted: how erroneous are the estimates of Farquhar et al. model parameters?Plant Cell Environment, 30, 1006–1022.CrossRefGoogle ScholarPubMed
Roden, J.S. and Ehleringer, J.R. (1999). Hydrogen and oxygen isotope ratio of tree-ring cellulose for riparian trees grown long term under hydroponically controlled environments. Oecologia, 121, 467–477.CrossRefGoogle ScholarPubMed
Roderick, M.L. (1999). Estimating the diffuse component from daily and monthly measurements of global radiation. Agricultural and Forest Meteorology, 95, 169–185.CrossRefGoogle Scholar
Roderick, M.L., Farquhar, G.D., Berry, S.L., et al. (2001). On the direct effect of clouds and atmospheric particles on the productivity and structure of vegetation. Oecologia, 129, 21–30.CrossRefGoogle ScholarPubMed
Rodrigues, M.L., Santos, T.P., Rodrigues, A.P., et al. (2008). Hydraulic and chemical signalling in the regulation of stomatal conductance and plant water use of field grapevines growing under deficit irrigation. Functional Plant Biology, 35 (7) 565–579.CrossRefGoogle Scholar
Rodriguez, D., Ewert, F., Goudriaan, J., et al. (2001). Modelling the response of wheat canopy assimilation to atmospheric CO2 concentrations. New Phytologist, 150, 337–346.CrossRefGoogle Scholar
Rodriguez-Moreno, L., Pineda, M., Soukupova, J., et al. (2008). Early detection of bean infection by Pseudomonas syringae in asymptomatic leaf areas using chlorophyll fluorescence imaging. Photosynthesis Research, 96, 27–35.CrossRefGoogle ScholarPubMed
Rodríguez-Pérez, J.R., Riaño, D., Carlisle, E., et al. (2007). Evaluation of hyperspectral reflectance indexes to detect grapevine water status in vineyards. American Journal of Enology and Viticulture, 58, 302–317.Google Scholar
Rogers, A., Ellsworth, D.S. and Humphries, S.W. (2001). Possible explanation of the disparity between the in vitro and in vivo measurements of Rubisco activity: a study in loblolly pine grown in elevated pCO2. Journal of Experimental Botany, 52, 1555–1561.CrossRefGoogle ScholarPubMed
Rogiers, N., Eugster, W., Furger, M., et al. (2005). Effect of land management on ecosystem carbon fluxes at a subalpine grassland site in the Swiss Alps. Theoretical and Applied Climatology, 80, 187–203.CrossRefGoogle Scholar
Rolfe, S.A. and Scholes, J.D. (2002). Extended depth-of-focus imaging of chlorophyll fluorescence from intact leaves. Photosynthesis Research, 72, 107–115.CrossRefGoogle ScholarPubMed
Rolletschek, H., Radchuk, R., Klukas, C., et al. (2005). Evidence of a key role for photosynthetic oxygen release in oil storage in developing soybean seeds. The New Phytologist, 167, 777–786.CrossRefGoogle ScholarPubMed
Rolletschek, H., Weber, H. and Borisjuk, L. (2003). Energy status and its control on embryogenesis of legumes. Embryo photosynthesis contributes to oxygen supply and is coupled to biosynthetic fluxes. Plant Physiology, 132, 1196–1206.CrossRefGoogle ScholarPubMed
Rolletschek, H., Weschke, W., Wobus, U., et al. (2004). Energy state and its control on seed development: starch accumulation is associated with high ATP and steep oxygen gradients within barley grains. Journal of Experimental Botany, 55, 1351–1359.CrossRefGoogle ScholarPubMed
Roloff, I., Scherm, H. and van Iersel, M.W. (2004). Photosynthesis of blueberry leaves as affected by Septoria leaf spot and abiotic leaf damage. Plant Disease, 88, 397–401.CrossRefGoogle Scholar
Romero-Puertas, M.C., Palma, J.M., Gómez, M., et al. (2002). Cadmium causes the oxidative modification of proteins in pea plants. Plant, Cell and Environment, 25, 677–686.CrossRefGoogle Scholar
Romero-Puertas, M.C., Rodríguez-Serrano, M., Corpas, F.J., et al. (2004). Cadmium-induced subcellular accumulation of O2.- and H2O2 in pea leaves. Plant, Cell and Environment, 27, 1122–1134.CrossRefGoogle Scholar
Roscher, A., Emsley, L., Raymond, P., et al. (1998). Unidirectional steady state rates of central metabolism enzymes measured simultaneously in a living plant tissue. The Journal of Biological Chemistry, 273, 25053–25061.CrossRefGoogle Scholar
Rosema, A. and Verhoef, W., (1991). Modeling of fluorescence light-canopy interaction. Proceeding of the 5th International Colloquium-Physical Signatures in Remote Sensing Courchevel France 14–18 January 1991 (ESA SP-319 May 1991).
Rosema, A., Snel, J.F.H., Zahn, H., et al. (1998). The relation between laser induced chlorophyll fluorescence and photosynthesis. Remote Sensing of the Environmentment, 65, 143–154.CrossRefGoogle Scholar
Rosema, A., Verhoef, W., Schroote, J., et al. (1991). Simulating fluorescence light-canopy interaction in support of laser-induced fluorescence measurements. Remote Sensing of Environment, 37, 117–130.CrossRefGoogle Scholar
Rosenberg, N.J., Blad, B.L. and Verma, S.B. (1983). Microclimate: The Biological Environment. Wiley, New York, USA.Google Scholar
Rosenstiel, T.N., Ebbets, A.L., Khatri, W.C., et al. (2004). Induction of poplar leaf nitrate reductase: a test of extrachloroplastic control of isoprene emission ratePlant Biology, 6, 12–21.Google ScholarPubMed
Rosenthal, S.I. and Camm, E.L. (1996). Effects of air temperature, photoperiod and leaf age on foliar senescence of western larch (Larix occidentalis Nutt.) in environmentally controlled chambers. Plant Cell Environment, 19, 1057–1065.CrossRefGoogle Scholar
Rosenthal, Y., Farnsworth, B., Rodrigo Romo, F.V., et al. (1999). High quality, continuous measurements of CO2 in Biosphere 2 to assess whole mesocosm carbon cycling. Ecological Engineering, 13, 249–262.CrossRefGoogle Scholar
Rosenzweig, C., Tubiello, F.N., Goldberg, R., et al. (2002). Increased crop damage in the US from excess precipitation under climate change. Global Environmental Change, 12, 197–202.CrossRefGoogle Scholar
Ross, J. (1981). The radiation regime and architecture of plant stands. Dr. W. Junk Publishers, The Hague.CrossRefGoogle Scholar
Ross, J. and Sulev, M. (2000). Sources of errors in measurements of PAR. Agricultural and Forest Meteorology, 100, 103.CrossRefGoogle Scholar
Rossa, B. and von Willert, D.J. (1999). Physiological characteristics of geophytes in a semi-arid Namaqualand, South Africa. Plant Ecology, 142, 121–132.CrossRefGoogle Scholar
Rossel, J.B., Wilson, P.B., Hussain, D., et al. (2007). Systemic and intracellular responses to photooxidative stress in Arabidopsis. Plant Cell, 19, 4091–4110.CrossRefGoogle ScholarPubMed
Rosso, D., Ivanov, A.G., Fu, A., Geisler-Lee, J., et al. (2006). IMMUTANS does not act as a stress-induced safety valve in the protection of the photosynthetic apparatus of Arabidopsis during steady state photosynthesis. Plant Physiology, 142, 574–585.CrossRefGoogle Scholar
Rotenberg, D., MacGuidwin, A.E., Saeed, I.A.M., et al. (2004). Interaction of spatially separated Pratylenchus penetrans and Verticillium dahliae on potato measured by impaired photosynthesis. Plant Pathology, 53, 294–302.CrossRefGoogle Scholar
Rotenberg, E. and Yakir, D. (2010). Contribution of semi-arid forests to the climate system. Science, 327, 451–454.CrossRefGoogle ScholarPubMed
Roth, I. (1992). Leaf Structure: Coastal Vegetation and Mangroves of Venezuela. Gebrüder Borntraeger, Berlin, Germany.Google Scholar
Roth-Nebelsick, A., Hassiotou, F. and Veneklaas, E.J. (2009). Stomatal Crypts have small effects on transpiration: A numerical model Analysis. Plant Physiology, 151, 2018–2027.CrossRefGoogle ScholarPubMed
Roth-Nebelsick, A., Uhl, D., Mosbrugger, V., et al. (2001). Evolution and function of leaf venation: a review. Annals of Botany, 87, 553–566.CrossRefGoogle Scholar
Roumet, C., Garnier, E., Suzor, H., et al. (2000). Short and long-term responses of whole-plant gas exchange to elevated CO2 in four herbaceous species. Environmental Experimental Botany, 43, 155–169.CrossRefGoogle Scholar
Rousseaux, M.C., Scopel, A.L., Searles, P.S., et al. (2001). Responses to solar ultraviolet-B radiation in a shrub-dominated natural ecosystem of Tierra del Fuego (southern Argentina). Global Change Biology, 7, 467–478.CrossRefGoogle Scholar
Roussel, M., Dreyer, E., Montpied, P., et al. (2009). The diversity of 13C isotope discrimination in a Quercus robur full-sib family is associated with differences in intrinsic water use efficiency, transpiration efficiency, and stomatal conductance. Journal of Experimental Botany, 60, 2419–2431.CrossRefGoogle Scholar
Rowell, D.M., Ades, P.K., Tausz, M., et al. (2009). Lack of genetic variation in tree ring δ13C suggests a uniform, stomatally driven response to drought stress across Pinus radiata genotypes. Tree Physiology, 29, 191–198.CrossRefGoogle ScholarPubMed
Rowland, D., Dorner, J., Sorensen, R., et al. (2005). Tomato spotted wilt virus in peanut tissue types and physiological effects related to disease incidence and severity. Plant Pathology, 54, 431–440.CrossRefGoogle Scholar
Roxas, V.P., Lodhi, S.A., Garrett, D.K., et al. (2000). Stress tolerance in transgenic tobacco seedlings that overexpress glutathione S-transferase/glutathione peroxidase. Plant and Cell Physiology, 41, 1229–1234.CrossRefGoogle ScholarPubMed
Royer, D.L., Osborne, C.P. and Beerling, D.J. (2005). Contrasting seasonal patterns of carbon gain in evergreen and deciduous trees of ancient polar forests. Paleobiology, 31, 141–150.2.0.CO;2>CrossRefGoogle Scholar
Royer, D.L., Sack, L., Wilf, P., et al. (2007). Fossil leaf economics quantified: calibration, Eocene case study, and implications. Paleobiology, 33, 574–589.CrossRefGoogle Scholar
Royer, D.L., Wing, S.L., Beerling, D.J., et al. (2001). Paleobotanical evidence for near present-day levels of atmospheric CO2 during part of the Tertiary. Science, 292, 2310–2313.CrossRefGoogle ScholarPubMed
Royo, C. (2002). Spanish research priorities and current approaches in plant physiology and breeding of crops to improve water economies. In: Identifying Priority Tools for Cooperation. INCO-MED Workshops (ed. Rodríguez, R.) European Comission-CSIC, Brussels (Belgium).
Ruban, A.V., Berera, R., Ilioana, C., et al. (2007). Identification of photoprotective energy dissipation in higher plants. Nature, 450, 575–579.CrossRefGoogle ScholarPubMed
Ruban, A.V., Pascal, A.A., Robert, B., et al. (2002). Activation of zeaxanthin is an obligatory event in the regulation of photosynthetic light harvesting. Journal of Biological Chemistry, 277, 7785–7789.CrossRefGoogle ScholarPubMed
Ruban, A.V., Young, A.J. and Horton, P. (1993). Induction of nonphotochemical energy dissipation and absorbance changes in leaves. Plant Physiology, 102, 741–750.CrossRefGoogle ScholarPubMed
Rubio, S., Rodrigues, A., Saez, A., et al. (2009). Triple loss of function of protein phosphatases TYPE 2C leads to partial constitutive response to endogenous abscisic acid. Plant Physiology, 150, 1345–1355.CrossRefGoogle ScholarPubMed
Rudall, P.J., Sokoloff, D.D., Remizowa, M.V., et al. (2007). Morphology of Hydatellaceae, an anomalous aquatic family recently recognized as an early divergent angiosperm lineage. American Journal of Botany, 94, 1073–1092.CrossRefGoogle ScholarPubMed
Rüdiger, W., Böhm, S., Helfrich, M., et al. (2005). Enzymes of the last steps of chlorophyll biosynthesis: modification of the substrate structure helps to understand the topology of the active centers. Biochemistry, 44, 10864–10872.CrossRefGoogle ScholarPubMed
Ruelland, E., Vaulthier, M.N., Zachowski, A., et al. (2009). Cold signalling and cold acclimation in plants. Advances in Botanical Research, 49, 35–150.CrossRefGoogle Scholar
Ruess, B.R., Ferrari, S. and Eller, B.M. (1988). Water economy and photosynthesis of the CAM plant Senecio medley woodii during increasing drought. Plant, Cell and Environment, 11, 583–589.CrossRefGoogle Scholar
Rumeau, D., Peltier, G. and Cournac, L. (2007). Chlororespiration and cyclic electron flow around PSI during photosynthesis and plant stress response. Plant Cell and Environment, 30, 1041–1051.CrossRefGoogle ScholarPubMed
Rundel, P.W. and Gibson, A.C. (1996). Ecological Communities and Processes in a Mojave Desert Ecosystem. Cambridge University Press, Cambridge, UK.Google Scholar
Rundel, P.W. and Sharifi, M.R. (1993). Carbon isotope discrimination and resource availability in the desert shrub Larrea tridentata. In: Stable Isotopes and Plant Carbon-Water Relations (eds Ehleringer, J.R., Hall, A.E. and Farquhar, J.D.), Academic Press, San Diego, USA, pp. 173–185.Google Scholar
Rundel, P.W., Cowling, R.M., Esler, K.J., et al. (1995). Winter growth phenology and leaf orientation in Pachypodium namaquanum (Apocynaceae) in the Succulent Karoo of the Richtersveld, South Africa. Oecologia, 101, 472–477.CrossRefGoogle ScholarPubMed
Rundel, P.W., Ehleringer, J.R. and Nagy, K.A. (1988). Stable Isotopes in Ecological Research. Berlin: Springer-Verlag.Google Scholar
Rundel, P.W., Esler, K.J. and Cowling, R.M. (1999). Ecological and phylogenetic patterns of carbon isotope discrimination in the winter-rainfall flora of the Richtersveld, South Africa. Plant Ecology, 142, 133–148.CrossRefGoogle Scholar
Rundel, P.W., Gibson, A.C. and Sharifi, M.R. (2005). Plant functional groups in alpine fellfield habitats of the White Mountains, California. Arctic, Antarctic, and Alpine Research, 37, 358–365.CrossRefGoogle Scholar
Rundel, P.W., Gibson, A.C., Midgley, G.S., et al. (2003). Ecological and ecophysiological patterns in a pre-altiplano shrubland of the Andean Cordillera in northern Chile. Plant Ecology, 169, 179–193.CrossRefGoogle Scholar
Running, S.R. and Nemani, R. (1988). Relating seasonal patterns of the AVHRR vegetation index to simulated photosynthesis and transpiration of forests in different climates. Remote Sensing of Environment, 24, 347–367.CrossRefGoogle Scholar
Running, S.R., Nemani, R.R., Heinsch, F.A., et al. (2004). A continuous satellite-derived measure of global terrestrial primary production. BioScience, 54, 547–560.CrossRefGoogle Scholar
Running, S.W., Baldocchi, D., Turner, D., et al. (1999). A global terrestrial monitoring network integrating tower fluxes, flask sampling, ecosystem modeling and EOS satellite data. Remote Sensing of Environment, 70, 108–127.CrossRefGoogle Scholar
Runyon, J., Waring, R.H., Goward, S.N., et al. (1994). Environmental limits on net primary production and light-use efficiency across the Oregon transect. Ecological Applications, 4, 226–237.CrossRefGoogle Scholar
Russell, G., Jarvis, P.G. and Monteith, J.L. (1989). Absorption of radiation by canopies and stand growth. In: Plant Canopies, Their Growth, Form and Function (eds Russell, G., Marshall, B. and Jarvis, P.G.), Cambridge University Press, Cambridge, UK, pp. 21–39.CrossRefGoogle Scholar
Russell, R.J. (1931). Dry climates of the Unites States: I climatic map. University of California, Publications in Geography, 5, 1–41.Google Scholar
Russo, M. and Martelli, G.P. (1982). Ultrastructure of turnip crinckle- and Saguaro cactus virus-infected tissues. Virology, 118, 109–116.CrossRefGoogle Scholar
Rutherford, A.W. and Boussac, A. (2004). Water photolysis in biology. Science, 303, 1782–1784.CrossRefGoogle ScholarPubMed
Rutherford, A.W., Crofts, A.R. and Inoue, Y. (1982). Thermoluminescence as a probe of Photosystem II photochemistry: the origin of the flash-induced glow peaks. Biochimica Biophysica Acta, 689, 457–465.CrossRefGoogle Scholar
Rutherford, A.W., Govindjee, J. and Inoue, Y. (1984). Charge accumulation and photochemistry in leaves studied by thermoluminescence and delayed light emission. Proceedings of the National Academy of Science USA, 81, 1107–1111.CrossRefGoogle ScholarPubMed
Ruuska, S.A., Schwender, J. and Ohlrogge, J.B. (2004). The capacity of green oilseeds to utilize photosynthesis to drive biosynthetic processes. Plant Physiology, 136, 2700–2709.CrossRefGoogle ScholarPubMed
Ryan, M., Bond, B.J., Law, B.E., et al. (2000). Transpiration and whole-tree conductance in ponderosa pine trees of different heights. Oecologia, 124, 553–560.CrossRefGoogle ScholarPubMed
Ryan, M.G. and Waring, R.H. (1992). Maintenance respiration and stand development in a subalpine lodgepole pine forest. Ecology, 73, 2100–2108.CrossRefGoogle Scholar
Ryan, M.G. and Yoder, B.J. (1997). Hydraulic limits to tree height and tree growth. What keeps trees from growing beyond a certain height?BioScience, 47, 235–242.CrossRefGoogle Scholar
Ryan, M.G., Binkley, D. and Fownes, J.H. (1997). Age-related decline in forest productivity: pattern and process. Advances in Ecological Research, 27, 213–262.CrossRefGoogle Scholar
Ryan, M.G., Binkley, D., Fownes, J.H., et al. (2004). An experimental test of the causes of forest growth decline with stand age. Ecological Monographs, 74, 393–414.CrossRefGoogle Scholar
Ryel, R.J. and Beyschlag, W. (1995). Benefits associated with steep foliage orientation in two tussock grasses of the American Intermountain West. A look at water-use-efficiency and photoinhibition. Flora, 190, 1–10.CrossRefGoogle Scholar
Ryel, R.J., Beyschlag, W. and Caldwell, M.M. (1994). Light field heterogeneity among tussock grasses – theoretical considerations of light harvesting and seedling establishment in tussocks and uniform tiller distributions. Oecologia, 98, 241–246.CrossRefGoogle ScholarPubMed
Sabar, M., De Paepe, R. and Kouchkovsky, Y. (2000). Complex I impairment, respiratory compensations, and photosynthetic decrease in nuclear and mitochondrial male sterile mutants of Nicotiana sylvestris. Plant Physiology, 124, 1239–1249.CrossRefGoogle ScholarPubMed
Sack, L. and Holbrook, N.M. (2006). Leaf hydraulics. Annual Review of Plant Biology, 57, 361–381.CrossRefGoogle ScholarPubMed
Sack, L., Cowan, P.D., Jaikumar, N.J., et al. (2003). The ‘hydrology’ of leaves: coordination of structure and function in temperate woody species. Plant, Cell and Environment, 26, 1343–1356.CrossRefGoogle Scholar
Sack, L., Dietrich, E.M., Streeter, C.M., et al. (2008). Leaf palmate venation and vascular redundancy confer tolerance of hydraulic disruption. Proceedings of the National Academy of Sciences, 105, 1567–1572.CrossRefGoogle ScholarPubMed
Sacksteder, C.A. and Kramer, D.A. (2000). Dark-interval relaxation kinetics (DIRK) of absorbance changes as a quantitative probe of steady state electron transfer. Photosynthesis Research, 66, 145–158.CrossRefGoogle ScholarPubMed
Sade, N., Gebretsadik, M., Seligmann, R., et al. (2010). The role of tobacco aquaporin1 in improving water use efficiency, hydraulic conductivity, and yield production under salt stress. Plant Physiology, 152, 245–254.CrossRefGoogle ScholarPubMed
Saeed, I.A.M., MacGuidwin, A.E., Rouse, D.I., et al. (1999). Limitation to photosynthesis in Pratylenchus penetrans- and Verticillium dahliae-infected potato. Crop Science, 39, 1340–1346.CrossRefGoogle Scholar
Sáez, A., Apostolova, N., Gonzalez-Guzman, M., et al. (2004). Gain-of-function and loss-of-function phenotypes of the protein phosphatase 2C HAB1 reveal its role as a negative regulator of abscisic acid signalling. The Plant Journal, 37, 354–369.CrossRefGoogle ScholarPubMed
Sáez, A., Robert, N., Maktabi, M.H., et al. (2006). Enhancement of abscisic acid sensitivity and reduction of water consumption in Arabidopsis by combined inactivation of the protein phosphatases type 2C ABI1 and HAB1. Plant Physiology, 141, 1389–1399.CrossRefGoogle ScholarPubMed
Sagardoy, R., Morales, F., López-Millán, A.F., et al. (2009). Effects of zinc toxicity in sugar beet (Beta vulgaris L.) plants grown in hydroponics. Plant Biology, 11, 339–350.CrossRefGoogle ScholarPubMed
Sagardoy, R., Vázquez, S., Florez-Sarasa, I.D. et al. (2010). Stomatal and mesophyll conductances to CO2 are the main limitations to photosynthesis in sugar beet (Beta vulgaris) plants grown with excess zinc. The New Phytologist, 187, 145–158.CrossRefGoogle ScholarPubMed
Sage, R.F. (1999) Why C4 photosynthesis?. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, CA, USA, pp. 3–16.Google Scholar
Sage, R.F. (2001). Environmental and evolutionary preconditions for the origin and diversification of the C4 phtosynthetic syndrome. Plant Biology, 3, 202–213.CrossRefGoogle Scholar
Sage, R.F. (2004). The evolution of C4 photosynthesis. New Phytologist, 161, 341–370.CrossRefGoogle Scholar
Sage, R.F. and Kubien, D.S. (2003). Quo vadis C4? An ecophysiological perspective on global change and the future of C4 plants. Photosynthesis Research, 77, 209–225.CrossRefGoogle Scholar
Sage, R.F. and Kubien, D.S. (2007). The temperature response of C3 and C4 photosynthesis. Plant, Cell and Environment, 30, 1086–1106.CrossRefGoogle Scholar
Sage, R.F. and Monson, R.K. (eds.). (1999). C4 Plant Biology. Academic Press, San Diego, USA.Google Scholar
Sage, R.F. and Pearcy, R.W. (1987a). The nitrogen use efficiency of C3 and C4 plants. I. Leaf nitrogen effects on the gas exchange characteristics of Chenopodium album L. and Amaranthus retroflexus L. Plant Physiology, 84, 954–958.CrossRefGoogle Scholar
Sage, R.F. and Pearcy, R.W. (1987b). The nitrogen use efficiency of C3 and C4 plants. II. Leaf nitrogen effects on the gas exchange characteristics of Chenopodium album L. and Amaranthus retroflexus L. Plant Physiology, 84, 959–963.CrossRefGoogle Scholar
Sage, R.F., Cen, Y.P. and Li, M. (2002). The activation state of Rubisco directly limits photosynthesis at low CO2 and low O2 partial pressures. Photosynthesis Research, 71, 241–250.CrossRefGoogle Scholar
Sage, R.F., Li, M. and Monson, R.K. (1999). The taxonomic distribution of C4 photosynthesis. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, USA, pp. 551–584.Google Scholar
Sage, R.F. and Sharkey, T.D. (1987).The effect of temperature on the occurrence of O2 and CO2 insensitive photosynthesis in field grown plants. Plant Physiology, 84, 658–664.CrossRefGoogle Scholar
Sage, R.F., Schäppi, B. and Körner, C. (1997). Effect of atmospheric CO2 enrichment on Rubisco content in herbaceous species from high and low altitude. Acta Ecologica, 18, 183–192.CrossRefGoogle Scholar
Sage, R.F., Sharkey, T.D. and Seeman, J.R. (1989). Acclimation of photosynthesis to elevated carbon dioxide in five C3 species. Plant Physiology, 89, 590–596.CrossRefGoogle Scholar
Sage, R.F., Way, D.A. and Kubien, D.S. (2008) Rubisco, Rubisco activase, and global climate change. Journal of Experimental Botany, 59, 1581–1595.CrossRefGoogle ScholarPubMed
Sage, R.F., Wedin, D.A. and Li, M. (1999). The biogeography of C4 photosynthesis, patterns and controlling factors. In: C4 Plant Biology (eds Sage, R.F. and Monson, R.K.), Academic Press, San Diego, USA, pp. 313–373.Google Scholar
Saito, T., Soga, K., Hoson, T. et al. (2006). The bulk elastic modulus and the reversible properties of cell walls in developing Quercus leaves. Plant and Cell Physiology, 47, 715–725.CrossRefGoogle ScholarPubMed
Saito, Y., Saito, R., Nomura, E., et al. (1999). Performance check of vegetation fluorescence imaging lidar through in vivo and remote estimation of chlorophyll concentration inside plant leaves. Optical Review, 6, 155–159.CrossRefGoogle Scholar
Sajnani, C., Zurita, J., Roncel, M., et al. (2007). Changes in photosynthetic metabolism induced by tobamovirus infection in Nicotiana benthamiana studied in vivo by chlorophyll thermoluminescence. New Phytologist, 175, 120–130.CrossRefGoogle Scholar
Sakai, A. (1960). Survival of the twigs of woody plants at –196°C. Nature, 185, 393–394.CrossRefGoogle Scholar
Sakai, A. and Larcher, W. (1987). Frost survival of plants. responses and adaptation to freezing stress. Ecological Studies, 62.CrossRefGoogle Scholar
Sakata, T. and Yokoi, Y. (2002). Analysis of the O2 dependency in leaf-level photosynthesisof two Reynoutria japonica populations growing at different altitudes. Plant, Cell and Environment, 25, 65–74.CrossRefGoogle Scholar
Sakata, T., Nakano, T., Iino, T., et al. (2006). Contrastive seasonal changes in ecophysiological traits of leaves of two perennial Polygonaceae herb species differing in leaf longevity and altitudinal distribution. Ecological Research, 21, 633–640.CrossRefGoogle Scholar
Sakugawa, H. and Cape, J.N. (2007). Harmful effects of atmospheric nitrous acid on the physiological status of Scots pine trees. Environmental Pollution, 147, 532–534.CrossRefGoogle ScholarPubMed
Sala, A. and Hoch, G. (2009). Height-related growth declines in ponderosa pine are not due to carbon limitation. Plant Cell Environment, 32, 22–30.CrossRefGoogle Scholar
Sala, A. and Tenhunen, J.D. (1996). Simulations of canopy net photosynthesis and transpiration in Quercus ilex L. under the influence of seasonal drought. Agricultural and Forest Meteorology, 78, 203–222.CrossRefGoogle Scholar
Sala, O.E., Parton, W.J., Joyce, L.A., et al. (1988). Primary production of the central grassland region of the United States. Ecology, 69, 40–45.CrossRefGoogle Scholar
Saleska, S.R., Miller, S.D., Matross, D.M., et al. (2003). Carbon in Amazon forests: unexpected seasonal fluxes and disturbance-induced losses. Science, 302, 1554–1557.CrossRefGoogle ScholarPubMed
Salleo, S. and Lo Gullo, M.A. (1990). Sclerophylly and plant water relations in three Mediterranean Quercus species. Annals of Botany, 65, 259–270.CrossRefGoogle Scholar
Salleo, S. and Nardini, A. (2000). Sclerophylly: evolutionary advantage or mere epiphenomenon?Plant Biosystems, 134, 247–259.CrossRefGoogle Scholar
Salleo, S., Nardini, A. and Lo Gullo, M.A. (1997). Is sclerophylly of Mediterranean evergreens and adaptation to drought?The New Phytologist, 135, 603–612.CrossRefGoogle Scholar
Salvetat, R., Juneau, P. and Popovic, R. (1998). Measurement of chlorophyll fluorescence by synchronous detection in integrating sphere: a modified analytical approach for the accurate determination of photosynthesis parameters for whole plants. Environmental Science and Technology, 32, 2640–2645.CrossRefGoogle Scholar
Salvi, S. and Tuberosa, R. (2005). To clone or not to clone plant QTLs: present and future challenges. Trends in Plant Science, 10, 297–304.CrossRefGoogle ScholarPubMed
Salvucci, M.E. and Crafts-Brandner, S.J. (2004a). Relationship between the heat tolerance of photosynthesis and the thermal stability of rubisco activase in plants from contrasting thermal environments. Plant Physiology, 134, 1460–1470.CrossRefGoogle ScholarPubMed
Salvucci, M.E., Crafts-Brandner, S.J. (2004b). Inhibition of photosynthesis by heat stress: the activation state of Rubisco as a limiting factor in photosynthesis. Physiologia Plantarum, 120, 179–186.CrossRefGoogle ScholarPubMed
Salvucci, M.E., DeRidder, B.P. and Portis, Jr A.R. (2006). Effect of activase level and isoform on the thermotolerance of photosynthesis in Arabidopsis. Journal of Experimental Botany, 57, 3793–3799.CrossRefGoogle ScholarPubMed
Sampol, B., Bota, J., Riera, D., et al. (2003). Analysis of the virus-induced inhibition of photosynthesis in malmsey grapevines. New Phytologist, 160, 403–412.CrossRefGoogle Scholar
Sampson, D.A., Janssens, I.A. and Ceulemans, R. (2001). Simulated soil CO2 efflux and net ecosystem exchange in a 70-year-old Belgian Scots pine stand using the process model SECRETS. Annals of Forest Science, 58, 31–36.CrossRefGoogle Scholar
Samson, G., Tremblay, N., Dudelzak, A.E., et al. (2000). Nutrient stress of corn plants: early detection and discrimination using a compact multiwavelength fluorescent lidar. In: EARSeL Proceedings (ed. Reuter, R.), EARSeL, Dresden, Germany, pp. 214–223.
Samsuddin, Z. and Impens, I. (1979). Photosynthesis and diffusion resistances to carbon dioxide in Hevea brasiliensis muel. agr. clones. Oecologia, 37, 361–363.CrossRefGoogle ScholarPubMed
Sandermann, H. Jr. (1996). Ozone and plant health. Annual Review of Phytopathology, 34, 347–366.CrossRefGoogle ScholarPubMed
Sandmann, G. and Boger, P. (1980). Copper-mediated lipid peroxidation processes in photosynthetic membranes. Plant Physiology, 66, 797–800.CrossRefGoogle ScholarPubMed
Sandquist, D.R. and Ehleringer, J.R. (2003). Population-and family level variation of brittlebush (Encelia farinosa, Asteraceae) pubescence: its relation to drought and implications for selection in variable environments. American Journal of Botany, 90, 1481–1486.CrossRefGoogle ScholarPubMed
Sandquist, D.R., Schuster, W.S.F., Donovan, L.A., et al. (1993). Differences in carbon isotope discrimination between seedlings and adults of southwestern desert perennial plants. The Southwestern Naturalist, 38, 212–217.CrossRefGoogle Scholar
Sands, P.J. (1995). Modelling canopy production. I. Optimal distribution of photosynthetic resources. Australian Journal of Plant Physiology, 22, 593–601.CrossRefGoogle Scholar
Sandström, J. (2000). Nutritional quality of phloem sap in relation to host plant-alternation in the bird cherry oat aphid. Chemoecology, 10, 17–24.Google Scholar
Sane, P.V., Ivanov, A.G., Hurry, V., et al. (2003). Changes in the redox potential of primary and secondary electron-accepting quinones in photosystem II confer increased resistance to photoinhibition in low-temperature-acclimated Arabidopsis. Plant Physiology, 132, 2144–2151.CrossRefGoogle Scholar
Sangsing, K., Poonpipope, K., Thanisawanyangkura, S., et al. (2004) Respiation rate and a two-component model of growth and maintenance respiration in leaves of rubber (Hevea brasiliensis Muell. Arg.). Kasetsart Journal (Natural Science), 38, 320–330.Google Scholar
Sankaran, M., Hanan, N.P., Scholes, R.J., et al. (2005). Determinants of woody cover in African savannas. Nature, 438, 846–849.CrossRefGoogle ScholarPubMed
Santarius, K.A. (1975). Sites of heat sensitivity in chloroplasts and differential inactivation of cyclic and noncyclic photophosphorylation by heating. Journal of Thermal Biology, 1, 101–107.CrossRefGoogle Scholar
Santarius, K.A. and Müller, M. (1979). Investigations on heat resistance of spinach leaves. Planta, 146, 529–538.CrossRefGoogle ScholarPubMed
Santiago, J., Dupeux, F., Round, A., et al. (2009). The abscisic acid receptor PYR1 in complex with abscisic acid. Nature, 462, 665–668.CrossRefGoogle ScholarPubMed
Santiago, L.S. and Wright, S.J. (2007). Leaf functional traits of tropical forest plants in relation to growth form. Functional Ecology, 21, 19–27.CrossRefGoogle Scholar
Sanitago, L.S., Goldstein, G., Meinzer, F.C., et al. (2004). Leaf photosynthetic traits scale with hydraulic conductivity and wood density in Panamanian forest canopy trees. Oecologia, 140, 543–550.CrossRefGoogle Scholar
Šantrůček, J. and Sage, R.F. (1996). Acclimation of stomatal conductance to a CO2 enriched atmosphere and elevated temperature in Chenopodium album. AustralianJournal of Plant Physiology, 23, 467–478.Google Scholar
Šantrůček, J., Šimáňová, E.Karbulková, J., et al. (2004). A new technique for measurement of water permeability of stomatous cuticular membranes isolated from Hedera helix leaves. Journal of Experimental Botany, 55, 1411–1422.CrossRefGoogle ScholarPubMed
Sarris, D., Christodoulakis, D. and Körner, C. (2007). Recent decline in precipitation and tree growth in the eastern Mediterranean. Global Change Biology, 13, 1187–1200.CrossRefGoogle Scholar
Sasaki, H., Samejima, M. and Ishii, R. (1996). Analysis by delta-13C measurement on mechanism of cultivar difference in leaf photosynthesis on rice (Oryza sativa L.). Plant and Cell Physiology, 37, 1161–1166.CrossRefGoogle Scholar
Sassenrath, G.F., Ort, D.R. and Portis, A.R. (1990). Impaired reductive activation of stromal bisphosphatases in tomato leaves following low-temperature exposure at high light. Archives of Biochemistry and Biophysics, 282, 302–308.CrossRefGoogle ScholarPubMed
Sassenrath-Cole, G.F., Pearcy, R.W. and Steinmaus, S. (1994). The role of enzyme activation state in limiting carbon assimilation under variable light conditions. Photosynthesis Research, 41, 295–302.CrossRefGoogle Scholar
Saugier, B., Roy, J. and Mooney, H.A. (2001). Estimation of global terrestrial productivity: converging toward a single number? In: Terrestrial Global Productivity (eds Mooney, H.A., Saugier, B. and Roy, J.), Academic Press, San Diego, USA, pp. 543–557.Google Scholar
Savé, R., Biel, C. and de Herralde, F. (2000). Leaf pubescence, water relations and chlorophyll fluorescence in two subspecies of Lotus creticus L. Biologia Plantarum, 43, 239–244.CrossRefGoogle Scholar
Saveyn, A., Steppe, K., Ubierna, N., et al. (2010). Woody tissue photosynthesis and bud development in young plants. Plant, Cell and Environment, 33, 1949–1958.CrossRefGoogle ScholarPubMed
Savitch, L.V., Allard, G., Seki, M., et al. (2005). The effect of over-expression of two Brassica CBF/DREB1-like transcription factors on photosynthetic capacity and freezing tolerance in Brassica napus. Plant and Cell Physiology, 46, 1525–1539.CrossRefGoogle Scholar
Savitch, L.V., Barker-Astrom, J., Ivanov, A.G., et al. (2001). Cold acclimation of Arabidopsis thaliana results in incomplete recovery of photosynthetic capacity, associated with an increased reduction of the chloroplast stroma. Planta, 214, 295–303.CrossRefGoogle ScholarPubMed
Savitch, L.V., Gray, G.R. and Huner, N.P.A. (1997). Feedback-limited photosynthesis and regulation of sucrose-starch accumulation during cold acclimation and low-temperature stress in a spring and winter wheat. Planta, 201, 18–26.CrossRefGoogle Scholar
Savitch, L.V., Harney, T. and Huner, N.P.A. (2000a). Sucrose metabolism in spring and winter wheat in response to high irradiance, cold stress and cold acclimation. Physiologia Plantarum, 108, 270–278.CrossRefGoogle Scholar
Savitch, L.V., Leonardos, E.D., Krol, M., et al. (2002). Two different strategies for light utilization in photosynthesis in relation to growth and cold acclimation. Plant Cell Environment, 25, 761–771.CrossRefGoogle Scholar
Savitch, L.V., Massacci, A., Gray, G.R., et al. (2000b). Acclimation to low temperature or high light mitigates sensitivity to photoinhibition: roles of the Calvin cycle and the Mehler reaction. Austral Journal of Plant Physiology, 27, 253–264.Google Scholar
Sawada, S., Usuda, H. and Tsukui, T. (1992). Participation of inorganic orthophosphate in regulation of the Ribulose-1,5-bisphosphate carboxylase activity in response to changes in the photosynthetic source-sink balance. Plant and Cell Physiology, 33, 943–949.Google Scholar
Saxe, H. (1986). Effects of NO, NO2 and CO2 on net photosynthesis, dark respiration and transpiration of pot plants. New Phytologist, 103, 185–197.CrossRefGoogle Scholar
Saxe, H. and Christensen, O. V. (1985). Effects of carbon dioxide with and without nitric oxide pollution on growth, morphogenesis and production time of pot plants. Environmental Pollution, 38, 159–169.CrossRefGoogle Scholar
Saxe, H., Cannell, M.G.R., Johnsen, Ã, et al. (2001). Tansley review 123: Tree and forest functioning in response to global warming. New Phytologist, 149, 369–400.CrossRefGoogle Scholar
Saxe, H., Ellsworth, D.S. and Heath, J. (1998). Tree and forest functioning in an enriched CO2 atmosphere. New Phytologist, 139, 395–436.CrossRefGoogle Scholar
Sayed, O.H. (2001). Crassulacean acid metablism 1975–2000, a check list. Photosynthetica, 39, 339–352.CrossRefGoogle Scholar
Scafaro, A.P., von Caemmerer, S., Evans, J.R. et al. (2011). Temperature response of mesophyll conductance in cultivated and wild Oryza species with contrasting mesophyll cell wall thickness. Plant, Cell and Environment, 34, 1999–2008CrossRef
Scanlon, T.M. and Albertson, J.D. (2004). Canopy scale measurements of CO2 and water vapor exchange along a precipitation gradient in southern Africa. Global Change Biology, 10, 329–341.CrossRefGoogle Scholar
Scarascia-Mugnozza, G., De Angelis, P., Matteucci, G., et al. (1996). Long term exposure to elevated CO2 in a natural Quercus ilex L. community: net photosynthesis and photochemical efficiency of PSII at different levels of water stress. Plant, Cell and Environment, 19, 643–654.CrossRefGoogle Scholar
Scartazza, A., Lauteri, M., Guido, M.C., et al. (1998). Carbon isotope discrimination in leaf and stem sugars, water-use efficiency and mesophyll conductance during different developmental stages in rice subjected to drought. Australian Journal of Plant Physiology, 25, 489–498.CrossRefGoogle Scholar
Scartazza, A., Mata, C., Matteucci, G., et al. (2004). Comparisons of δ13C of photosynthetic products and ecosystem respiratory CO2 and their response to seasonal climate variability. Oecologia, 140, 340–351.CrossRefGoogle Scholar
Schaefer, H.M. and Wilkinson, D.M. (2004). Red leaves, insects and coevolution: a red herring?Trends in Ecology and Evolution, 19, 616–618.CrossRefGoogle ScholarPubMed
Schaeffer, S.M., Anderson, D.E., Burns, S.P., et al. (2008). Canopy structure and atmospheric flows in relation to the δ13C of respired CO2 in a subalpine coniferous forest. Agricultural and Forest Meteorology, 148, 592–605.CrossRefGoogle Scholar
Schäfer, K.V.R., Oren, R. and Tenhunen, J.D. (2000). The effect of tree height on crown level stomatal conductance. Plant Cell Environment, 23, 365–375.CrossRefGoogle Scholar
Schans, J. and Arntzen, F.K. (1991). Photosynthesis, transpiration and plant growth characters of different potato cultivars at various densities of Globodera pallida. Netherland Journal of Plant Pathology, 97, 297–310.CrossRefGoogle Scholar
Schansker, G., Tóth, S.Z. and Strasser, R. J. (2005). Methylviologen and dibromothymoquinone treatments of pea leaves reveal the role of photosystem I in the Chl a fluorescence rise OJIP. Biochimica et Biophysica Acta, 1706, 250–261.CrossRefGoogle Scholar
Scharte, J., Schon, H. and Weiss, E. (2005). Photosynthesis and carbohydrate metabolism in tobacco leaves during an incompatible interaction with Phytophthora nicotianae. Plant, Cell and Environment, 28, 1421–1435.CrossRefGoogle Scholar
Schäufele, R., Santrucek, J. and Schynder, H. (2011). Dynamic changes of canopy scale mesophyll conductance to CO2 diffusion of sunflower as affected by CO2 concentration and abscisic acid. Plant, Cell and Environment, 34, 127–136.CrossRefGoogle Scholar
Scheibe, R. (1990). Light/dark modulation: regulation of chloroplast metabolism in a new light. Botanica Acta, 103, 327–334.CrossRefGoogle Scholar
Scheibe, R. (2004). Malate valves to balance cellular energy supply. Physiologia Plantarum, 120, 21–26.CrossRefGoogle ScholarPubMed
Scheibe, R., Backhausen, J. E., Emmerlich, V., et al. (2005). Strategies to maintain redox homeostasis during photosynthesis under changing conditions. Journal of Experimental Botany, 56, 1481–1489.CrossRefGoogle ScholarPubMed
Scheible, W.R., Krapp, A. and Stitt, M. (2000). Reciprocal diurnal changes of PEPc expression, cytosolic pyruvate kinase, citrate synthase and NADP-isocitrate dehydrogenase expression regulate organic acid metabolism during nitrate assimilation in tobacco leaves. Plant, Cell Environment, 23, 1155–1167.CrossRefGoogle Scholar
Scheirs, J., De Bruyn, L. and Verhagen, R. (2001). A test of the C3-C4 hypothesis with two grass miners. Ecology, 82, 410–421.Google Scholar
Scheller, H.V. and Haldrup, A. (2005). Photoinhibition of photosystem I. Planta, 221, 5–8.CrossRefGoogle ScholarPubMed
Schidlowski, M. (1988). A 3,800-million-year isotopic record of life from carbon in sedimentary rocks. Nature, 333, 313–318.CrossRefGoogle Scholar
Schidlowski, M. (1995). Isotope fractionations in the terrestrial carbon cycle: a brief over view. Advances in Space Research, 15, 441–449.CrossRefGoogle Scholar
Schimel, D. (2007). Carbon cycle conundrums. Proceedings of the National Academy of Sciences of the USA, 104, 18353–18354.CrossRefGoogle ScholarPubMed
Schimel, D.S., Braswell, B.H., McKeown, R., et al. (1996). Climate and nitrogen controls on the geography and timescales of terrestrial biogeochemical cycling. Global Biogeochemical Cycles, 10, 677–692.CrossRefGoogle Scholar
Schlegel, H., Godbold, D.L. and Hüttermann, A. (1987). Whole plant aspects of heavy metal induced changes in CO2 uptake and water relations of spruce (Picea abies) seedlings. Physiologia Plantarum, 69, 265–270.CrossRefGoogle Scholar
Schlesinger, W.H., Belnap, J. and Marion, G. (2009). On carbon sequestration in desert ecosystems. Global Change Biolog, 15, 1488–90.CrossRefGoogle Scholar
Schleucher, J., Vanderveer, P., Markley, J.L., et al. (1999). Intramolecular deuterium distributions reveal disequilibrium of chloroplast phosphoglucose isomerase. Plant, Cell and Environment, 22, 525–533.CrossRefGoogle Scholar
Schmid, H.P. (2002). Footprint modeling for vegetation atmosphere exchange studies: a review and perspective. Agricultural and Forest Meteorology, 113, 159–183.CrossRefGoogle Scholar
Schmitgen, S., Ciais, P., Geiss, H., et al. (2004). Carbon dioxide uptake of a forested region in southwest France derived from airborne CO2 and CO observations in a Lagrangian budget approach. Journal of Geophysical Research, 109, D14302.CrossRefGoogle Scholar
Schmitz, A., Tartachnyk, I., Kiewnick, S., et al. (2006). Detection of Heterodera schachtii infestation in sugar beet by means of laser-induced and pulse amplitude modulated chlorophyll fluorescente. Nematology, 8, 273–286.CrossRefGoogle Scholar
Schnablová, R., Synková, H. and Čeřovská, N. (2005). The influence of potato virus Y infection on the ultrastructure of Pssu-ipt transgenic tobaccoInternational Journal of Plant Sciences, 166, 713–721.CrossRefGoogle Scholar
Schneckenburger, H. and Frenz, M. (1986). Time-resolved fluorescence of conifers exposed to environmental pollutants. Radiation Environment Biophysics, 25, 289–295.CrossRefGoogle ScholarPubMed
Schoefs, B. and Franck, F. (2003). Protochlorophyllide reduction: mechanisms and evolution. Photochemistry and Photobiology, 78, 543–557.2.0.CO;2>CrossRefGoogle Scholar
Schoettle, A.W. and Smith, W.K. (1991). Interrelation between shoot characteristics and solar irradiance in the crown of Pinus contorta ssp. latifolia. Tree Physiology, 9, 245–254.CrossRefGoogle ScholarPubMed
Schoettle, A.W. and Smith, W.K. (1999). Interrelationships among light, photosynthesis and nitrogen in the crown of mature Pinus contorta ssp. latifolia. Tree Physiology, 19, 13–22.CrossRefGoogle ScholarPubMed
Scholberg, J., McNeal, B.L., Jones, J.W., et al. (2000). Growth and canopy characteristics of field-grown tomato. Agronomy Journal, 92, 152–159.CrossRefGoogle Scholar
Scholes, J.D. and Rolfe, S.A. (1996). Photosynthesis in localised regions of oat leaves infected with crown rust (Puccinia coronata): quantitative imaging of chlorophyll fluorescence. Planta, 199, 573–582.CrossRefGoogle Scholar
Scholes, J.D., Lee, P.J., Horton, P. et al (1994). Invertase: understanding changes in the photosynthetic and carbohydrate metabolism of barley leaves infected with powdery mildew. New Phytologist, 126, 213–222.CrossRefGoogle Scholar
Scholes, R.J. and Archer, S. (1997) Tree-grass interactions in savannas. Annual Review of Ecology and Systematics, 28, 517–544.CrossRefGoogle Scholar
Scholes, R.J., Frost, P.G. and Tian, Y. (2004). Canopy structure in savannas along a moisture gradient on Kalahari sands. Global Change Biology, 10, 292–302.CrossRefGoogle Scholar
Scholze, M., Ciais, P. and Heimann, M. (2008). Modeling terrestrial 13C cycling: climate, land use and fire. Global Biogeochemical Cycles, 22, GB1009.CrossRefGoogle Scholar
Schomburg, M. and Kluge, M. (1994). Phenotypic adaptation to elevated temperatures of tonoplast fluidity in the CAM plant Kalanchoë daigremontiana is caused by membrane proteins. Botanica Acta, 107, 328–332.CrossRefGoogle Scholar
Schöner, S. and Krause, G.H. (1990). Protective systems against active oxygen species in spinach: response to cold acclimation in excess light. Planta, 180, 383–389.CrossRefGoogle ScholarPubMed
Schrader, S.M., Wise, R.R., Wacholtz, W.F., et al. (2004). Thylakoid membrane responses to moderately high leaf temperature in pima cotton. Plant, Cell and Environment, 27, 725–735.CrossRefGoogle Scholar
Schrauder, M., Langebartels, C. and Sandermann, H. (1997). Changes in the biochemical status of plant cells induced by the environmental pollutant ozone. Physiologia Plantarum, 100, 274–280.CrossRefGoogle Scholar
Schreiber, U. and Berry, J.A. (1977). Heat-induced changes in chlorophyll fluorescence in intact leaves correlated with damage of the photosynthetic apparatus. Planta, 136, 233–238.CrossRefGoogle ScholarPubMed
Schreiber, U. and Neubauer, C. (1987). The polyphasic rise of chlorophyll upon onset of continuous illumination. II. Partial control by the photosystem II donor side and possible ways of interpretation. Zeitschrift für Naturforschung, 42c, 1255–1264.Google Scholar
Schreiber, U., Bilger, W. and Neubauer, C. (1994). Chlorophyll fluorescence as a noninvasive indicator for rapid assessment of in vivo photosynthesis. In: Ecophysiology of Photosynthesis (eds Schulze, E.D. and Caldwell, M.M.), Springer-Verlag, Berlin, Germany, pp. 49–70.
Schreiber, U., Klughammer, C. and Neubauer, C. (1989). Measuring P700 absorbance changes around 830 nm with a new type of pulse modulation system. Zeitschrift für Naturforschung, 43c, 686–698.Google Scholar
Schreiber, U., Neubauer, C. and Schliwa, U. (1993). PAM fluorometer based on medium-frequency pulsed Xe-flash measuring light: a highly sensitive new tool in basic and applied photosynthesis research. Photosynthesis Research, 36, 65–72.CrossRefGoogle ScholarPubMed
Schreiber, U., Schliwa, U. and Bilger, W. (1986). Continuous recording of photochemical and non-photochemical fluorescence quenching with a new type of pulse modulation fluorometer. Photosynthesis Research, 10, 51–62.CrossRefGoogle ScholarPubMed
Schröder, R., Forstreuter, M. and Hilker, M. (2005). A plant notices insect egg deposition and changes its rate of photosynthesis. Plant Physiology, 138, 470–477.CrossRefGoogle ScholarPubMed
Schroeder, J.I. (2003). Knockout of the guard cell K+ out channel and stomatal movements. Proceedings of the National Academy of Sciences USA, 100, 4976–4977.CrossRefGoogle ScholarPubMed
Schroeder, J.I., Allen, G.J., Hugouvieux, V., et al. (2001a). Guard cell signal transduction. Annual Review of Plant Physiology and Plant Molecular Biology, 52, 627–658.CrossRefGoogle ScholarPubMed
Schroeder, J.I., Kwak, J.M. and Allen, G.J. (2001b). Guard cell abscisic acid signalling and engineering drought hardiness in plants. Nature, 410, 327–330.CrossRefGoogle ScholarPubMed
Schuerger, A.C., Capelle, G.A., Di Benedetto, J.A., et al. (2003). Comparison of two hyperspectral imaging and two laser-induced fluorescence instruments for the detection of zinc stress and chlorophyll concentration in bahia grass (Paspalum notatum Flugge.). Remote Sensing of Environment, 84, 572–588.CrossRefGoogle Scholar
Schulze, E.D. (1986). Carbon dioxide and water vapor exchange in response to drought in the atmosphere and in the soil. Annual Review of Plant Physiology, 37, 247–274.CrossRefGoogle Scholar
Schulze, E.D. and Hall, A.E. (1982). Stomatal responses to water loss and CO2 assimilation rates in plants of contrasting environments. In: Encyclopedia of Plant Physiology. Physiological Plant Ecology. Vol. 12B (eds Lange, O. L., Nobel, P., Osmond, C.B. and Ziegler, H.), Springer-Verlag, Berlin, Germany, pp. 181–230.Google Scholar
Schulze, E.D. and Kock, W. (1971). Measurement of primary production with cuvettes. In: Productivity of Forest Ecosystems (ed. Duvigneaud, P.), UNESCO, Paris, France, pp. 141–157.Google Scholar
Schulze, E.D., Ellis, R., Schulze, W., et al. (1996). Diversity, metabolic types and delta 13C carbon isotope ratios in the grass flora of Namibia in relation to growth form, precipitation and habitat conditions. Oecologia, 106, 352–369.CrossRefGoogle Scholar
Schulze, E.D., Fuchs, M. and Fuchs, M.I. (1977). Spacial distribution of photosynthetic capacity and performance in a mountain spruce forest of northern Germany. III. The significance of the evergreen habit. Oecologia, 30, 239–248.CrossRefGoogle Scholar
Schulze, E.D., Hall, A.E., Lange, O.L., et al. (1982). A portable steady state porometer for measuring the carbon dioxide and water vapour exchanges of leaves under natural conditions. Oecologia, 53, 141–145.CrossRefGoogle Scholar
Schulze, E.D., Kelliher, F.M., Körner, C., et al. (1994). Relationships among maximum stomatal conductance, ecosystem surface conductance, carbon assimilation rate, and plant nitrogen nutrition: a global scaling exercise. Annual Review of Ecological Systems, 25, 629–660.CrossRefGoogle Scholar
Schürmann, P. and Buchanan, B. (2008). The ferredoxin-thioredoxin system of oxygenic photosynthesis. Antioxidants and Redox Signalling, 10, 1235–1273.CrossRefGoogle ScholarPubMed
Schurr, U. (1997). Growth physiology: approaches to spatially and temporarily varying problem. Progress in Botany, 59, 355–373.Google Scholar
Schurr, U. and Schulze, E.D. (1996). Effects of drought on nutrient and ABA transport in Ricinus communis. Plant, Cell and Environment, 19, 665–674.CrossRefGoogle Scholar
Schurr, U., Gollan, T. and Schulze, E.D. (1992). Stomatal response to drying soil in relation to changes in the xylem sap composition of Helianthus annuus. II. Stomatal sensitivity to abscisic acid imported from the xylem sap. Plant, Cell and Environment, 15, 561–567.CrossRefGoogle Scholar
Schurr, U., Heckenberger, U., Herdel, K., et al. (2000). Leaf development in Ricinus communis during drought stress: dynamics of growth processes, of cellular structure and of sink-source-transition. Journal of Experimental Botany, 51, 1515–1529.CrossRefGoogle Scholar
Schuster, W.S. and Monson, R.K. (1990). An examination of the advantages of C3-C4 intermediate photosynthesis in warm environments. Plant, Cell and Environment, 13, 903–912.CrossRefGoogle Scholar
Schuster, W.S.F., Sandquist, D.R., Phillips, S.L., et al. (1992). Comparisons of carbon isoptope discrimination in populations of aridland plant species differing in lifespan. Oecologia, 91, 332–337.CrossRefGoogle Scholar
Schwaller, M.R., Schnetzler, C.C. and Marshall, P.E. (1983). The changes in leaf reflectance of sugar maple (Acer saccharum Marsh) seedlings in response to heavy metal stress. International Journal of Remote Sensing, 4, 93–100.CrossRefGoogle Scholar
Schwender, J., Goffman, F., Ohlrogge, J.B., et al. (2004). Rubisco without the Calvin cycle improves the carbon efficiency of developing green seeds. Nature, 432, 779–782.CrossRefGoogle ScholarPubMed
Scott, L. and Vogel, J.C. (2000). Evidence for environmental conditions during the last 20,000 years in Southern Africa from C13 in fossil hyrax dung. Global and Planetary Change, 26, 207–215.CrossRefGoogle Scholar
Seel, W.E. and Press, M.C. (1996). Effects of repeated parasitism by Rhinanthus minor on the growth and photosynthesis of a perennial grass, Poa alpina. New Phytologist, 134, 495–502.CrossRefGoogle Scholar
Seel, W.E., Cooper, R.E. and Press, M.C. (1993). Growth, gas exchange and water use efficiency of the facultative hemiparasite Rhinanthus minor associated with hosts differing in foliar nitrogen concentration. Physiologia Plantarum, 89, 64–70.CrossRefGoogle Scholar
Seelert, H., Poetsch, A., Dencher, N.A., et al. (2000). Proton-powered turbine of a plant motor. Nature, 405, 418–419.CrossRefGoogle ScholarPubMed
Seibt, U., Rajabi, A., Griffiths, H., et al. (2008). Carbon isotopes and water use efficiency: sense and sensitivity. Oecologia, 155, 441–454.CrossRefGoogle ScholarPubMed
Seki, M., Narusaka, M., Ishida, J., et al. (2002). Monitoring the expression profiles of 7000 Arabidopsis genes under drought, cold and high-salinity stresses using a full-length cDNA microarray. Plant Journal, 31, 279–292.CrossRefGoogle ScholarPubMed
Sellers, P.J. (1987). Canopy reflectance, photosynthesis and transpiration: II. The role of biophysics in the linearity of their interdependence. Remote Sensing of Environment, 21, 143–183.CrossRefGoogle Scholar
Sellers, P.J., Berry, J.A., Collatz, G.J., et al. (1992). Canopy reflectance, photosynthesis and transpiration. Part III: a reanalysis using enzyme kinetics-electron transport models of leaf physiology. Remote Sensing of Environment, 42, 187–216.CrossRefGoogle Scholar
Sellers, P.J., Bounoua, L, Collatz, G.J., et al. (1996). Comparison of radiative and physiological effects of doubled atmospheric CO2 on climate. Science, 271, 1402–1406.CrossRefGoogle Scholar
Sellers, P.J., Dickinson, R.E., Randall, D.A., et al. (1997). Modeling the exchanges of energy, water and carbon between continents and the atmosphere. Science, 275, 502–509.CrossRefGoogle ScholarPubMed
Sepulcre-Cantó, G., Zarco-Tejada, P.J., Jiménez-Muñoz, J.C., et al. (2007). Monitoring yield and fruit quality parameters in open-canopy tree crops under water stress. Implications for ASTER. Remote Sensing of the Environment, 107, 455–470.CrossRefGoogle Scholar
Serraj, R., Hash, C.T., Rizvi, S.M.H., et al. (2005). Recent advances in marker-assisted selection for drought tolerance in pearl millet. Plant Production Science, 8, 334–337.CrossRefGoogle Scholar
Sersen, F., Kralova, K. and Bumbalova, A. (1998). Action of mercury on the photosynthetic apparatus of spinach chloroplasts. Photosynthetica, 35, 551–559.CrossRefGoogle Scholar
Sestak, Z., Catsky, J. and Jarvis, P. (1971). Plant Photosynthetic Production. Manual of Methods. Dr. Junk NV, W., Publishers, The Hague 818.Google Scholar
Setterdahl, A.T., Chivers, P.T., Hirasawa, M., et al. (2003). Effect of pH on the oxidation-reduction properties of thioredoxins. Biochemistry, 42, 14877–14884.CrossRefGoogle ScholarPubMed
Seyfried, M. and Fukshansky, L. (1983). Light gradients in plant tissue. Applied Optics, 22, 1402–1408.CrossRefGoogle Scholar
Shaberg, P.G., DeHayes, D.H., Hawley, G.J., et al. (2002). Effects of chronic N fertilization on foliar membranes, cold tolerance, and carbon storage in montane red spruce. Canadian Journal of Forest Research, 32, 1351–1359.CrossRefGoogle Scholar
Shahbazi, M., Gilbert, M., Labouré, A.M., et al. (2007). Dual role of the plastid terminal oxidase in tomato. Plant Physiology, 145, 691–702.CrossRefGoogle ScholarPubMed
Shalitin, D. and Wolf, S. (2000). Cucumber mosaic virus infection affects sugar transport in melon plants. Plant Physiology, 123, 597–604.CrossRefGoogle ScholarPubMed
Shang, W. and Feierabend, J. (1998). Slow turnover of the D1 reaction center protein of photosystem II in leaves of high mountain plants. FEBS Letters, 425, 97–100.CrossRefGoogle ScholarPubMed
Shantz, H.J. and Piemeisel, L.N. (1927). The water requirement of plants at Akron, Colo. Journal of Agriculture Research, 34, 1093–1190.Google Scholar
Shapiro, J.B., Griffin, K.L., Lewis, J.D., et al. (2004). Response of Xanthium strumarium leaf respiration in the light to elevated CO2 concentration, nitrogen availability and temperature. New Phytologist, 162, 377–386.CrossRefGoogle Scholar
Sharifi, M.R., Gibson, A.C. and Rundel, P.W. (1997). Surface dust impacts on gas exchange in Mojave Desert shrubs. Journal of Applied Ecology, 34, 837–46.CrossRefGoogle Scholar
Sharifi, M.R., Nilsen, E.T. and Rundel, P.W. (1982). Biomass and net primary production of Prosopis glandulosa (Fabaceae) in the Sonoran Desert of California. American Journal of Botany, 69, 760–767.CrossRefGoogle Scholar
Sharkey, T.D. (1979). Stomatal responses to light in Xanthium strumarium and other species. Ph.D. Thesis. Michigan State University, East Lansing, p. 95.
Sharkey, T.D. (1985). Photosynthesis in intact leaves of C3 plants: physics, physiology and rate limitations. Botanical Review, 51, 53–105.CrossRefGoogle Scholar
Sharkey, T.D. (1988). Estimating the rate of photorespiration in leaves. Physiologia Plantarum, 73, 147–152.CrossRefGoogle Scholar
Sharkey, T.D. and Schrader, S.M. (2006). High temperature stress. In: Physiology and Molecular Biology of Stress Tolerance in Plants (eds Rao, K.V.M., Raghavendra, A.S. and Reddy, K.J.), Springer, Dordrecht, Netherlands, pp. 101–130.CrossRef
Sharkey, T.D. and Vassey, T.L. (1989). Low oxygen inhibition of photosynthesis is caused by inhibition of starch synthesis. Plant Physiology, 90, 385–387.CrossRefGoogle ScholarPubMed
Sharkey, T.D. and Yeh, S. (2001). Isoprene emission from plants. Annual Review of Plant Physiology and Plant Molecular Biology, 52, 407–436.CrossRefGoogle ScholarPubMed
Sharkey, T.D., Badger, M.R., von Caemmerer, S., et al. (2001). Increased heat sensitivity of photosynthesis in tobacco plants with reduced Rubisco activase. Photosynthesis Research, 67, 147–156.CrossRefGoogle ScholarPubMed
Sharkey, T.D., Bernacchi, C.J., Farquhar, G.D., et al. (2007). Fitting photosynthetic carbon dioxide response curves for C3 leaves. Plant, Cell and Environment, 30, 1035–1040.CrossRefGoogle Scholar
Sharkey, T.D., Imai, K., Farquhar, G.D., et al. (1982). A direct confirmation of the standard method of estimating intercellular partial pressure of CO2. Plant Physiology, 69, 657–659.CrossRefGoogle Scholar
Sharkey, T.D., Vassey, T.L., Vanderveer, P.J., et al. (1991). Carbon metabolism enzymes and photosynthesis in transgenic tobacco (Nicotiana tabaccum L.) having excess phytochrome. Planta, 185, 287–296.CrossRefGoogle Scholar
Sharkey, T.D., Wiberley, A.E. and Donohue, A.R. (2008). Isoprene emission from plants: why and how. Annals of Botany, 101, 5–18.CrossRefGoogle Scholar
Sharkey, T.D., Wise, S.E., Standish, A.J., et al. (2004). CO2 processing, chloroplast to leaf. In: Photosynthetic Adaptation, Chloroplast to Landscape (eds Smith, W.K., Vogelmann, T.C. and Critchley, C.), Springer, New York, USA, pp. 171–206.Google Scholar
Sharma, P.N., Tripathi, A. and Bisht, S.S. (1995). Zinc requirement for stomatal opening in cauliflower. Plant Physiology, 107, 751–756.CrossRefGoogle ScholarPubMed
Sharma-Natu, P. and Ghildiyal, M.C.. (2005). Potential targets for improving photosynthesis and crop yield. Curr Sci., 88, 1918–1928.Google Scholar
Sharp, R.E. (2002). Interaction with ethylene: changing views on the role of abscisic acid in root and shoot growth responses to water stress. Plant, Cell and Environment, 25, 211–222.CrossRefGoogle ScholarPubMed
Sharwood, R.E., von Caemmerer, S., Maliga, P., et al. (2008). The catalytic properties of hybrid Rubisco comprising tobacco small and sunflower large subunits mirror the kinetically equivalent source Rubiscos and can support tobacco growth. Plant Physiology, 146, 83–96.CrossRefGoogle ScholarPubMed
Shatil-Cohen, A., Attia, Z. and Moshelion, M. (2011). Bundle-sheath cell regulation of xylem-mesophyll water transport via aquaporins under drought stress: a target of xylem-borne ABA?The Plant Journal, 67, 72–80.CrossRefGoogle ScholarPubMed
Shaw, M. and MacLachan, G.A. (1954). Chlorophyll content and carbon dioxide uptake of stomatal cells. Nature, 173, 29–30.CrossRefGoogle ScholarPubMed
Sheahan, J.J. (1996). Sinapate esters provide greater UV-B attenuation than flavonoids in Arabidopsis thaliana (Brassicaceae). American Journal of Botany, 83, 679–686.CrossRefGoogle Scholar
Sheard, L.B. and Zheng, N. (2009). Signal advance for abscisic acid. Nature, 462, 575–576.CrossRefGoogle ScholarPubMed
Sheehy, J.E., Mitchell, P.L. and Ferrer, A.B. (2004). Bi-phasic growth patterns in rice. Annals of Botany, 94, 811–817.CrossRefGoogle Scholar
Shen, B., Jensen, R.G. and Bohnert, H.J. (1997). lncreased resistance to oxidative stress in transgenic plants by targeting mannitol biosynthesis to chloroplasts. Plant Physiology, 113, 177–183.CrossRefGoogle Scholar
Shepherd, T. and Griffiths, D.W. (2006). The effects of stress on plant cuticular waxes. New Phytologist, 171, 469–499.CrossRefGoogle ScholarPubMed
Shesták, Z. (ed) (1985). Photosynthesis during leaf development. Dr. W. Junk Publishers, Dordrecht – Boston – Lancaster.
Shi, Z., Liu, S., Liu, X., et al. (2006). Altitudinal variation in photosynthetic capacity, diffusional conductance and δ13C of butterfly bush (Buddleja davidii) plants growing at high elevations. Physiologia Plantarum, 128, 722–731.CrossRefGoogle Scholar
Shikanai, T. (2007). Cyclic electron transport around Photosystem I: genetic approaches. Annual Review of Plant Biology, 58, 199–217.CrossRefGoogle ScholarPubMed
Shimazaki, K., Doi, M., Assmann, S.M., et al. (2007). Light regulation of stomatal movement. Annual Reviews Plant Biology, 58, 219–247.CrossRefGoogle ScholarPubMed
Shinozaki, K., Yamaguchi-Shinozaki, K. and Seki, M. (2003). Regulatory network of gene expression in the drought and cold stress responses. Current Opinion in Plant Biology, 6, 410–417.CrossRefGoogle ScholarPubMed
Shirano, Y., Shimada, H., Kanamaru, K., et al. (2000). Chloroplast development in Arabidopsis thaliana requires the nuclear-encoded transcription factor Sigma B. FEBS Lett, 485, 178–182.CrossRefGoogle ScholarPubMed
Shope, J.C., Peak, D. and Mott, K.A. (2008). Stomatal responses to humidity in isolated epidermis. Plant, Cell and Environment, 31, 1290–1298.CrossRefGoogle Scholar
Shtienberg, D. (1992). Effects of foliar diseases on gas exchange processes: a comparative study. Phytopathology, 82, 760–765.CrossRefGoogle Scholar
Shu, G.G., Baum, D.A. and Mets, L.J. (1999). Detection of gene expression patterns in various plant tissues using non-radioactive mRNA in situ hybridization. The World Wide Web Journal of Biology 4, .Google Scholar
Shulaev, V. and Oliver, D.J. (2006). Metabolic and proteomic markers for oxidative stress. New tools for reactive oxygen species research. Plant Physiology, 141, 367–372.CrossRefGoogle ScholarPubMed
Sicher, R.C., Bahr, J.T. and Jensen, R.G. (1979). Measurement of ribulose-1,5-biphosphate from spinach chloroplasts. Plant Physiology, 64, 876–879.CrossRefGoogle Scholar
Siddiqui, Z.A. and Mahmood, I. (1999). Effects of Meloidogyne incognita, Fusarium oxysporum f.sp. pisi, Rhizobium sp. And different soil types on growth, chlorophyll, and carotenoid pigments of pea. Israel Journal of Plant Sciences, 47, 251–256.CrossRefGoogle Scholar
Siebke, K. and Weis, E. (1995). Imaging of chlorophyll a fluorescence in leaves: topography of photosynthetic oscillations in leaves of Glechoma hederacea. Photosynthesis Research, 45, 225–237.CrossRefGoogle ScholarPubMed
Siegenthaler, U. and Sarmiento, J.L. (1993). Atmospheric carbon dioxide and the ocean. Nature, 365, 119–125.CrossRefGoogle Scholar
Sienkiewicz- Porzucek, A., Sulpice, R., Osorio, S., et al. (2010). Mild reductions in mitochondrial NAD-dependent isocitrate dehydrogenase activity result in altered nitrate assimilation and pigmentation but do not impact growth. Molecular Plant, 3, 156–173.CrossRefGoogle Scholar
Sienkiewicz-Porzucek, A., Nunes-Nesi, A., Sulpice, R., et al. (2008). Mild reductions in mitochondrial citrate synthase activity result in a compromised nitrate assimilation and reduced leaf pigmentation but have no effect on photosynthetic performance or growth. Plant Physiology, 147, 115–127.CrossRefGoogle ScholarPubMed
Sierra-Almeida, A., Cavieres, L.A. and Bravo, L.A. (2009). Freezing resistance varies within the growing season and with elevation in high-Andean species of central Chile. New Phytologist, 182, 461–469.CrossRefGoogle ScholarPubMed
Sigfridsson, K.G.V., Bernát, G., Mamedov, F., et al. (2004). Molecular interference of Cd2+ with Photosystem II. Biochimica et Biophysica Acta, 1659, 19–31.CrossRefGoogle Scholar
Silim, S.N., Guy, R.D., Patterson, T.B. and Livingston, N.J. (2001). Plasticity in water-use efficiency of Picea sitchensis, P. glauca and their natural hybrids. Oecologia, 128, 317–325.CrossRefGoogle ScholarPubMed
Silvera, K., Santiago, L.S. and Winter, K. (2005). Distribution of crassulacean acid metabolism in orchids of Panama: evidence of selection for strong and weak modes. Functional Plant Biology, 32, 397–407.CrossRefGoogle Scholar
Silvera, K., Santiago, L.S., Cushman, J.C., et al. (2009). Crassulacean acid metabolism and epiphytism linked to adaptive radiations in the Orchidaceae. Plant Physiology, 149, 1838–1847.CrossRefGoogle ScholarPubMed
Simioni, G., Le Roux, X., Gignoux, J., et al. (2004). Leaf gas exchange characteristics and water- and nitrogen-use efficiencies of dominant grass and tree species in a West African savanna. Plant Ecology, 173, 233–246.CrossRefGoogle Scholar
Simpson, D. (1995). Biogenic emissions in Europe. 2. Implications for ozone control strategies. Journal of Geophysical Research, 100, 22891–22906.CrossRefGoogle Scholar
Sims, D.A. and Pearcy, R.W. (1992). Response of leaf anatomy and photosynthetic capacity in Alocasia macrorrhiza (Araceae) to a transfer from low to high light. American Journal of Botany, 79, 449–455.CrossRefGoogle Scholar
Sims, D.A., Luo, H., Hastings, S., et al. (2006). Parallel adjustments in vegetation greenness and ecosystem CO2 exchange in response to drought in a Southern California chaparral ecosystem. Remote Sensing of the Environment, 103, 289–303.CrossRefGoogle Scholar
Sims, D.A., Pearcy, and , R.W. (1991). Photosynthesis and respiration in Alocasia macrorrhiza following transfers to high and low light. Oecologia, 86, 447–453.CrossRefGoogle ScholarPubMed
Sims, D.A., Seemann, J.R. and Luo, Y. (1998a). Elevated CO2 concentration has independent effects on expansion rates and thickness of soybean leaves across light and nitrogen gradients. Journal of Experimental Botany, 49, 583–591.CrossRefGoogle Scholar
Sims, D.A., Seemann, J.R. and Luo, Y. (1998b). The significance of differences in the mechanisms of photosynthetic acclimation to light, nitrogen and CO2 for return on investment in leaves. Functional Ecology, 12, 185–194.CrossRefGoogle Scholar
Sinclair, T.R. and Muchow, R.C. (1999). Radiation use efficiency. Advances in Agronomy, 65, 215–215.CrossRefGoogle Scholar
Sinclair, T.R. and Purcell, L.C. (2005). Is a physiological perspective relevant in a “genocentric” age?Journal of Experimental Botany, 56, 2777–2782.CrossRefGoogle Scholar
Sinclair, T.R., Purcell, L.C. and Sneller, C.H. (2004). Crop transformation and the challenge to increase yield potential. Trends in Plant Science, 9, 70–75.CrossRefGoogle ScholarPubMed
Singh, A.K., Li, H. and Sherman, L.A. (2004). Microarray analysis and redox control of gene expression in the cyanobacterium Synechocystis sp. PCC 6803. Physiologia Plantarum, 120, 27–35.CrossRefGoogle ScholarPubMed
Singh, K.K., Chen, C. and Gibbs, M. (1993a). Photoregulation of fructose and glucose respiration in the intact chloroplasts of Chlamydomonas reinhardtii F-60 and spinach. Plant Physiology, 101, 1289–1294.CrossRefGoogle ScholarPubMed
Singh, KK., Chen, C. and Epstein, D.K., et al. (1993b). Respiration of sugars in spinach (Spinacia oleracea), maize (Zea mays), and Chlamydomonas reinhardtii F-60 chloroplasts with emphasis on the hexose kinases. Plant Physiology, 102, 587–593.CrossRefGoogle Scholar
Singh, R. (1993). Photosynthesis characteristics of fruiting structures of cultivated crops. In: Photosynthesis Photoreactions to Plant Productivity (eds Abrol, Y.P., Mohanty, P. and Govindjee, J.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 390–415.Google Scholar
Singsaas, E.L. and Sharkey, T.D. (2000). The effects of high temperature on isoprene synthesis in oak leaves. Plant, Cell and Environment, 23, 751–757.CrossRefGoogle Scholar
Singsaas, E.L., Laporte, M.M., Shi, J.Z., et al. (1999). Kinetics of leaf temperature fluctuation affect isoprene emission from red oak (Quercus rubra) leaves. Tree Physiology, 19, 917–924.CrossRefGoogle ScholarPubMed
Singsaas, E.L., Ort, D.R. and DeLucia, E.H. (2001). Variation in measured values of photosynthetic quantum yield in ecophysiological studies. Oecologia, 128, 15–23.CrossRefGoogle ScholarPubMed
Singsaas, E.L., Ort, D.R. and Delucia, E.H. (2004). Elevated CO2 effects on mesophyll conductance and its consequences for interpreting photosynthetic physiology. Plant, Cell and Environment, 27, 41–50.CrossRefGoogle Scholar
Sinoquet, H., Le Roux, X., Adam, B., et al. (2001). RATP: a model for simulating the spatial distribution of radiation absorption, transpiration and photosynthesis within canopies: application to an isolated tree crown. Plant Cell Environment, 24, 395–406.CrossRefGoogle Scholar
Sinoquet, H., Sonohat, G., Phattaralerphong, J., et al. (2005). Foliage randomness and light interception in 3D digitized trees: an analysis of 3D discretization of the canopy. Plant Cell Environment, 29, 1158–1170.CrossRefGoogle Scholar
Sinoquet, H., Stephan, J., Sonohat, G., et al. (2007). Simple equations to estimate light interception by isolated trees from canopy structure features: assessment with three-dimensional digitized apple trees. New Phytologist, 175, 94–106.CrossRefGoogle ScholarPubMed
Sioris, C.E. and Evans, W.F.J. (1999). Filling in of Fraunhofer and gas-asorption lines in sky spectra as caused rotational Raman scattering. Applied Optics, 38, 2706–2713.CrossRefGoogle ScholarPubMed
Sitch, S., Smith, B., Prentice, I.C., et al. (2003). Evaluation of ecosystem dynamics, plant geography and terrestrial carbon cycling in the LPJ dynamic global vegetation model. Global Change Biology, 9, 161–185.CrossRefGoogle Scholar
Siwko, M.E., Marrink, S.J., de Vries, A.H., et al. (2007). Does isoprene protect plant membranes from thermal shock? A molecular dynamics study. Biochimica et Biophysica Acta – Biomembranes, 1768, 198–206.CrossRefGoogle Scholar
Skillman, J.B., Strain, B.R. and Osmond, C.B. (1996). Contrasting patterns of photosynthetic acclimation and photoinhibition in two evergreen herbs from a winter deciduous forest. Oecologia, 107, 446–455.CrossRefGoogle ScholarPubMed
Skre, O. and Oechel, W.C. (1981). Moss functioning in different taiga ecosystems in interior Alaska. I. Seasonal, phenotypic, and drought effects on photosynthesis and response patterns. Oecologia, 48, 50–59.CrossRefGoogle Scholar
Slafer, G.A., Satorre, E.H. and Andrade, F.H. (1994). Increases in grain yield in bread wheat from breeding and associated physiological changes. In: Genetic Improvement of Field Crops (ed. Slafer, G.A.), Marcel Dekker, New York, USA, pp. 1–68.Google Scholar
Slaton, M.R. and Smith, W.K. (2002). Mesophyll architecture and cell exposure to intercellular air space in alpine desert, and forest species. International Journal of Plant Sciences, 163, 937–948.CrossRefGoogle Scholar
Slatyer, R.O. (1967). Plant-Water Relationships. Academic Press, London, UK.Google Scholar
Slatyer, R.O. (1977a). Altitudinal variation in the photosynthetic characteristics of snow gum, Eucalyptus pauciflora Sieb ex Spreng. III. Temperature response of material grown in contrasting thermal environments. Australian Journal of Plant Physiology, 4, 301–312.CrossRefGoogle Scholar
Slatyer, R.O. (1977b). Altitudinal variation in the photosynthetic characteristics of snow gum, Eucalyptus pauciflora Sieb. ex Spreng. VI. Comparison of field and phytotron responses to growth temperature. Australian Journal of Plant Physiology, 4, 901–916.CrossRefGoogle Scholar
Slatyer, R.O. and Ferrar, P.J. (1977). Altitudinal variation in the photosynthetic characteristics of snow gum, Eucalyptus pauciflora Sieb. ex Spreng. II. Effects of growth temperature under controlled conditions. Australian Journal of Plant Physiology, 4, 289–299.CrossRefGoogle Scholar
Slot, M., Wirth, C., Schumacher, J., et al. (2005). Regeneration patterns in boreal Scots pine glades linked to cold-induced photoihibition. Tree Physiology, 25, 1139–1150.CrossRefGoogle Scholar
Slovik, S. (1996). Early needle senescence and thinning of the crown structure of Picea abies as induced by chronic SO2 pollution. 1. Model deduction and analysis. Global Change Biology, 2, 443–458.CrossRefGoogle Scholar
Smeets, K., Cuypers, A., Lambrechts, A., et al. (2005). Induction of oxidative stress and anti-oxidative mechanisms in Phaseolus vulgaris after Cd application. Plant Physiology and Biochemistry, 43, 437–444.CrossRefGoogle Scholar
Smillie, R.M., Hetherington, S.E. and Davies, W.J. (1999). Photosynthetic activity of the calyx, green shoulder, pericarp, and locular parenchyma of tomato fruit. Journal of Experimental Botany, 50, 707–718.CrossRefGoogle Scholar
Smirnoff, N. (1998). Plant resistance to environmental stress. Current Opinion in Biotechnology, 9, 6.CrossRefGoogle ScholarPubMed
Smith, A.M., Zeeman, S.C. and Smith, S.M. (2005). Starch degradation. Annual Review of Plant Biology, 56, 73–98.CrossRefGoogle ScholarPubMed
Smith, B.N. and Epstein, S. (1971). Two categories of 13C/12C ratios for higher plants. Plant Physiology, 47, 380–384.CrossRefGoogle Scholar
Smith, D.M. and Allen, S.J. (1996). Measurement of sap flow in plant trunks. Journal of Experimental Botany, 47, 1833–1844.CrossRefGoogle Scholar
Smith, J.A.C. (1989). Epiphytic bromeliads. In: Vascular Plants as Epiphytes. Evolution and Ecophysiology. Ecological Studies, vol. 76 (ed. Lüttge, U.), Springer-Verlag, Berlin – Heidelberg – New York, pp. 109–138.CrossRef
Smith, J.A.C. and Lüttge, U. (1985). Day night changes in leaf water relations associated with the rhythm of crassulacean acid metabolism in Kalanchoë daigremontiana. Planta, 163, 272–283.CrossRefGoogle ScholarPubMed
Smith, L.H., Langdale, J.A. and Chollet, R. (1998). A functional Calvin cycle is not indispensable for the light activation of C4 phosphoenolpyruvate carboxylase kinase and its target enzyme in the maize mutant bundle sheath defective2-mu89941tbl. Plant Physiology, 118, 191–197.CrossRefGoogle ScholarPubMed
Smith, M.D., Wilcox, J.C., Kelly, T., et al. (2004). Dominance not richness determines invasibility of tallgrass prairie. Oikos, 106, 253–262.CrossRefGoogle Scholar
Smith, P.R. and Neales, T.F. (1977). Analysis of the effects of virus infection on the photosynthetic properties of peach leaves. Australian Journal of Plant Physiology, 4, 723–732.CrossRefGoogle Scholar
Smith, S.D., Hartsock, T.L. and Nobel, P.S. (1983). Ecophysiology of Yucca brevifolia, an arborescent monocot of the Mojave Desert. Oecologia, 60, 10–17.CrossRefGoogle Scholar
Smith, S.D., Monson, R.K. and Anderson, J.E. (1997). Physiological Ecology of North American Desert Plants. Springer-Verlag, Berlin, Germany.CrossRefGoogle Scholar
Smith, W.K., Knapp, A.K. and Reiners, W.A. (1989). Penumbral effects on sunlight penetration in plant communities. Ecology, 70, 1603–1609.CrossRefGoogle Scholar
Smith, W.K., Vogelmann, T.C., DeLucia, E.H., et al. (1997). Leaf form and photosynthesis. Do leaf structure and orientation interact to regulate internal light and carbon dioxide?BioScience, 47, 785–793.CrossRefGoogle Scholar
Snyder, A.M., Clarck, B.M. and Bungard, R.A. (2005). Light-dependent conversion of carotenoids in the parasitic angiosperm Cuscuta reflexa L. Plant, Cell and Environment, 28, 1326–1333.CrossRefGoogle Scholar
Soar, C.J., Speirs, J., Maffei, S.M., et al. (2004). Gradients in stomatal conductance, xylem sap ABA and bulk leaf ABA along canes of Vitis vinifera cv. Shiraz: molecular and physiological studies investigating their source. Functional Plant Biology, 31, 659–669.CrossRefGoogle Scholar
Soares, A.S., Driscoll, S.P., Olmos, E., et al. (2008). Adaxial/abaxial specification in the regulation of photosynthesis and stomatal opening with respect to light orientation and growth with CO2 enrichment in the C4 species Paspalum dilatatum. New Phytologist, 177, 186–198.Google Scholar
Sobrado, M.A. (1996). Leaf photosynthesis and water loss as influenced by leaf age and seasonal drought in an evergreen tree. Photosynthetica, 32, 563–568.Google Scholar
Sofo, A., Dichio, B. Xiloyannis C., et al. (2004). Effects of different irradiance levels on some antioxidant enzymes and on malondialdehyde content during rewatering in olive tree. Plant Science, 166, 293–302.CrossRefGoogle Scholar
Soitamo, A.J., Piippo, M., Allahverdiyeva, Y., Battchikova, N., Aro, E.M. (2008). Light has a specific role in modulating Arabidopsis gene expression at low temperature. BMC Plant Biology, 8, 13.CrossRefGoogle Scholar
Solhaug, K.A. and Haugen, J. (1998). Seasonal variation of photosynthess in bark from Populus tremula L. Photosynthetica, 35, 411–417.CrossRefGoogle Scholar
Solti, A., Gáspár, L., Mészáros, I., et al. (2008). Impact of iron supply on the kinetics of recovery of photosynthesis in Cd-stressed poplar (Populus glauca). Annals of Botany, 102, 771–782.CrossRefGoogle Scholar
Soltis, D.E. and Soltis, P.S. (2003). The role of phylogenetics in comparative genetics. Plant Physiology, 132, 1790–1800.CrossRefGoogle ScholarPubMed
Soltis, D.E., Soltis, P.S. and Zanis, M.J. (2002). Phylogeny of seed plants based on evidence from eight genes. American Journal of Botany, 89, 1670–1681.CrossRefGoogle ScholarPubMed
Soltis, P.S. (2005). Ancient and recent polyploidy in angiosperms. The New Phytologist, 166, 5–8.CrossRefGoogle ScholarPubMed
Soltis, P.S., Soltis, D.E., Wolf, P.G., et al. (1999). The phylogeny of land plants inferred from 18S rDNA sequences: pushing the limits of rDNA signal?Molecular Biology and Evolution, 16, 1774–1784.CrossRefGoogle ScholarPubMed
Sonohat, G., Sinoquet, H., Varlet-Grancher, C., et al. (2002). Leaf dispersion and light partitioning in three-dimensionally digitized tall fescue-white clover mixtures. Plant Cell Environment, 25, 529–538.CrossRefGoogle Scholar
Sonoike, K. (1996). Photoinhibition of photosystem I: its physiological significance in the chilling sensitivity of plants. Plant and Cell Physiology, 37, 239–247.CrossRefGoogle Scholar
Soolanayakanahally, R.Y., Guy, R.D., Silim, S.N., et al. (2009). Enhanced assimilation rate and water use efficiency with latitude through increased photosynthetic capacity and internal conductance in balsam poplar (Populus balsamifera L.). Plant Cell and the Environment, 32, 1821–1832.CrossRefGoogle Scholar
Soriano, M.A., Orgaz, F., Villalobos, F., et al. (2004). Efficiency of water use of early planting of sunflower. European Journal of Agronomy, 21, 465–476.CrossRefGoogle Scholar
Soukupová, J., Cséfalvay, L., Urbam, O., et al. (2008). Annual variation of the steady state chlorophyll fluorescence emission of evergreen plants in temperate zone. Functional Plant Biology, 35, 63–76.CrossRefGoogle Scholar
Soukupová, J., Smatanová, S., Nedbal, L., et al. (2003). Plant response to destruxins visualized by imaging of chlorophyll fluorescence. Physiologia Plantarum, 118, 399–405.CrossRefGoogle Scholar
Soussana, J.F., Allard, V., Pilegaard, K., et al. (2007). Full accounting of the greenhouse gas (CO2, N2O, CH4) budget of nine European grassland sites. Agriculture, Ecosystems and Environment, 121, 121–134.CrossRefGoogle Scholar
Souza, C.R., Maroco, J., Santos, T., et al. (2005). Impact of deficit irrigation on water use efficiency and carbon isotope composition (δ13C) of field-grown grapevines under Mediterranean Climate. Journal of Experimental Botany, 56, 2163–2172.CrossRefGoogle ScholarPubMed
Souza, R.P. de, Machado, E.C., Silva, J.A.B., et al. (2004). Photosynthetic gas exchange, chlorophyll fluorescence and some associated metabolic changes in cowpea (Vigna unguiculata) during water stress and recovery. Environmental and Experimental Botany, 51, 45–56.CrossRefGoogle Scholar
Sowinska, M., Cunin, B., Heisel, F., et al. (1999). New UV-A laser-induced fluorescence imaging system for near-field remote sensing of vegetation: characteristics and performances. In: Proceedings of the SPIE Conference on Laser Radar Technology, Orlando Florida: The International Society for Optical Engineering, pp. 91–102.CrossRef
Sowinska, M., Heisel, F., Miehe, J.A., et al. (1996). Remote sensing of plants by streak camera lifetime measurements of the chlorophyll a emission. Journal of Plant Physiology, 148, 638–644.CrossRefGoogle Scholar
Specht, R.L. and Moll, E.J. (1983). Mediterranean-type shrublands and sclerophyllous shrublands of the world: an overview. In: Mediterranean-type Ecosystems. The Role of Nutrients (eds Kruger, F.J., Mitchell, D.T. and Jarvis, J.U.M.), Springer-Verlag, Berlin – Heidelberg – New York – Tokyo, pp. 41–65.
Spicer, R.A. (1993). Palaeoecology, past climate systems, and C3/C4 photosynthesis. Chemosphere, 27, 947–978.CrossRefGoogle Scholar
Spikes, J.D. and Stout, M. (1955). Photochemical activity of chloroplasts isolated from sugar beet infected with virus yellows. Science, 12, 375–376.CrossRefGoogle Scholar
Spitters, C.J.T., Toussaint, H.A.J.M. and Goudriaan, J. (1986). Separating the diffuse and direct component of global radiation and its implications for modelling canopy photosynthesis. Part I. Components of incoming radiation. Agricultural and Forest Meteorology, 38, 217–229.CrossRefGoogle Scholar
Spreitzer, R.J. and Salvucci, M.E. (2002). Rubisco: structure, regulatory interactions, and possibilities for a better enzyme. Annual Review of Plant Biology, 53, 449–475.CrossRefGoogle ScholarPubMed
Stanghellini, C. (2005). Irrigation water: use, efficiency and economics. In: Improvement in Water Use Efficiency in Protected Crops. Junta de Andalucía, Sevilla, Spain, pp. 23–33.
Stanhill, G. and Cohen, S. (2001). Global dimming: a review of the evidence for a widespread and significant reduction in global radiation with discussion of its probable causes and possible agricultural consequences. Agricultural and Forest Meterorology, 107, 255–278.CrossRefGoogle Scholar
Starr, G. and Oberbauer, S.F. (2003). Photosynthesis of Arctic evergreens under snow: implications for tundra ecosystem carbon balance. Ecology, 84, 1415–1420.CrossRefGoogle Scholar
Staudt, M. and Lhoutellier, L. (2007). Volatile organic compound emission from holm oak infested by gypsy moth larvae: evidence for distinct responses in damaged and undamaged leaves. Tree Physiology, 27, 1433–1440.CrossRefGoogle ScholarPubMed
Steduto, P. and Hsiao, T.C. (1998). Maize canopies under two soil water regimes – II. Seasonal trends of evapotranspiration, carbon dioxide assimilation and canopy conductance, and as related to leaf-area index. Agricultural and Forest Meteorology, 89, 185–200.CrossRefGoogle Scholar
Steele, K., Price, A.H., Shashidhar, H.E., et al. (2006). Marked-assisted selection to intogress rice QTLs controlling root traits into an Indian upland rice variety. Theoretical and Applied Genetics, 112, 208–221.CrossRefGoogle Scholar
Steele, M.R., Gitelson, A.A. and Rundquist, D. (2008). A comparison of two techniques for nondestructive measurement of chlorophyll content in grapevine leaves. Agronomy Journal, 100, 779–782.CrossRefGoogle Scholar
Steigmiller, S., Turina, P. and Gräber, P. (2008). The thermodynamic H+/ATP ratios of the H+-ATP synthases from chloroplasts and Escherichia coli. Proceedings of the National Academy of Sciences USA, 105, 3745–3750.CrossRefGoogle Scholar
Stenberg, P., DeLucia, E.H., Schoettle, A.W., et al. (1995). Photosynthetic light capture and processing from cell to canopy, In: Resource Physiology of Conifers Acquisition, Allocation, and Utilization (eds Smith, W.K. and Hinckley, T.M.), Academic Press, London, pp. 3–38.CrossRef
Stenström, A., Jónsdóttir, I.S. and Augner, M. (2002). Genetic and Environmental Effects on Morphology in Clonal Sedges in the Eurasian Arctic. American Journal of Botany, 89, 1410–1421.CrossRefGoogle ScholarPubMed
Stepien, P. and Johnson, G.N. (2009). Contrasting responses of photosynthesis to salt stress in the glycophyte Arabidopsis and the halophyte Thellungiella: role of the plastid terminal oxidase as an alternative electron sink. Plant physiology, 149, 1154–1165.CrossRefGoogle ScholarPubMed
Sterck, F.J. and Bongers, F. (1998). Ontogenetic changes in size, allometry, and mechanical design of tropical rain forest trees. American Journal of Botany, 85, 266–272.CrossRefGoogle ScholarPubMed
Stern, D.B., Hanson, M.R. and Barkan, A. (2004). Genetics and genomics of chloroplast biogenesis: maize as a model system. Trends Plant Science, 9, 293–301.CrossRefGoogle ScholarPubMed
Sternberg, L. and De Niro, M. (1983). Biogeochemical implications of the isotopic equilibrium fractionation factor between oxygen atoms of acetone and water. Geochimica et Cosmochimica Acta, 47, 2271–2274.CrossRefGoogle Scholar
Sternberg, L., De Niro, M. and Savidge, R. (1986). Oxygen isotope exchange between metabolites and water during biochemical reactions leading to cellulose synthesis. Plant Physiology, 82, 423–427.CrossRefGoogle Scholar
Stevens, C.L.R., Schultz, D., Van Baalen, C., et al. (1975). Oxygen isotope fractionation during photosynthesis in a blue-green and a green alga. Plant Physiology, 65, 126–129.CrossRefGoogle Scholar
Stewart, G.R. and Press, M.C. (1990). The physiology and biochemistry of parasitic angiosperms. Annual Review of Plant Physiology and Plant Molecular Biology, 41, 127–151.CrossRefGoogle Scholar
Stewart, G.R., Turnbull, M.H., Schmidt, S., et al. (1995). C13 Natural abundance in plant communities along a rainfall gradient – a biological integrator of water availability. Australian Journal of Plant Physiology, 22, 51–55.CrossRefGoogle Scholar
Stewart, J.B. (1988). Modelling surface conductance of Pine forest. Agricultural and Forest Meteorology, 43, 19–35.CrossRefGoogle Scholar
Steyn, W.J., Wand, S.J.E., Holcroft, D.M., et al. (2002). Anthocyanins in vegetative tissues: a proposed unified function in photoprotection. New Phytologist, 155, 349–361.CrossRefGoogle Scholar
Steyn, W.J., Wand, S.J.E., Jacobs, G., et al. (2009). Evidence for a photoprotective function of low-temperature-induced anthocyanin accumulation in apple and pear peel. Physiologia Plantarum, 136, 461–472.CrossRefGoogle ScholarPubMed
Still, C.J., Berry, J.A., Collatz, G.J., et al. (2003). Global distribution of C3 and C4 vegetation: carbon cycle implications. Global Biogeochemical Cycles, 17, Article number 1006.CrossRefGoogle Scholar
Stiller, I., Dulai, S., Kondrák, M., et al. (2008). Effects of drought on water content and photosynthetic parameters in potato plants expressing the trehalose-6-phosphate synthase gene of Saccharomyces cerevisiae. Planta, 227, 299–308.CrossRefGoogle ScholarPubMed
Stiller, W.N., Read, J.J., Constable, G.A., et al. (2005). Selection for water use efficiency traits in a cotton breeding program: cultivar differences. Crop Sciences, 45, 1107–113.CrossRefGoogle Scholar
Stimler, K., Montzka, S.A., Berry, J.A., et al. (2010). Relationships between carbonyl sulfide (COS) and CO2 during leaf gas exchange. New Phytologist, 186, 869–878.CrossRefGoogle ScholarPubMed
Stirling, C.M., Aguilera, C., Baker, N.R., et al. (1994). Changes in the photosynthetic light response curve during leaf development of field grown maize with implications for modelling canopy photosynthesis. Photosynthesis Research, 42, 217–225.CrossRefGoogle ScholarPubMed
Stitt, M. (1991). Rising CO2 levels and their potential significance for carbon flow in photosynthetic cells. Plant, Cell and Environment, 14, 741–762.CrossRefGoogle Scholar
Stitt, M. and Fernie, A.R. (2003). From measurements of metabolites to metabolomics: an ‘on the fly’ perspective illustrated by recent studies of carbon-nitrogen interactions. Current Opinion in Biotechnology, 14, 136–144.CrossRefGoogle ScholarPubMed
Stitt, M. and Hurry, V. (2002). A plant for all seasons: alterations in photosynthetic carbon metabolism during cold acclimation in Arabidopsis. Current Opinion in Plant Biology, 5, 199–206.CrossRefGoogle ScholarPubMed
Stitt, M. and Krapp, A. (1999). The interaction between elevated carbon dioxide and nitrogen nutrition: the physiological and molecular background. Plant, Cell and Environment, 22, 583–621.CrossRefGoogle Scholar
Stitt, M., Müller, C., Matt, P., et al. (2002). Steps towards an integrated view of nitrogen metabolism. Journal of Experimental Botany, 53, 959–970.CrossRefGoogle ScholarPubMed
Stobart, A.K., Griffiths, W.T., Ameen-Bukhari, , et al. (1985). The effect of Cd2+ on the biosynthesis of chlorophyll in leaves of barley. Physiologia Plantarum, 63, 293–298.CrossRefGoogle Scholar
Stokes, V.J., Morecroft, M.D. and Morison, J.I.L. (2006). Boundary layer conductance for contrasting leaf shapes in a deciduous broadleaved forest canopy. Agriculture for Meteorology, 139, 40–54.CrossRefGoogle Scholar
Stoll, M.P., Court, A., Smorenburg, K., et al. (1999). FLEX – Fluorescence Explorer. In: Remote Sensing for Earth Sciences, Ocean and Sea Ice Applications, SPIE, Florence, Italy, pp. 487–494.Google Scholar
Strain, B.R. and Bazzaz, F.A. (1983). Terrestrial plant communities. In: CO2 and Plants: The Response of Plants to Rising Levels of Atmospheric Carbon Dioxide (ed. Lemon, E.R.), Westview Press, Boulder, USA, pp. 177–222.Google Scholar
Strand, A., Foyer, C.H., Gustafsson, P., et al. (2003). Altering flux through the sucrose biosynthesis pathway in transgenic Arabidopsis thaliana modifies photosynthetic acclimation at low temperatures and the development of freezing tolerance. Plant Cell Environment, 26, 523–535.CrossRefGoogle Scholar
Strand, A., Hurry, V., Gustafsson, P., et al. (1997). Development of Arabidopsis thaliana leaves at low temperatures releases the suppression of photosynthesis and photosynthetic gene expression despite the accumulation of soluble carbohydrates. Plant Journal, 12, 605–614.CrossRefGoogle ScholarPubMed
Strand, A., Hurry, V., Henkes, S., et al. (1999). Acclimation of Arabidopsis leaves developing at low temperatures. Increasing cytoplasmic volume accompanies increased activities of enzymes in the Calvin cycle and in the sucrose-biosynthesis pathway. Plant Physiology, 119, 1387–1397.CrossRefGoogle ScholarPubMed
Strand, M., Lundmark, T., Söderbergh, I., et al. (2002). Impacts of seasonal air and soil temperatures on photosynthesis in Scots pine trees. Tree Physiology, 22, 839–847.CrossRefGoogle ScholarPubMed
Strasser, R.J., Srivasta, A. and Govindjee, J. (1995). Polyphasic chlorophyll-alpha fluorescence transient in plants and cyanobacteria. Photochemistry and Photobiology, 61, 32–42.CrossRefGoogle Scholar
Strauss-Debenedetti, S. and Bazzaz, F.A. (1991). Plasticity and acclimation to light in tropical Moraceae of different sucessional positions. Oecologia, 87, 377–387.CrossRefGoogle ScholarPubMed
Streb, P. and Feierabend, J. (1999). Significance of antioxidants and electron sinks for the cold-hardening-induced resistance of winter rye leaves to photo-oxidative stress. Plant, Cell Environment, 22, 1225–1237.CrossRefGoogle Scholar
Streb, P., Aubert, S. and Bligny, R. (2003b). High temperature effects on light sensitivity in the two high mountain plant species Soldanella alpina (L) and Ranunculus glacialis (L). Plant Biology, 5, 432–440.CrossRefGoogle Scholar
Streb, P., Aubert, S., Gout, E., et al. (2003a). Cold- and light-induced changes of metabolite and antioxidant levels in two high mountain plant species Soldanella alpina and Ranunculus glacialis and a lowland species Pisum sativum. Physiologia Plantarum, 118, 96–104.CrossRefGoogle Scholar
Streb, P., Aubert, S., Gout, E., et al. (2003c). Reversibility of cold- and light-stress tolerance and accompanying changes of metabolite and antioxidant levels in the two high mountain plant species Soldanella alpina and Ranunculus glacialis. Journal of Experimental Botany, 54, 405–418.CrossRefGoogle ScholarPubMed
Streb, P., Feierabend, J. and Bligny, R. (1997). Resistance to photoinhibition of photosystem II and catalase and antioxidative protection in high mountain plants. Plant Cell and Environment, 20, 1030–1040.CrossRefGoogle Scholar
Streb, P., Josse, E.M., Gallouët, E., et al. (2005). Evidence for alternative electron sinks to photosynthetic carbon assimilation in the high mountain plant species Ranunculus glacialis. Plant, Cell Environment, 28, 1123–1135.CrossRefGoogle Scholar
Streb, P., Shang, W. and Feierabend, J. (1999). Resistance of cold-hardened winter rye leaves (Secale cereale L.) to photo-oxidative stress. Plant, Cell Environment, 22, 1211–1223.CrossRefGoogle Scholar
Streb, P., Shang, W., Feierabend, J., et al. (1998). Divergent strategies of photoprotection in high-mountain plants. Planta, 207, 313–324.CrossRefGoogle Scholar
Strimbeck, G.R., Kjellsen, T.D., Schaberg, P.G., et al. (2007). Cold in the common garden: comparative low-temperature tolerance of boreal and temperate conifer foliage. Trees, 21, 557–567.CrossRefGoogle Scholar
Strong, G.L., Bannister, P. and Burritt, D. (2000). Are mistletoes shade plants? CO2 assimilation and chlorophyll fluorescence of temperate mistletoes and their hosts. Annals of Botany, 85, 511–519.CrossRefGoogle Scholar
Studart-Guimaraes, C., Fait, A., Nunes- Nesi, A., et al. (2007). Reduced expression of succinyl-coenzyme a ligase can be compensated for by up-regulation of the γ-aminobutyrate shunt in illuminated tomato leaves. Plant Physiology, 145, 626–639.CrossRefGoogle ScholarPubMed
Stuntz, S. and Zotz, G. (2001). Photosynthesis in vascular epiphytes: a survey of 27 species of diverse taxonomic origin. Flora, 196, 132–141.CrossRefGoogle Scholar
Stylinski, C.D., Gamon, J.A. and Oechel, W.C. (2002). Seasonal patterns of reflectance indices, carotenoid pigments and photosynthesis of evergreen chaparral species. Oecologia, 131, 366–374.CrossRefGoogle ScholarPubMed
Stylinski, C.D., Oechel, W.C., Gamon, J.A., et al. (2000). Effects of lifelong [CO2] enrichment on carboxylation and light utilization of Quercus pubescens Willd. examined with gas exchange, biochemistry and optical techniques. Plant, Cell and Environment, 23, 1353–1362.CrossRefGoogle Scholar
Suárez, L., Zarco-Tejada, P.J., Berni, J.A.J., et al. (2009) Modeling PRI for water stress detection using radiative transfer models. Remote Sensing of the Environment, 113, 730–744.CrossRefGoogle Scholar
Suárez, L., Zarco-Tejada, P.J., Sepulcre-Cantó, G., et al. (2008). Assessing canopy PRI for water stress detection with diurnal airborne imagery. Remote Sensing of the Environment, 112, 560–575.CrossRefGoogle Scholar
Suess, H.E. (1955). Radiocarbon concentration in modern wood. Science, 122, 415–417.CrossRefGoogle Scholar
Sugano, S.S., Shimada, T., Imai, Y., et al. (2010). Stomagen positively regulates stomatal density in Arabidopsis. Nature, 463, 241–244.CrossRefGoogle ScholarPubMed
Sugita, M. and Sugiura, M. (1996). Regulation of gene expression in chloroplasts of higher plants. Plant Molecular Biology, 32, 315–326.CrossRefGoogle ScholarPubMed
Suh, H.J., Kim, C.S., Lee, J.Y., et al. (2002). Photodynamic effect of iron excess on photosystem II function in pea plants. Photochemistry and Photobiology, 75, 513–518.2.0.CO;2>CrossRefGoogle ScholarPubMed
Sulpice, R., Tschoep, H., Von Korff, M., et al. (2007). Description and applications of a rapid and sensitive non-radioactive microplate-based assay for maximum and initial activity of D-ribulose-1,5-bisphosphate carboxylase/oxygenase. Plant Cell Environment, 30, 1163–1175.CrossRefGoogle ScholarPubMed
Sun, G., Ji, Q., Dilcher, D.L., et al. (2002). Archaefructaceae, a new basal angiosperm family. Science, 296, 899–904.CrossRefGoogle ScholarPubMed
Sun, J., Nishio, J.N. and Vogelmann, T.C. (1996). High-light effects on CO2 fixation gradients across leaves. Plant Cell Environment, 19, 1261–1271.CrossRefGoogle Scholar
Sun, J., Nishio, J.N. and Vogelmann, T.C. (1998). Green light drives CO2 fixation deep within leaves. Plant Cell Physiology, 39, 1020–1026.CrossRefGoogle Scholar
Sun, J., Okita, T.W. and Edwards, G.E. (1999). Modification of carbon partitioning, photosynthetic capacity, and O2 sensitivity in Arabidopsis plants with low ADP-glucose pyrophosphorylase activity. Plant Physiology, 119, 367–276.CrossRefGoogle ScholarPubMed
Sun, P., Grignetti, A., Liu, S., et al. (2008). Associated changes in physiological parameters and spectral reflectance indices in olive (Olea europaea L.) leaves in response to different levels of water stress. International Journal of Remote Sensing, 29, 1725–1743.CrossRefGoogle Scholar
Sun, Q., Yoda, K., Suzuki, M., et al. (2003). Vascular tissue in the stems and roots of woody plants can conduct light. Journal of Experimental Botany, 54, 1627–1635.CrossRefGoogle ScholarPubMed
Sun, W.H, Verhoeven, A.S., Bugos, R.C., et al. (2001). Suppression of zeaxanthin formation does not reduce photosynthesis and growth of transgenic tobacco under field conditions. Photosynthesis Research, 67, 41–50.CrossRefGoogle Scholar
Sun, Z.J., Livingston, N.J., Guy, R.D., et al. (1996). Stable carbon isotopes as indicators of increased water use efficiency and productivity in white spruce (Picea glauca (Moench) Voss) seedlings. Plant, Cell and Environment, 19, 887–894.CrossRefGoogle Scholar
Sundberg, M.D. (1986). A comparison of stomatal distribution and length in succulent and non-succulent desert plants. Phytomorphology, 36, 53–66.Google Scholar
Sundblad, L.G., Schröder, W.P. and Akerlung, H.E. (1988). S-state distribution and redox state of QA in barley in relation to luminescence decay kinetics. Biochimica et Biophysica Acta, 973, 47–52.CrossRefGoogle Scholar
Sundbom, E., Strand, M. and Hällgreen, J.E. (1982). Temperature-induced fluorescence changes. A screening method for frost tolerance of potato (Solanum sp.). Plant Physiology, 70, 1299–1302.CrossRefGoogle Scholar
Sundby, C. and Andersson, B. (1985). Temperature-induced reversible migration along the thylakoid membrane of photosystem II regulates it association with LHC-II. FEBS Letters, 191, 24–28.CrossRefGoogle Scholar
Suni, T., Berninger, F., Vesala, T., et al. (2003). Air temperature triggers the commencement of evergreen boreal forest photosynthesis in spring. Global Change Biology, 9, 1410–1426.Google Scholar
Susiluoto, S., Perämäki, M., Nikinmaa, E., et al. (2007). Effects of sink removal on transpiration at the treeline: implications for growth limitation hypothesis. Environmental and Experimental Botany, 60, 334–339.CrossRefGoogle Scholar
Susiluoto, S., Pumpanen, J. and Berninger, F. (2008). Effects of grazing on the vegetation structure and carbon balance of Scandinavian fell vegetation. Arctic, Antarctic, and Alpine Research, 40, 422–431.CrossRefGoogle Scholar
Suwignyo, R.A., Nose, A., Kawamitsu, Y., e al. (1995). Effects of manipulations of source and sink on carbon exchange rate and some enzymes of sucrose metabolism in leaves of soybean [Glycine max (L.) Merr.]. Plant Cell Physiology, 36, 1439–1446.Google Scholar
Suyker, A.E. and Verma, S.B. (1993). Eddy correlation measurements of CO2 flux using a closed-path sensor: theory and field tests against an open-path sensor. Boundary Layer Meteorology, 64, 391–407.CrossRefGoogle Scholar
Suyker, A.E. and Verma, S.B. (2001). Year-round observations of the net ecosystem exchange of carbon dioxide in a native tallgrass prairie. Global Change Biology, 7, 279–289.CrossRefGoogle Scholar
Suzuki, T., Eiguchi, M., Kumamaru, T., et al. (2008). MNU-induced mutant pools and high performance TILLING enable finding of any gene mutation in rice. Molecular Genetics and Genomics, 279, 213–223.CrossRefGoogle Scholar
Sveshnikov, D., Ensminger, I., Ivanov, A.G., et al. (2005). Excitation energy partitioning and quenching during cold acclimation in Scots pine. Tree Physiology, 26, 325–336.CrossRefGoogle Scholar
Swarbrick, P.J., Schulze-Lefert, P. and Scholes, J.D. (2006). Metabolic consequences of susceptibility and resistance (race-specific and broad-spectrum) in barley leaves challenged with powdery mildew. Plant, Cell and Environment, 29, 1061–1076.CrossRefGoogle ScholarPubMed
Swarthout, D., Harper, E., Judd, S., et al. (2009). Measures of leaf-level water-use efficiency in drought stressed endophyte infected and non-infected tall fescue grasses. Environmental and Experimental Botany, 66, 88–93.CrossRefGoogle Scholar
Sweetlove, L.J., Lytovchenko, A., Morgan, M., et al. (2006). Mitochondrial uncoupling protein is required for efficient photosynthesis. Proceedings of the National Academy of Sciences USA, 103, 19, 587–19, 592.CrossRefGoogle ScholarPubMed
Swinbank, W.C. (1951). The measurement of vertical transfer of heat and water vapor by eddies in the lower atmosphere. Journal of Meterology, 8, 135–145.2.0.CO;2>CrossRefGoogle Scholar
Sykes, M.T. and Prentice, I.C. (1996). Climate change, tree species distributions and forest dynamics: a case study in the mixed conifer/northern hardwoods zone of northern Europe. Climatic Change, 34, 161–177.CrossRefGoogle Scholar
Syvertsen, J.P., Lloyd, J., McConchie, C., et al. (1995). On the relationship between leaf anatomy and CO2 diffusion through the mesophyll of hypostomatous leaves. Plant Cell and Environment, 18, 149–157.CrossRefGoogle Scholar
Szarek, S.R. and Ting, I.P. (1975). Physiological responses to rainfall in Opuntia basiliaris (Cactaceae). American Journal of Botany, 62, 602–609.CrossRefGoogle Scholar
Szarek, S.R., Holthe, P.A. and Ting, I.P. (1987). Minor physiological response to elevated CO2 by the CAM plant Agave vilmoriniana. Plant Physiology, 83, 938–940.CrossRefGoogle Scholar
Sziráki, I.B., Mustarky, L.A., Faludi-Daniel, A., et al. (1984). Alterations in chloroplast ultraestructure and chlorophyll content in rust infected pinto beans at different stages of disease development. Phytopathology, 74, 77–84.CrossRefGoogle Scholar
Tadaki, Y. (1977). Forest biomass. Leaf biomass. In: Primary Productivity of Japanese Forests. Productivity of Terrestrial Communities. Japanese Commitee for the International Biological Program (eds Shidei, T. and Kira, T.), University of Tokyo Press, Tokyo, pp. 39–44.Google Scholar
Takahashi, H. and Ehara, Y. (1992). Changes in the activity and the polypeptide composition of the oxygen-evolving complex in photosystem II of tobacco leaves infected with cucumber mosaic virus strain Y. Molecular Plant-Microbe Interactions, 5, 269–272.CrossRefGoogle ScholarPubMed
Takahashi, N., Ling, P.P. and Frantz, J.M. (2008). Considerations for accurate whole plant photosynthesis measurement. Environmental Control in Biology, 46, 91–101.CrossRefGoogle Scholar
Takahashi, S. and Badger, M.R. (2011). Photoprotection in plants: a new light on photosystem II damage. Trends in Plant Science, 16, 53–60.CrossRefGoogle ScholarPubMed
Takahashi, S. and Murata, N. (2008). How do environmental stresses accelerate photoinhibition?Trends in Plant Science, 13, 179–182.CrossRefGoogle ScholarPubMed
Takahashi, S., Milward, S.E., Fan, D.Y., et al. (2009). How does cyclic electron flow alleviate photoinhibition in Arabidopsis?Plant Physiology, 149, 1560–1567.CrossRefGoogle ScholarPubMed
Takamiya, K.I., Tsuchiya, T. and Ohta, H. (2000). Degradation pathway(s) of chlorophyll: what has gene cloning revealed?Trends Plant Science, 10, 426–431.CrossRefGoogle Scholar
Takashima, T., Hikosaka, K. and Hirose, T. (2004). Photosynthesis or persistence: nitrogen allocation in leaves of evergreen and deciduous Quercus species. Plant, Cell and Environment, 27, 1047–1054.CrossRefGoogle Scholar
Takenaka, A. (1994). Effects of leaf blade narrowness and petiole length on the light capture efficiency of a shoot. Ecologcal Research, 9, 109–114.CrossRefGoogle Scholar
Takenaka, A., Takahashi, K. and Kohyama, T. (2001). Optimal leaf display and biomass partitioning for efficient light capture in an understory palm, Licuala arbuscula. Functional Ecology, 15, 660–668.CrossRefGoogle Scholar
Takeuchi, Y., Kubiske, M.E., Isebrands, J.G., et al. (2001). Photosynthesis, light and nitrogen relationships in a young deciduous forest canopy under open-air CO2 enrichment. Plant, Cell and Environment, 24, 1257–1268.CrossRefGoogle Scholar
Takizawa, K., Kanazawa, A. and Kramer, D.M. (2008). Depletion of stromal Pi induces high ‘energy dependent’ antenna exciton quenching (qE) by decreasing proton conductivity at CFO-CF1 ATP synthase. Plant, Cell and Environment, 31, 235–243.CrossRefGoogle Scholar
Talbott, L.D. and Zeiger, E. (1998). The role of sucrose in guard cell osmoregulation. Journal of Experimental Botany, 49, 329–337.CrossRefGoogle Scholar
Tan, W. and Hogan, G.D. (1995). Limitations to net photosynthesis as affected by nitrogen status in jack pine (Pinus banksiana Lamb.) seedlings. Journal of Experimental Botany, 46, 407–413.CrossRefGoogle Scholar
Tanaka, R. and Tanaka, A. (2007). Tetrapyrrole biosynthesis in higher plants. Annual Review of Plant Biology, 58, 321–346.CrossRefGoogle ScholarPubMed
Taneda, H. and Tateno, M. (2005). Hydraulic conductivity, photosynthesis and leaf water balance in six evergreen woody species from fall to winter. Tree Physiology, 25, 299–306.CrossRefGoogle ScholarPubMed
Tang, A.Ch. and Boyer, J.S. (2007). Leaf shrinkage decreases porosity at low water potentials in sunflower. Functional Plant Biology, 34, 24–30.CrossRefGoogle Scholar
Tang, J.Y., Zielinski, R.E., Zangerl, A.R., et al. (2006). The differential effects of herbivory by first and fourth instars of Trichoplusia ni (Lepidoptera: Noctuidae) on photosynthesis in Arabidopsis thaliana. Journal of Experimental Botany, 57, 27–536.CrossRefGoogle ScholarPubMed
Tang, Y.H., Kachi, N., Furukawa, A., et al. (1999). Heterogeneity of light availability and its effects on simulated carbon gain of tree leaves in a small gap and the understory in a tropical rain forest. Biotropica, 31, 268–278.CrossRefGoogle Scholar
Tanner, C.B. and Sinclair, T.R. (1983). Efficient water use in crop production: research or re-search? In: Limitations to efficient water use in crop production (ed. Taylor, H.M.. Jordan, W.R. and Sinclair T.R.), ASA, CSSA, SSSA, Madison, Wisconsin, USA, pp. 1–25.
Tans, P.P., Berry, J.A. and Keeling, R.F. (1993). Oceanic 13C data: a new window on CO2 uptake by oceans. Global Biogeochemical Cycles, 7, 353–368.CrossRefGoogle Scholar
Tans, P.P., Fung, I.Y. and Takahashi, T. (1990). Observational constraints on the global atmospheric CO2 budget. Science, 247, 1431–1438.CrossRefGoogle Scholar
Tanz, S.K., Tetu, S.G., Vella, N.G.F., et al. (2009). Loss of the transit peptide and an increase in gene expression of an ancestral chloroplastic carbonic anhydrase were instrumental in the evolution of the cytosolic C4 carbonic anhydrase in Flaveria. Plant Physiology, 150, 1515–1529.CrossRefGoogle ScholarPubMed
Tao, Y., Xie, Z., Chen, W., et al. (2003). Quantitative nature of Arabidopsis responses during compatible and incompatible interactions with the bacterial pathogen Pseudomonas syringae. Plant Cell, 15, 317–330.CrossRefGoogle ScholarPubMed
Tardieu, F. and Simonneau, T. (1998). Variability among species of stomatal control under fluctuating soil water status and evaporative demand: modelling isohydric and anisohydric behaviours. Journal of Experimental Botany, 49, 419–432.CrossRefGoogle Scholar
Tardieu, F., Granier, C. and Muller, B. (1999). Research review. Modelling leaf expansion in a fluctuating environment: are changes in specific leaf area a consequence of changes in expansion rate?New Phytologist, 143, 33–44.CrossRefGoogle Scholar
Tardieu, F., Katerji, N., Bethenod, O., et al. (1991a). Leaf stomatal conductance in the field: its relationship with measured plant water potentials, mechanical constraints and ABA concentration in the xylem sap. Plant, Cell and Environment, 14, 121–126.CrossRefGoogle Scholar
Tardieu, F., Katerji, N., Bethenod, O., et al. (1991b). Maize stomatal conductance in the field – its relationship with soil and plant water potentials, mechanical constraints and aba concentration in the xylem sap. Plant, Cell and Environment, 14, 121–126.CrossRefGoogle Scholar
Taschler, D. and Neuner, G. (2004). Summer frost resistance and freezing patterns measured in situ in leaves of major alpine plant growth forms in relation to their upper distribution boundary. Plant Cell and Environment, 27, 737–746.CrossRefGoogle Scholar
Tattini, M., Lombardini, L. and Gucci, R. (1997). The effect of NaCl stress and relief on gas exchange properties of two olive cultivars differing in tolerance to salinity. Plant and Soil, 197, 87–93.CrossRefGoogle Scholar
Taub, D.R. (2000). Climate and the US distribution of C4 grass subfamilies and decarboxylation variants of C4 photosynthesis. American Journal of Botany, 87, 1211–1215.CrossRefGoogle Scholar
Taub, D.R., Seemann, J.R. and Coleman, J.S. (2000). Growth in elevated CO2 protects photosynthesis against high-temperature damage. Plant, Cell and Environment, 23, 649–656.CrossRefGoogle Scholar
Tausz, M., Warren, C.R. and Adams, M.A. (2005a). Is the bark of shining gum (Eucalyptus nitens) a sun or shade leaf?Trees, 19, 415–421.CrossRefGoogle Scholar
Tausz, M., Warren, C.R. and Adams, M.A. (2005b). Dynamic light use and protection from excess light in upper canopy and coppice leaves of Nothofagus cunninghamii in an old growth, cool temperate rainforest in Victoria, Australia. New Phytologist, 165, 143–155.CrossRefGoogle Scholar
Taylor, A., Marin, J. and Seel, W.E. (1996). Physiology of the parasitic association between maize and witchweed (Striga hermonthica): is ABA involved. Journal of Experimental Botany, 47, 1057–1065.CrossRefGoogle Scholar
Taylor, G. and Davies, W.J. (1986). Leaf growth of Betula and Acer in simulated shadelight. Oecologia, 69, 589–593.CrossRefGoogle ScholarPubMed
Taylor, S.E. and Terry, N. (1986). Variation in photosynthetic electron transport capacity and its effect on the light modulation of ribulose bisphosphate carboxylase. Photosynthesis Research, 8, 249–256.CrossRefGoogle Scholar
Taylor, S.H., Ripley, B.S., Woodward, F.I. et al. (2011). Drought limitation of photosynthesis differs between C3 and C4 grasses in a comparative experiment. Plant, Cell and Environment, 34, 65–75.CrossRefGoogle Scholar
Taylor, W.C., Rosche, E., Marshall, J.S., et al. (1997). Diverse mechanisms regulate the expression of genes coding for C4 enzymes. Australian Journal of Plant Physiology, 24, 437–442.CrossRefGoogle Scholar
Tazoe, Y., von Caemmerer, S., Badger, M.R., et al. (2009). Light and CO2 do not affect the mesophyll conductance to CO2 diffusion in wheat leaves. Journal of Experimental Botany, 60, 2291–2301.CrossRefGoogle Scholar
Tazoe, Y., von Caemmerer, S., Estavillo, G. et al. (2011). Using tunable diode laser spectroscopy to measure carbon isotope discrimination and mesophyll conductance to CO2 diffusion dynamically at different CO2 concentrations. Plant, Cell and Environment, 34, 580–591.CrossRefGoogle Scholar
Tcherkez, G. and Farquhar, G.D. (2007). On the 16O/18O isotope effect associated with photosynthetic O2 production. Functional Plant Biology, 34, 1049–1052.CrossRefGoogle Scholar
Tcherkez, G. and Farquhar, G.D. (2008). On the effect of heavy water (D2O) on carbon isotope fractionation in photosynthesis. Functional Plant Biology, 35, 201–212.CrossRefGoogle Scholar
Tcherkez, G. and Hodges, M. (2008). How isotopes may help to elucidate primary nitrogen metabolism and its interactions with (photo)respiration in C3 leaves. Journal of Experimental Botany, 59, 1685–1693.CrossRefGoogle Scholar
Tcherkez, G., Cornic, G., Bligny, R., et al. (2005). In vivo respiratory metabolism of illuminated leaves. Plant Physiology, 138, 1596–1606.CrossRefGoogle ScholarPubMed
Tcherkez, G., Cornic, G., Bligny, R., et al. (2008). Respiratory metabolism of illuminated leaves depends on CO2 and O2 conditions. Proceedings of the Natural Academy of Sciences USA, 105, 797–802.CrossRefGoogle ScholarPubMed
Tcherkez, G., Farquhar, G.D., Badek, F., et al. (2004). Theoretical considerations about carbon isotope distribution in glucose of C3 plants. Functional Plant Biology, 31, 857–877.CrossRefGoogle Scholar
Tcherkez, G.G.B., Farquhar, G.D. and Andrews, T.J. (2006). Despite slow catalysis and confused substrate specificity, all ribulose bisphosphate carboxylases may be nearly perfectly optimized. Proceedings of the National Academy of Sciences, 103, 7246–7251.CrossRefGoogle ScholarPubMed
Técsi, L.I., Smith, A.M., Maule, A.J., et al. (1996). A spatial analysis of physiological changes associated with infection of cotyledons of marrow plants with Cucumber mosaic virus. Plant Physiology, 111, 975–985.CrossRefGoogle ScholarPubMed
Tedeschi, V., Rey, A., Manca, G., et al. (2006). Soil respiration in a Mediterranean oak forest at different developmental stages after coppicing. Global Change Biology, 12, 110–121.CrossRefGoogle Scholar
Teklemariam, T.A. and Sparks, J.P. (2004). Gaseous fluxes of peroxyacetyl nitrate (PAN) into plant leaves. Plant, Cell and Environment, 27, 1149–1158.CrossRefGoogle Scholar
Telfer, A. (2005). Too much light? How beta-carotene protects the photosystem II reaction centre. Photochemical and Photobiological Sciences, 4, 950–956.CrossRefGoogle ScholarPubMed
Telfer, A., Bottin, H., Barber, J., et al. (1984). The effect of magnesium and phosphorylation of light-harvesting chlorophyll a/b-protein on the yield of P-700-photooxidation in pea chloroplasts. Biochimica et Biophysica Acta, 764, 324–330.CrossRefGoogle Scholar
Tenhunen, J.D., Pearcy, R.W. and Lange, O.L. (1987). Diurnal variations in leaf conductance and gas exchange in natural environments. In: Stomatal Function (eds Zeiger, E., Farquhar, G.D. and Cowan, I.R.), Stanford University Press, Stanford, USA, pp. 323–352.Google Scholar
Tenhunen, J.D., Falge, E., Ryel, R., et al. (2001). Modelling of fluxes in a spruce forest catchment of the Fichtelgebirge. In: Ecosystem Approaches to Landscape Management in Central Europe (eds Tenhunen, J.D., Lenz, R. and Hantschel, R.), Springer-Verlag, Berlin, Germany, pp. 417–462.CrossRefGoogle Scholar
Tenhunen, J.D., Hanano, R., Abril, M., et al. (1994). Above- and below-ground environmental influences on leaf conductance of Ceanothus thyrsiflorus growing in a chaparral environment: drought response and the role of abscisic acid. Oecologia, 99, 306–314.CrossRefGoogle Scholar
Tenhunen, J.D., Lange, O.L., Gebel, J., et al. (1984). Changes in photosynthetic capacity, carboxylation efficiency, and CO2 compensation point associated with midday stomatal closure and midday depression of net CO2 exchange of leaves of Quercus suber. Planta, 162, 193–203CrossRefGoogle ScholarPubMed
Tenhunen, J.D., Sala Serra, A., Harley, P.C., et al. (1990). Factors influencing carbon fixation and water use by mediterranean sclerophyll shrubs during summer drought. Oecologia, 82, 381–393.CrossRefGoogle ScholarPubMed
Terahima, I. and Ono, K. (2002). Effects of HgCl2 on CO2 dependence of leaf photosynthesis: evidence indicating involvement of aquaporins in CO2 diffussion across the plasma membrane. Plant Cell Physiology, 43, 70–78.CrossRefGoogle Scholar
Terashima, I. (1992). Anatomy of non-uniform leaf photosynthesis. Photosynthesis Research, 31, 195–212.CrossRefGoogle ScholarPubMed
Terashima, I. and Hikosaka, K. (1995). Comparative ecophysiology of leaf and canopy photosynthesis. Plant Cell Environment, 18, 1111–1128.CrossRefGoogle Scholar
Terashima, I. and Inoue, Y. (1985a). Palisade tissue chloroplasts and spongy tissue chloroplasts in spinach: Biochemical and ultrastructural differences. Plant Cell Physiology, 26, 63–75.Google Scholar
Terashima, I. and Inoue, Y. (1985b). Vertical gradient in photosynthetic properties of spinach chloroplasts dependent on intra-leaf light environment. Plant Cell Physiology, 26, 781–785.CrossRefGoogle Scholar
Terashima, I. and Ono, K. (2002). Effects of HgCl2 on CO2 dependence of leaf photosynthesis: evidence indicating involvement of aquaporins in CO2 diffusion across the plasma membrane. Plant and Cell Physiology, 43, 70–78.CrossRefGoogle Scholar
Terashima, I., Araya, T., Miyazawa, S.I., et al. (2005), Construction and maintenance of the optimal photosynthetic systems of the leaf, herbaceous plant and tree: an eco-developmental treatise. Annals of Botany, 95, 507–519.CrossRefGoogle ScholarPubMed
Terashima, I. and Evans, J.R. (1988). Effects of light and nitrogen nutrition on the organization of the photosynthetic apparatus in spinach. Plant and Cell Physiology, 29, 143–155.Google Scholar
Terashima, I., Fujita, T., Inoue, T., et al. (2009). Green light drives leaf photosynthesis more efficiently than red light in strong white light: revisiting the enigmatic question of why leaves are green. Plant Cell Physiology, 50, 684–697.CrossRefGoogle ScholarPubMed
Terashima, I., Hanba, Y.T., Tholen, D. et al. (2011). Leaf functional anatomy in relation to photosynthesis. Plant Physiology, 155, 108–116.CrossRefGoogle ScholarPubMed
Terashima, I., Masuzawa, T., Ohba, H., et al. (1995). Is photosynthesis suppressed at higher elevations due to low CO2 pressure?Ecology, 76, 2663–2668.CrossRefGoogle Scholar
Terashima, I., Noguchi, K., Itoh-Nemoto, T., et al. (1998). The cause of PSI photoinhibition at low temperatures in leaves of Cucumis sativus, a chilling-sensitive plant. Physiologia Plantarum, 103, 295–303.CrossRefGoogle Scholar
Terashima, I., Sakaguchi, S. and Hara, N. (1986). Intra-leaf and intracellular gradients in chloroplast ultrastructure of dorsiventral leaves illuminated from the adaxial or abaxial side during their development. Plant Cell Physiology, 27, 1023–1031.Google Scholar
Terjung, F. (1998). Reabsorption of chlorophyll fluorescence and its effects on the spectral distribution and the picosecond decay of higher plant leaves. Z Naturforsch C, 53, 924–926.Google Scholar
Terry, N. (1980). Limiting factors in photosynthesis I. Use of iron stress to control photochemical capacity in vivo. Plant Physiology, 65, 114–120.CrossRefGoogle ScholarPubMed
Terry, N. and Abadía, J. (1986). Function of iron in chloroplasts. Journal of Plant Nutrition, 9, 609–646.CrossRefGoogle Scholar
Terry, N. and Ulrich, A. (1973a). Effects of phosphorus deficiency on the photosynthesis and respiration of leaves of sugar beet. Plant Physiology, 51, 43–47.CrossRefGoogle ScholarPubMed
Terry, N. and Ulrich, A. (1973b). Effects of potassium deficiency on the photosynthesis and respiration of leaves of sugar beet. Plant Physiology, 51, 783–786.CrossRefGoogle ScholarPubMed
Terry, N. and Ulrich, A. (1974). Photosynthetic and respiratory CO2 exchange of sugar beet as influenced by manganese deficiency. Crop Science, 14, 502–504.CrossRefGoogle Scholar
Terzaghi, W.B., Fork, D.C., Berry, J.A., et al. (1989). Low and high temperature limits to PSII. Plant Physiology, 91, 1494–1500.CrossRefGoogle ScholarPubMed
Teskey, R.O. (1997). Combined effects of elevated CO2 and air temperature on carbon assimilation of Pinus taeda trees. Plant, Cell and Environment, 20, 373–380.CrossRefGoogle Scholar
Teskey, R.O., Grier, C.C. and Hinckley, T.M. (1984). Change in photosynthesis and water relations with age and season in Abies amabilis. Canadian Journal of Forest Research, 14, 77–84.CrossRefGoogle Scholar
Teskey, R.O., Saveyn, A., Steppe, K., et al. (2008). Origin, fate and significance of CO2 in tree stems. New Phytologist, 177, 17–32.Google ScholarPubMed
Tew, R.K. (1970). Seasonal variation in the nutrient content of aspen foliage. Journal of Wildlife Management, 34, 475–478.CrossRefGoogle Scholar
Tezara, W., Mitchell, V.J., Driscoll, S.D., et al. (1999). Water stress inhibits plant photosynthesis by decreasing coupling factor and ATP. Nature, 401, 914–917.CrossRefGoogle Scholar
Thiele, A., Krause, G.H. and Winter, K. (1998). In situ study of photoinhibition of photosynthesis and xanthophyll cycle activity in plants growing in natural gaps of the tropical forest. Australian Journal of Plant Physiology, 25, 189–195.CrossRefGoogle Scholar
Tholen, D. and Zhu, X.-G. (2011). The mechanistic basis of internal conductance: a theoretical analysis of mesophyll cell photosynthesis and CO2 diffusion. Plant Physiology, 156, 90–105.CrossRefGoogle ScholarPubMed
Tholen, D., Boom, C., Noguchi, K., et al. (2008). The chloroplast avoidance response decreases internal conductance to CO2 diffusion in Arabidopsis thaliana leaves. Plant Cell and Environment, 31, 1688–1700.CrossRefGoogle ScholarPubMed
Tholen, D., Pons, T.L., Voesenek, L.A.C.J., et al. (2007). Ethylene insensitivity results in down-regulation of Rubisco expression and photosynthetic capacity in tobacco. Plant Physiology, 144, 1305–1315.CrossRefGoogle ScholarPubMed
Tholl, D., Boland, W., Hansel, A., et al. (2006). ‘Practical approaches to plant volatile analysis’. The Plant Journal, 45(4), 540–560.CrossRefGoogle Scholar
Thomas, D.S., Montagu, K.D. and Conroy, J.P. (2006). Leaf inorganic phosphorus as a potential indicator of phosphorus status, photosynthesis and growth of Eucalyptus grandis seedlings. Forest Ecology and Management, 223, 267–274.CrossRefGoogle Scholar
Thomas, H., Ougham, H.J., Wagstaff, C., et al. (2003). Defining senescence and death. Jounral of Experimental Botany, 385, 1127–1132.CrossRefGoogle Scholar
Thomas, R.B., Reid, C.D., Ybema, R., et al. (1993). Growth and maintenance components of leaf respiration of cotton grown in elevated carbon dioxide partial pressure. Plant Cell Environment, 16, 539–546.CrossRefGoogle Scholar
Thomas, S.C. and Winner, W.E. (2002). Photosynthetic differences between saplings and adult trees: an integration of field results by meta-analysis. Tree Physiology, 22, 117–127.CrossRefGoogle ScholarPubMed
Thomashow, M.F. (1999). Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annual Review of Plant Physiology and Plant Molecular Biology, 50, 571–599.CrossRefGoogle ScholarPubMed
Thompson, L.K., Blaylock, R., Sturtevant, J.M., et al. (1989). Molecular basis of the heat denaturation of photosystem II. Biochemistry, 28, 6686–6695.CrossRefGoogle ScholarPubMed
Thomson, V.P., Cunningham, S.A., Ball, M.C., et al. (2003). Compensation for herbivory by Cucumis sativus through increased photosynthetic capacity and efficiency. Oecologia, 134, 167–175.CrossRefGoogle Scholar
Thornley, J.M. and Johnson, I.R. (1990). Plant and crop modelling. A mathematical approach to plant and crop physiology. Clarendon Press, Oxford, UK.Google Scholar
Tieszen, L.L., Reed, B.C., Bliss, N.B., et al. (1997). NDVI, C3 and C4 production, and distributions in Great Plains grassland land cover classes. Ecological Applications, 7, 59–78.Google Scholar
Tikhonov, A.N., Khomutov, G.B. and Ruuge, E.K. (1984). Electron transport control in chloroplasts. Effects of magnesium ions on the electron flow between two photosystems. Photobiochemistry and Photobiology, 8, 261–269.Google Scholar
Tikkanen, M., Grieco, M. and Aro, E.-M. (2011). Novel insights into plant light-harvesting complex II phosphorylation and ‘state transitions’. Trends in Plant Science, 16, 126–131.CrossRefGoogle ScholarPubMed
Tikkanen, M., Nurmi, M., Suorsa, M., et al. (2008). Phosphorylation-dependent regulation of excitation energy distribution between the two photosystems in higher plants. Biochimica et Biophysica Acta, 1777, 425–432.CrossRefGoogle ScholarPubMed
Tilman, D. and Wedin, D. (1991). Plant traits and resource reduction for five grasses growing on a nitrogen gradient. Ecology, 72, 685–700.CrossRefGoogle Scholar
Tilman, D., Cassman, K.G., Matson, P.A., et al.. (2002) Agricultural sustainability and intensive production practices. Nature, 418, 671–677.CrossRefGoogle ScholarPubMed
Timlin, D., Rahman, S.M.L., Baker, J., et al. (2006). Whole plant photosynthesis, development, and carbon partitioning in potato as a function of temperature. Agronomy Journal, 98, 1195–1203.CrossRefGoogle Scholar
Timperio, A.M., D’Amici, G.M., Barta, C., et al. (2007). Proteomics, pigment composition, and organization of thylakoid membranes in iron-deficient spinach leaves. Journal of Experimental Botany, 58, 3695–3710.CrossRefGoogle ScholarPubMed
Tinoco-Ojanguren, C. and Pearcy, R.W. (1992). Dynamic stomatal behavior and its role in carbon gain during lightflecks of a gap phase and an understory Piper species acclimated to high and low light. Oecologia, 92, 222–228.CrossRefGoogle ScholarPubMed
Tinoco-Ojanguren, C. and Pearcy, R.W. (1993a). Stomatal dynamics and its importance to carbon gain in two rainforest Piper species. II. Stomatal versus biochemical limitations during photosynthetic induction. Oecologia, 94, 395–402.CrossRefGoogle ScholarPubMed
Tipple, B.J. and Pagani, M. (2007). The early origins of terrestrial C4 photosynthesis. Annual Review of Earth Planetary Science, 35, 435–461.CrossRefGoogle Scholar
Tissue, D.T., Griffin, K.L., Turnbull, M.H., et al. (2005). Stomatal and non-stomatal limitations to photosynthesis in four tree species in a temperate rainforest dominated by Dacrydium cupressinum in New Zealand. Tree Physiology, 25, 447–456.CrossRefGoogle Scholar
Titel, C., Woehlecke, H., Afifi, I., et al. (1997). Dynamics of limiting cell wall porosity in plant suspension cultures. Planta, 203, 320–326.CrossRefGoogle Scholar
Tjoelker, M.G., Oleksyn, J. and Reich, P.B. (1999). Acclimation of respiration to temperature and CO2 in seedlings of boreal tree species in relation to plant size and relative growth rate. Global Change Biology, 5, 679–691.CrossRefGoogle Scholar
Tjoelker, M.G., Oleksyn, J. and Reich, P.B. (2001). Modelling respiration of vegetation: evidence for a general temperature dependent Q10. Global Change Biology, 7, 223–230.CrossRefGoogle Scholar
Tocquin, P. and Périlleux, C. (2004). Design of a versatile device for measuring whole plant gas exchanges in Arabidopsis thaliana. New Phytologist, 162, 223–229.CrossRefGoogle Scholar
Toft, N. and Pearcy, R.W. (1982). Gas exchange characteristics and temperature relations of two desert annuals: a comparison of a winter-active and a summer-active species. Oecologia, 55, 170–7.CrossRefGoogle Scholar
Tognetti, R., Johnson, J.D. and Michelozzi, M. (1995). The response of european beech (Fagus sylvatica L) seedlings from 2 italian populations to drought and recovery. Trees, Structure and Function, 9, 348–354.CrossRefGoogle Scholar
Tognetti, R., Longobucco, A., Miglietta, F., et al. (1999). Water relations, stomatal response and transpiration of Quercus pubescens trees during summer in a Mediterranean carbon dioxide spring. Tree Physiology, 25, 261–270.CrossRefGoogle Scholar
Tognetti, R., Minnocci, A., Peñuelas, J., et al. (2000). Comparative field water relations of three Mediterranean shrub species co-occurring at a natural CO2 vent. Journal of Experimental Botany, 51, 1131–1146.CrossRefGoogle Scholar
Tollenaar, M. and Aguilera, A. (1992). Radiation use efficiency of an old and a new maize hybrid. Agronomy Journal, 84, 536–41.CrossRefGoogle Scholar
Tollenaar, M.,. McCullough, D.E. and Dwyer, L.M. (1994). Physiological basis of the genetic improvement in corn. In: Genetic Improvement of Field Crops (ed. Slafer, G.A.), Marcel Dekker, New York, USA, pp. 183–236.Google Scholar
Tomescu, A.M.F. (2008). Megaphylls, microphylls and the evolution of leaf development. Trends in Plant Science, 14, 5–12.CrossRefGoogle ScholarPubMed
Tominaga, M., Kinoshita, T. and Shimazaki, K. (2001). Guard cell chloroplasts provide ATP required for H+ pumping in the plasma membrane and stomatal opening. Plant and Cell Physiology, 42, 795–802.CrossRefGoogle ScholarPubMed
Tomlinson, J.A. and Webb, M.J.W. (1978). Ultrastructural changes in chloroplasts of lettuce infected with beet western yellows virus. Physiological Plant Pathology, 12, 13–18.CrossRefGoogle Scholar
Tong, H. and Hipps, L.E. (1996). The effect of turbulence on the light environment of alfalfa. Agricultural and Forest Meteorology, 80, 249–261.CrossRefGoogle Scholar
Törnroth-Horsefield, S., Wang, Y., Hedfalk, K., et al. (2006). Structural mechanism of plant aquaporin gating. Nature, 439, 688–694.CrossRefGoogle ScholarPubMed
Török, Z., Tsvetkova, N. M., Balogh, G., et al. (2003). Heat shock protein coinducers with no effect on protein denaturation specifically modulate the membrane lipid phase. Proceedings of the National Academy of Sciences of the United States of America, 100, 3131–3136.CrossRefGoogle ScholarPubMed
Tortell, P.D. (2000). Evolutionary and ecological perspectives on carbon acquisition in phytoplankton. Limnology and Oceanography, 45, 744–750.CrossRefGoogle Scholar
Tóth, S.T., Schansker, G., Garab, G., et al. (2007). Photosynthetic electron transport activity in heat-treated barley leaves: the role of internal alternative electron donors to photosystem II. Biochimica et Biophysica Acta, 1767, 295–305.CrossRefGoogle ScholarPubMed
Tóth, S.Z., Puthur, J.T., Nagy, V., et al. (2009). Experimental evidence for ascorbate-dependent electron transport in leaves with inactive oxygen-evolving complexes. Plant Physiology, 149, 1568–1578.CrossRefGoogle ScholarPubMed
Tournaire-Roux, C., Sutka, M., Javot, H., et al. (2003). Cytosolic pH regulates root water transport during anoxic stress through gating of aquaporins. Nature, 425, 393–397.CrossRefGoogle ScholarPubMed
Tournebize, R. and Sinoquet, H. (1995). Light interception and partitioning in a shrub/grass mixture. Agricultural and Forest Meteorology, 72, 277–294.CrossRefGoogle Scholar
Tovar-Mendez, A., Miernyk, J.A. and Randall, D.D. (2003). Regulation of pyruvate dehydrogenase complex activity in plant cells. European Journal of Biochemistry, 270, 1043–1049.CrossRefGoogle ScholarPubMed
Townsend, A.R., Asner, G.P. and White, J.W.C. (2002). Land use effects of atmosphere 13C imply a sizable terrestrial CO2 sink in tropical latitudes. Geophysical Research Letters, 29, 1426.CrossRefGoogle Scholar
Tranquilini, W. (1957). Standortsklima, Wasserbilanz un CO2-gaswechsel junger Zirben (Pinus cembra L.) an de alpinen Waldgrenze. Planta, 49, 612–661.CrossRefGoogle Scholar
Trejo, C.L. and Davies, W.J. (1991). Drought-induced closure of Phaseolus vulgaris L. stomata precedes leaf water deficit and any increase in xylem ABA concentration. Journal of Experimental Botany, 42, 1507–1515CrossRefGoogle Scholar
Trenberth, K.E. and Jones, P.D. (2007). Observations: Surface and atmospheric climate change. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (eds Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M. and Miller, H.L.), Cambridge University Press, Cambridge, UK, pp. 235–336.Google Scholar
Triantaphylidès, C., Krischke, M., Hoeberichts, F.A., et al. (2008). Singlet oxygen is the major reactive oxygen species involved in photooxidative damage to plants. Plant Physiology, 148, 960–968.CrossRefGoogle ScholarPubMed
Triggs, J., Kimball, B., Pinter, P., et al. (2004). Free-air CO2 enrichment effects on the energy balance and evapotranspiration of sorghum. Agricultural and Forest Meteorology, 124, 63–79.CrossRefGoogle Scholar
Troughton, J.H. and Slatyer, R.O. (1969). Plant water status, leaf temperature, and the calculated mesophyll resistance to carbon dioxide of cotton leaves. Australian Journal of Biological Sciences, 22, 815–827.CrossRefGoogle Scholar
Truman, W., Torres de Zabala, M. and Grant, M. (2006). Type III effectors orchestrate a complex interplay between transcriptional networks to modify basal defence responses during pathogenesis and resistance. The Plant Journal, 46, 14–33.CrossRefGoogle ScholarPubMed
Trumble, J.T., Kolodny Hirsch, D.M. and Ting, I.P. (1993). Plant compensation for arthropod herbivory. Annual Reviews of Entomology, 38, 93–119.CrossRefGoogle Scholar
Tseng, M.J., Liu, C.W. and Yiu, J.C. (2007). Enhanced tolerance to sulfur dioxide and salt stress of transgenic Chinese cabbage plants expressing both superoxide dismutase and catalase in chloroplasts. Plant Physiology and Biochemistry, 45, 822–833.CrossRefGoogle ScholarPubMed
Tsialtas, J.T., Handley, L.L., Kassioumi, M.T., et al. (2001). Interspecific variation in potential water-use efficiency and its relation to plant species abundance in a water-limited grassland. Functional Ecology, 15, 605–614.CrossRefGoogle Scholar
Tsuchihira, A., Hanba, Y.T., Kato, N., et al. (2010). Effect of overexpression of radish plasma membrane aquaporins on water-use efficiency, photosynthesis and growth of Eucalyptus trees. Tree Physiology, 30, 417–430.CrossRefGoogle ScholarPubMed
Tsukaya, H., Okada, H. and Mohamed, M. (2004). A novel feature of structural variegation in leaves of the tropical plant Schismatoglottis calyptrate. Journal of Plant Research, 117, 477–480.CrossRefGoogle Scholar
Tsuyama, M., Shibata, M. and Kobayashi, Y. (2003). Leaf factors affecting the relationship between chlorophyll fluorescence and the rate of photosynthetic electron transport as determined from CO2 uptake. Journal of Plant Physiology, 160, 1131–1139.CrossRefGoogle ScholarPubMed
Tsvetkova, N.M., Horváth, I., Török, Z., et al. (2002). Small heat-shock proteins regulate membrane lipid polymorphism. Proceedings of the National Academy of Sciences of the United States of America, 99, 13504–13509.CrossRefGoogle ScholarPubMed
Tu, J.C. and Ford, R.E. (1968). Effect of maize dwarf mosaic virus infection on respiration and photosynthesis of corn. Phytopathology, 58, 282–284.Google Scholar
Tucker, C.J., Townshend, J.R.G. and Goff, T.E. (1985). African landcover classification using satellite data. Science, 227, 369–374.CrossRefGoogle Scholar
Turgeon, R. and Wolf, S. (2009). Phloem transport: Cellular pathways and molecular trafficking. Annual Review of Plant Biology, 60, 207–221.CrossRefGoogle ScholarPubMed
Turgut, R. and Kadioglu, A., (1998). The effect of drought, temperature and irradiation on leaf rolling in Ctenanthe setosa. Biologia Plantarum, 41, 629–633.CrossRefGoogle Scholar
Turnbull, T.L., Adams, M.A. and Warren, C.R. (2007a). Increased photosynthesis following partial defoliation of field-grown Eucalyptus globulus is not caused by increased leaf nitrogen. Tree Physiology, 27, 1481–1492.CrossRefGoogle Scholar
Turnbull, TL., Warren, C.R. and Adams, M.A. (2007b). Novel mannose-sequestration technique reveals variation in subcellular orthophosphate pools do not explain the effects of phosphorus nutrition on photosynthesis in Eucalyptus globulus seedlings. New Phytologist, 176, 849–861.CrossRefGoogle Scholar
Turner, I.M. (1994). Sclerophylly: primarily protective?Functional Ecology, 8, 669–675.CrossRefGoogle Scholar
Turner, N.C. (2004). Agronomic options for improving rainfall-use efficiency of crops in dryland farming systems. Journal of Experimental Botany, 55, 2413–2425.CrossRefGoogle ScholarPubMed
Tyree, M.T. and Sperry, J.S. (1989). Vulnerability of xylem to cavitation and embolism. Annual Review of Plant Physiology and Plant Molecular Biology, 40, 19–38CrossRefGoogle Scholar
Tyree, M.T. and Wilmot, T.R. (1990). Errors in the calculation of evaporation and leaf conductance in steady state porometry: the importance of accurate measurement of leaf temperature. Canadian Journal of Forestry Research, 20, 1031–1035.CrossRefGoogle Scholar
Tyree, M.T. and Zimmermann, M.H. (2002). Xylem Structure and Ascent of Sap. Springer-Verlag, Berlin, Germany, p. 283.CrossRefGoogle Scholar
Tyystjärvi, E. and Vass, I. (2004). Light emission as a probe of charge separation and recombination in the photosynthetic apparatus: relation of prompt fluorescence to delayed light emission and thermoluminescence. In: Chlorophyll a Fluorescence: A Signature of Photosynthesis. Advances in photosynthesis and respiration, vol 19 (eds. Papageorgiou, G.C. and Govindjee, J.), Kluwer Academic Publishers, Dordrecht, Nederlands, pp. 363–388.CrossRef
Tyystjärvi, E., Koski, A., Keranen, M., et al. (1999). The Kautsky curve is a built-in barcode. Biophysical Journal, 77, 1159–1167.CrossRefGoogle ScholarPubMed
Tzvetkova-Chevolleau, T., Franck, T., Alawady, A.E., et al. (2007). The light stress-induced protein ELIP2 is a regulator of chlorophyll synthesis in Arabidopsis thaliana. The Plant Journal, 50, 795–809.CrossRefGoogle ScholarPubMed
U.S. DOE. (2006). Breaking the biological barriers to cellulosic ethanol: A joint research agenda, DOE/SC/EE-0095, US Department of Energy Office of Science and Office of Energy Efficiency and Renewable Energy, .
Ubierna, N. and Marshall, J.D. (2011). Estimation of canopy average mesophyll conductance using δ13C of phloem contents. Plant, Cell and Environment, 34, 1521–1535.CrossRef
Uedan, K. and Sugiyama, T. (1976). Purification and characterization of phosphoenolpyruvate carboxylase from maize leaves. Plant Physiology, 57, 906–910.CrossRefGoogle ScholarPubMed
Uehlein, N., Lovisolo, C., Siefritz, F., et al. (2003). The tobacco aquaporin NtAQP1 is a membrane CO2 transporter with physiological functions. Nature, 425, 734–737.CrossRefGoogle ScholarPubMed
Uehlein, N., Otto, B., Hanson, D.T., et al. (2008). Function of Nicotiana tabacum aquaporins as chloroplast gas pores challenges the concept of membrane CO2 permeability. Plant Cell, 20, 648–657.CrossRefGoogle ScholarPubMed
Uemura, A., Ishida, A., Nakano, T., et al. (2000). Acclimation of leaf characteristics of Fagus species to previous-year and current-year solar irradiances. Tree Physiology, 20, 945–951.CrossRefGoogle ScholarPubMed
Uemura, K., Anwaruzzaman, M.S. and Yokota, A. (1997). Ribulose-1,5-bisphosphate carboxylase/oxygenase from thermophilic red algae with a strong specificity for CO2 fixation. Biochemical and Biophysical Research Communications, 233, 568–571.CrossRefGoogle ScholarPubMed
Uemura, K., Suzuki, Y., Shikani, T., et al. (1996). A rapid and sensitive method for determination of relative specificity of RuBisCO from various species by anion-exchange chromatography. Plant and Cell Physiology, 37, 325–331.CrossRefGoogle Scholar
Uhl, C., Clark, K., Dezzeo, N., et al. (1988).Vegetation dynamics in amazonian treefall gaps. Ecology, 69, 751–763.CrossRefGoogle Scholar
UNIS, (2000). Secretary General address to Developing Countries “South Summit”, UN Information Service Press Release, 13 April 2000. (See .)
Upchurch, Jr., G.R. (1984). Cuticular anatomy of angiosperm leaves from the Lower Cretaceous Potomac group. I. Zone I leaves. American Journal of Botany, 71, 192–202.CrossRefGoogle Scholar
Urban, L., Jegouzo, L., Damour, G., et al. (2008). Interpreting the decrease in leaf photosynthesis during flowering in mango. Tree Physiology, 28, 1025–1036.CrossRefGoogle ScholarPubMed
Urbanczyk-Wochniak, E., Baxter, C., Kolbe, A., et al. (2005). Profiling of diurnal patterns of metabolite and transcript abundance in potato (Solanum tuberosum) leaves. Planta, 221, 891–903.CrossRefGoogle ScholarPubMed
Urey, H.C. (1947). The thermodynamic properties of isotopic substances. Journal of Chemical Society, 1947, 562–581.CrossRefGoogle Scholar
Uribe, E.G. and Stark, B. (1982). Inhibition of photosynthetic energy conversion by cupric ion. Evidence for Cu2+-coupling factor 1 interaction. Plant Physiology, 69, 1040–1045.CrossRefGoogle ScholarPubMed
Ustin, S.L., Roberts, D.R., Gamon, J.A., et al. (2004). Using imaging spectroscopy to study ecosystem processes and properties. BioScience, 54, 523–534.CrossRefGoogle Scholar
Usuda, H. (2004). Evaluation of the effect of photosynthesis on biomass production with simultaneous analysis of growth and continuous monitoring of CO2 exchange in the whole plants of radish cv Kosena under ambient and elevated CO2. Plant Production Science, 7, 386–396.CrossRefGoogle Scholar
Vahisalu, T., Kollist, H., Wang, Y.F., et al. (2008). SLAC1 is required for plant guard cell S- type anion channel function in stomatal signalling. Nature, 452, 487–491.CrossRefGoogle Scholar
Vaillant, N., Monnet, F., Hitmi, A., et al. (2005). Comparative study of responses in four Datura species to a zinc stress. Chemosphere, 59, 1005–1013.CrossRefGoogle ScholarPubMed
Val, J., Sanz, M., Montañés, L., et al. (1995). Application of chlorophyll fluorescence to study iron and manganese deficiencies in peach tree. Acta Horticulturae, 383, 201–209.CrossRefGoogle Scholar
Valcke, R. (2003). Fluorescence imaging: the stethoscope of the plant physiologist. Advances in Plant Physiology, 6, 445–462.Google Scholar
Valente, M.A.S., Faria, J.A.Q.A., Soares-Ramos, J.R.L., et al. (2009). The ER luminal binding protein (BiP) mediates an increase in drought tolerance in soybean and delays drought-induced leaf senescence in soybean and tobacco. Journal of Experimental Botany, 60, 533–546.CrossRefGoogle ScholarPubMed
Valentini, R. (ed.) (2003). Fluxes of Carbon, Water and Energy of European Forests, Ecological Studies 163, Springer-Verlag, Berlin Heidelberg, Germany, pp. 1–450.CrossRefGoogle Scholar
Valentini, R., Cecchi, G., Mazzinghi, P., et al. (1994). Remote sensing of chlorophyll a fluorescence of vegetation canopies: 2. Physiological significance of fluorescence signal in response to environmental stresses. Remote Sensing of the Environment, 47, 29–35.CrossRefGoogle Scholar
Valentini, R., Deangelis, P., Matteucci, G., et al. (1996). Seasonal net carbon dioxide exchange of a beech forest with the atmosphere. Global Change Biology, 2, 199–207.CrossRefGoogle Scholar
Valentini, R., Epron, D., De Angelis, P., et al. (1995a). In situ estimation of net CO2 assimilation, photosynthetic electron flow and photorespiration in Turkey oak (Quercus cerris L.) leaves: diurnal cycles under different levels of water supply. Plant Cell Environment, 18, 631–640.CrossRefGoogle Scholar
Valentini, R., Gamon, J.A. and Field, C.B. (1995b). Ecosystem gas exchange in a California grassland: seasonal patterns and implications for scaling. Ecology, 76, 1940–1952.CrossRefGoogle Scholar
Valentini, R., Matteucci, G., Dolman, A.J., et al. (2000). Respiration as the main determinant of carbon balance in European forests. Nature, 404, 861–865.CrossRefGoogle ScholarPubMed
Valentini, R., Scarascia Mugnozza, G.E., De Angelis, P., et al. (1991). An experimental test of the eddy correlation technique over a Mediterranean macchia canopy. Plant Cell Environment, 14, 987–994.CrossRefGoogle Scholar
Valiela, I., Bowen, J.L. and York, J.K. (2001) Mangrove forests: one of the world’s threatened major tropical environments. BioScience, 51, 807–815.CrossRefGoogle Scholar
Valjakka, M., Luomala, E.M., Kangasjärvi, J., et al. (1999). Expression of photosynthesis- and senescence-related genes during leaf development and senescence in silver birch (Betula pendula) seedlings. Physiologia Plantarum, 106, 302–310.CrossRefGoogle Scholar
Valladares, F. (2003). Light heterogeneity and plants: from ecophysiology to species coexistence and biodiversity. In: Progress in Botany, vol 64 (eds Esser, K., Lüttge, U., Beyschlag, W. and Hellwig, F.), Springer Verlag, Heidelberg, Germany, pp. 439–471.CrossRef
Valladares, F. and Niinemets, Ü. (2007). The architecture of plant crowns: from design rules to light capture and performance. In: Handbook of Functional Plant Ecology (eds Pugnaire, F. I. and Valladares, F.), CRC Press, Boca Raton, Florida, USA, pp. 101–149.CrossRefGoogle Scholar
Valladares, F. and Niinemets, Ü. (2008). Shade tolerance, a key plant feature of complex nature and consequences. Annual Review of Ecology, Evolution and Systematics 39, 237–257.CrossRefGoogle Scholar
Valladares, F. and Pearcy, R.W. (1997). Interactions between water stress, sun-shade acclimation, heat tolerance and photoinhibition in the sclerophyll Heteromeles arbutifolia. Plant, Cell and Environment, 20, 25–36.CrossRefGoogle Scholar
Valladares, F. and Pearcy, R.W. (1998). The functional ecology of shoot architecture in sun and shade plants of Heteromeles arbutifolia M. Roem., a Californian chaparral shrub. Oecologia, 114, 1–10.CrossRefGoogle Scholar
Valladares, F. and Pearcy, R.W. (1999). The geometry of light interception by shoots of Heteromeles arbutifolia: morphological and physiological consequences for individual leaves. Oecologia, 121, 171–182.CrossRefGoogle ScholarPubMed
Valladares, F. and Pearcy, R.W. (2000). The role of crown architecture for light harvesting and carbon gain in extreme light environments assessed with a realistic 3-D model. Anales Jardin Botanico de Madrid, 58, 3–16.Google Scholar
Valladares, F. and Pugnaire, F.I. (1999). Tradeoffs between irradiance capture and avoidance in semi-arid environments assessed with a crown architecture model. Annals of Botany, 83, 459–469.CrossRefGoogle Scholar
Valladares, F., Allen, M.T. and Pearcy, R.W. (1997). Photosynthetic response to dynamic light under field conditions in six tropical rainforest shrubs occurring along a light gradient. Oecologia, 111, 505–514.CrossRefGoogle Scholar
Valladares, F., Dobarro, I., Sánchez-Gómez, D., et al. (2005). Photoinhibition and drought in Mediterranean woody saplings: scaling effects and interactions in sun and shade phenotypes. Journal of Experimental Botany, 56, 483–494.CrossRefGoogle Scholar
Valladares, F., Gianoli, E. and Gómez, J.M. (2007). Ecological limits to plant phenotypic plasticity. Tansley review. New Phytologist, 176, 749–763.CrossRefGoogle Scholar
Valladares, F., Martinez-Ferri, E., Balaguer, L., et al. (2000a). Low leaf-level response to light and nutrients in Mediterranean evergreen oaks: a conservative resource-use strategy?New Phytologist, 148, 79–91.CrossRefGoogle Scholar
Valladares, F., Sancho, L.G. and Ascaso, C. (1996). Functional analysis of intrathalline and intracellular chlorophyll concentrations in the lichen family Umbilicariaceae. Annals of Botany, 78, 471–477.CrossRefGoogle Scholar
Valladares, F., Skillman, J. and Pearcy, R.W. (2002). Convergence in light capture efficiencies among tropical forest understory plants with contrasting crown architectures: a case of morphological compensation. American Journal of Botany, 89, 1275–1284.CrossRefGoogle ScholarPubMed
Valladares, F., Wright, S.J., Lasso, E., et al. (2000b). Plastic phenotypic response to light of 16 congeneric shrubs from a Panamanian rainforest. Ecology, 81, 1925–1936.CrossRefGoogle Scholar
van Assche, F.V. and Clijsters, H. (1986a). Inhibition of photosynthesis by treatment of Phaseolus vulgaris with toxic concentration of zinc: effects on electron transport and photophosphorylation. Physiologia Plantarum, 66, 717–721.CrossRefGoogle Scholar
van Assche, F.V. and Clijsters, H. (1986b). Inhibition of photosynthesis by treatment of Phaseolus vulgaris with toxic concentration of zinc: effect on ribulose-1,5-bisphosphate carboxylase/oxygenase. Journal of Plant Physiology, 125, 355–360.CrossRefGoogle Scholar
van Assche, F.V. and Clijsters, H. (1990). Effects of metals on enzyme activity in plants. Plant, Cell and Environment, 13, 195–206.CrossRefGoogle Scholar
van Camp, W., Capiau, K., Van Montagu, M., et al. (1996). Enhancement of oxidative stress tolerance in transgenic tobacco plants overproducing Fe-superoxide dismutase in chloroplasts. Plant Physiology, 112, 1703–1714.CrossRefGoogle ScholarPubMed
van der Tol, C., Verhoef, W., Timmermans, J., et al. (2009). An integrated model of soil-canopy spectral radiances, photosynthesis, fluorescence, temperature and energy balance. Biogeosciences, 6, 3109–3129.CrossRef
van der Velde, M., Green, S.R., Vanclooster, M., et al. (2006). Transpiration of squash under a tropical maritime climate. Plant and Soil, 280, 323–337.CrossRefGoogle Scholar
van Doorn, W.G. (2008). Is the onset of senescence in leaf cells of intact plants due to low or high sugar levels?Journal of Experimental Botany, 59, 1963–1972.CrossRefGoogle ScholarPubMed
van Gardingen, P., Grace, J. and Jeffree, C.E. (1991). Abrasive damage by wind to the needles surface of Pinus sylvestris L. and Picea stichensis (Bong.). Plant Cell Environment, 14, 185–193.CrossRefGoogle Scholar
van Gorsel, E., Delpierre, N., Leuning, R., et al. (2009). Estimating nocturnal ecosystem respiration from the vertical turbulent flux and change in storage of CO2. Agricultural and Forest Meteorology, 149, 1919–1930.CrossRefGoogle Scholar
van Gorsel, E., Leuning, R., Cleugh, H.A., et al. (2007). Nocturnal carbon efflux: Reconciliation of eddy covariance and chamber measurements using an alternative to the u*-threshold filtering. technique. Tellus, 59B, 397–403.CrossRefGoogle Scholar
van Gorsel, E., Leuning, R., Cleugh, H.A., et al. (2008). Application of an alternative method to derive reliable estimates of nighttime respiration from eddy covariance measurements in moderately complex topography. Agricultural and Forest Meteorology, 148, 1174–1180.CrossRefGoogle Scholar
van Iersel, M.W. (2003). Carbon use efficiency depends on growth respiration, maintenance respiration, and relative growth rate. A case study with lettuce. Plant Cell Environment, 26, 1441–1446.CrossRefGoogle Scholar
van Iersel, M.W. and Lindstrom, O.M. (1999). Temperature response of whole plant CO2 exchange rates of three magnolia (Magnolia grandiflora L.) cultivars. Journal of the American Society of Horticultural Science, 124, 277–282.Google Scholar
van Kooten, O., Meurs, C. and van Loon, L.C. (1990). Photosynthetic electron transport in tobacco leaves infected with Tobacco mosaic virus. Physiologia Plantarum, 80, 446–452.CrossRefGoogle Scholar
van Oosten, J.J.M., Gerbaud, A., Huijser, C., et al. (1997). An Arabidopsis mutant showing reduced feedback inhibition of photosynthesis. Plant Journal, 12, 1011–1020.CrossRefGoogle ScholarPubMed
van Volkenburgh, E. (1999). Leaf expansion – an integrating plant behaviour. Plant Cell Environment, 22, 1463–1473.CrossRefGoogle Scholar
van Wijk, M.T., Clemmensen, K.E., Shaver, G.R., et al. (2004). Long-term ecosystem level experiments at Toolik Lake, Alaska, and at Abisko, Northern Sweden: Generalizations and differences in ecosystem and plant type responses to global change. Global Change Biology, 10, 105–123.CrossRefGoogle Scholar
van Wijk, M.T., Dekker, S.C., Bouten, W., et al. (2000). Modeling daily gas exchange of Douglas-fir forest: comparison of three stomatal conductance models with and without a soil water stress function. Tree Physiology, 20, 115–122.CrossRefGoogle ScholarPubMed
Vandenbroucke, K., Robbens, S., Vandepoele, K., et al. (2007). Hydrogen peroxide-induced gene expression across kingdoms: a comparative analysis. Molecular Biology and Evolution, 25, 507–516.CrossRefGoogle Scholar
Vandervoet, H. and Mohren, G.M.J. (1994). An uncertainty analysis of the process-based growth-model FORGRO. Forest Ecology and Management, 69, 157–166.CrossRefGoogle Scholar
Vanninen, P., Ylitalo, H., Sievänen, R., et al. (1996). Effects of age and site quality on the distribution of biomass in Scots pine (Pinus sylvestris L.). Trees, 10, 231–238.Google Scholar
Vapaavuori, E.M., Vuorinen, A.J., Aphalo, P.J., et al. (1995). Relationship between net photosynthesis and nitrogen in Scots pine: seasonal variation in seedlings and shoots. Plant and Soil, 168–169, 263–270.CrossRefGoogle Scholar
Vareschi, V. (1980). Vegetationsökologie der Tropen. Stuttgart, Eugen Ulmer.Google Scholar
Vass, I. and Govindjee, J. (1996). Thermoluminescence of the photosynthetic apparatus. Photosynthesis Research, 48, 117–126.CrossRefGoogle ScholarPubMed
Vass, I., Horvath, G., Herczeg, T. and Demeter, S. (1981). Photosynthetic energy conservation investigated by thermoluminescence. activation energies and half-lives of thermoluminescence bands of chloroplasts determined by mathematical resolution of glow curve. Biochimica et Biophysica Acta, 634, 140–152.CrossRefGoogle Scholar
Vaughn, K.C., Ligrone, R., Owen, H.A., et al. (1992). The anthocerote chloroplast :a review. The New Phytologist, 120, 169–190.CrossRefGoogle Scholar
Vavilin, D.V. and Ducruet, J.M. (1998). The origin of 120–130°C thermoluminescence bands in chlorophyll-containing material. Photochemistry Photobiology, 68, 191–198.Google Scholar
Vavilin, D.V., Matorin, D.N., Kafarov, R.S., et al. (1991). High-temperature thermoluminescence of chlorophyll in lipid peroxidation. Biologischke Membranyii, 8, 89–98.Google Scholar
Vavrus, S.J., Walsh, J.E., Chapman, W.L., et al. (2006). The behavior of extreme cold air outbreaks under greenhouse warming. International Journal of Climatology, 26, 1133–1147.CrossRefGoogle Scholar
Vedrova, E.F., Pleshikov, F.I. and Kaplunov, V.Y. (2006). Net ecosystem production of boreal larch ecosystems on the Yenisei transect. Mitigation And Adaptation Strategies For Global Change, 11, 173–190CrossRefGoogle Scholar
Veenendaal, E.M., Kolle, O. and Lloyd, J. (2004) Seasonal variation in energy fluxes and carbon dioxide exchange for a broad-leaved semi-arid savanna (Mopane woodland) in Southern Africa. Global Change Biology, 10, 318–328.CrossRefGoogle Scholar
Veenendaal, E.M., Shushu, D.D. and Scurlock, J.M.O. (1993). Responses to shading of seedlings of savanna grasses (with different C4 photosynthetic pathways) in Botswana. Journal of Tropical Ecology, 9, 213–229.CrossRefGoogle Scholar
Veeranjaneyulu, K., Charland, M., Charlebois, D.C.N., et al. (1991). Photosynthetic energy storage of photosystems I and II in the spectral range of photosynthetically active radiation in intact sugar maple leaves. Photosynthesis Research, 30, 131–138.CrossRefGoogle ScholarPubMed
Velikova, V., Tsonev, T., Barta, C., et al. (2009). BVOC emissions, photosynthetic characteristics and changes in chloroplast ultrastructure of Platanus orientalis L. exposed to elevated CO2 and high temperature. Environmental Pollution, 157, 2629–2637.CrossRefGoogle ScholarPubMed
Vendramini, F., Díaz, S., Gurvich, D.E., et al. (2001). Leaf traits as indicators of resource-use strategy in floras with succulent species. New Phytologist, 154, 147–157.CrossRefGoogle Scholar
Vener, A.V., van Kan, P.J., Rich, P.R., et al. (1997). Plastoquinol at the quinol oxidation site of reduced cytochrome bf mediates signal transduction between light and protein phosphorylation: Thylakoid protein kinase deactivation by a single-turnover flash. Proceedings of the National Academy of Sciences (USA), 94, 1585–1590.CrossRefGoogle ScholarPubMed
Venmos, S.N. and Goldwin, G.K. (1993). Stomatal and chlorophyll distribution of Cox’s orange pippin apple flowers relative to other cluster parts. Annals of Botany, 71, 245–250.CrossRefGoogle Scholar
Verboven, P., Kerckhofs, G., Mebatsion, H.K., et al. (2008). Three-dimensional gas exchange pathways in pome fruit characterized by synchrotron X-ray computed tomography. Plant Physiology, 147, 518–527.CrossRefGoogle ScholarPubMed
Verhoef, A. (1997). The effect of temperature differences between porometer head and leaf surface on stomatal conductance measurements. Plant, Cell and Environment, 20, 641–646.CrossRefGoogle Scholar
Verhoeven, A.S., Adams, W.W., III and Demmig-Adams, B. (1999). The xanthophyll cycle and acclimation of Pinus ponderosa and Malva neglecta to winter stress. Oecologia, 118, 277–287.CrossRefGoogle ScholarPubMed
Verhoeven, A.S., Demmig-Adams, B. and Adams, W.W, III. (1997). Enhanced employment of the xanthophyll cycle and thermal energy dissipation in spinach exposed to high light and N stress. Plant Physiology, 113, 817–824.CrossRefGoogle ScholarPubMed
Verma, S.B., Baldocchi, D.D., Anderson, D.E., et al. (1986). Eddy fluxes of CO2, water vapor and sensible heat over a deciduous forest. Bounday Layer Meteorology, 36, 71–91.CrossRefGoogle Scholar
Verma, S.B., Dobermann, A., Cassman, K.G., et al. (2005). Annual carbon dioxide exchange in irrigated and rainfed maize-based agroecosystems. Agricultural and Forest Meteorology, 131, 77–96.CrossRefGoogle Scholar
Verma, S.B., Sellers, P.J., Walthall, C.L., et al. (1993). Photosynthesis and stomatal conductance related to reflectance on the canopy scale. Remote Sensing of the Environment, 44, 103–116.CrossRefGoogle Scholar
Vernon, L.P. (1960). Spectrophotometric determination of chlorophylls and pheophytins in plant extracts. Analytical Chemistry, 32, 1144–1150.CrossRefGoogle Scholar
Verslues, P.E. and Bray, E.A. (2006). Role of abscisic acid (ABA) and Arabidopsis thaliana ABA-insensitive loci in low water potential-induced ABA and proline accumulation. Journal of Experimental Botany, 57, 201–212.CrossRefGoogle ScholarPubMed
Vervuren, P.J.A., Blom, C.W.P.M. and De Kroon, H. (2003). Extreme flooding events on the Rhine and the survival and distribution of riparian plant species. Journal of Ecology, 91, 135–146.CrossRefGoogle Scholar
Vicentini, A., Barber, J.C., Aloscioni, S.S., et al. (2008). The age of the grasses and clusters of origins of C4 photosynthesis. Global Change Biology, 14, 2963–2977.CrossRefGoogle Scholar
Vieira Dos Santos, C. and Rey, P., (2006). Plant thioredoxins are key actors in the oxidative stress response. Trends in Plant Sciences, 11, 329–334.CrossRefGoogle ScholarPubMed
Vieira, F.C.B., He, Z.L., Wilson, P.C., et al. (2008). Response of representative cover crops to aluminum toxicity, phosphorus deprivation, and organic amendment. Australian Journal of Agricultural Research, 59, 52–61.CrossRefGoogle Scholar
Vierling, E. (1991). The roles of heat shock proteins in plants. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 579–620.CrossRefGoogle Scholar
Vígh, L., Combos, Z., Horváth, I., et al. (1989). Saturation of membrane lipids by hydrogenation induces thermal stability in chloroplast inhibiting the heat-dependent stimulation of photo-system I-mediated electron transport. Biochimica et Biophysica Acta (BBA) – Bioenergetics, 979, 361–364.CrossRefGoogle Scholar
Villar, R. and Merino, J. (2001). Comparison of leaf construction costs in woody species with differing leaf life-spans in contrasting ecosystems. New Phytologist, 151, 213–226.CrossRefGoogle Scholar
Vincent, D., Ergül, A., Bohlman, M.C., et al. (2007). Proteomic analysis reveals differences between Vitis vinifera L. cv. Chardonnay and cv. Cabernet Sauvignon and their responses to water deficit and salinity. Journal of Experimental Botany, 58, 1873–1892.CrossRefGoogle Scholar
Vingarzan, R. (2004) A review of surface ozone background levels and trends. Atmospheric Environment, 38, 3431–3442.CrossRefGoogle Scholar
Vitousek, P.M. and Denslow, J.S. (1986). Nitrogen and Phosphorus Availability in Treefall Gaps of a Lowland Tropical Rain-Forest. Journal of Ecology, 74, 1167–1178.CrossRefGoogle Scholar
Voesenek, L.A.C.J., Colmer, T.D., Pierik, R., et al. (2006). How plants cope with complete submergence. New Phytologist, 170, 213–226.CrossRefGoogle ScholarPubMed
Voesenek, L.A.C.J., Runders, J.H.G.M., Peeters, A.J.M., et al. (2004). Plant hormones regulate fast shoot elongation under water: from genes to communities. Ecology, 85, 16–27.CrossRefGoogle Scholar
Vogel, J.C. (1980). Fractionation of the carbon isotopes during photosynthesis. In: Sitzungsberichte der Heidelberger Akademie der wissenschaften, matematisch-naturwissenschaftliche Klasse Jahrgang 1980, 3, Abhandlung, Springer-Verlag, Heidelberg; Berlin; New York, USA, pp. 111–135.Google Scholar
Vogel, J.C. and Lerman, J.C. (1969). Groningen radiocarbon dates. VIII. Radiocarbon, 11, 351–390.CrossRefGoogle Scholar
Vogel, J.C., Fuls, A. and Ellis, R.P. (1978). Geographical distribution of Kranz grasses in South Africa. South African Journal of Science, 74, 209–215.Google Scholar
Vogelmann, T.C. (1989). Penetration of light into plants. Photochemistry and Photobiology, 50, 895–902.CrossRefGoogle Scholar
Vogelmann, T.C. (1993). Plant tissue optics. Annual Review of Plant Physiology and Plant Molecular Biology, 44, 231–251.CrossRefGoogle Scholar
Vogelmann, T.C. and Evans, J.R. (2002). Profiles of light absorption and chlorophyll within spinach leaves from chlorophyll fluorescence. Plant, Cell and Environment, 25, 1313–1323.CrossRefGoogle Scholar
Vogelmann, T.C. and Han, T. (2000). Measurement of gradients of absorbed light in spinach leaves from chlorophyll fluorescence profiles. Plant Cell Environment, 23, 1303–1311.CrossRefGoogle Scholar
Vogelmann, T.C. and Martin, G. (1993). The functional significance of palisade tissue: penetration of directional versus diffuse light. Plant Cell Environment, 16, 65–72.CrossRefGoogle Scholar
Vogelmann, T.C., Bornman, J.F. and Josserand, S. (1989). Photosynthetic light gradients and spectral regime within leaves of Medicago sativa. Philosophical Transactions of the Royal Society of London B, 323, 411–421.CrossRefGoogle Scholar
Vogelmann, T.C., Nishio, J.N. and Smith, W.K. (1996). Leaves and light capture: light propagation and gradients of carbon fixation within leaves. Trends in Plant Science, 1, 65–71.CrossRefGoogle Scholar
Volz, A. and Kley, D. (1988). Evaluation of the Montsouris series of ozone measurements made in the nineteenth century. Nature, 332, 240–242.CrossRefGoogle Scholar
von Caemmerer, S. (2000). Biochemical Models of Leaf Photosynthesis, Techniques in Plant Sciences No. 2. CSIRO Publishing, Collingwood, Victoria, Australia.Google Scholar
von Caemmerer, S. and Evans, J.R. (1991). Determination of the average partial pressure of CO2 in chloroplasts from leaves of several C3 plants. Australian Journal Plant Physiology, 18, 287–305.CrossRefGoogle Scholar
von Caemmerer, S. and Furbank, R.T. (1999). The modelling of C4 photosynthesis. In: The Biology of C4 Photosynthesis (ed. Sage, R.M.), Academic Press, New York, USA, pp. 173–211.Google Scholar
von Caemmerer, S. and Furbank, R.T. (2003). The C4 pathway: an efficient CO2 pump. Photosynthesis Research, 77, 191–207.CrossRefGoogle Scholar
von Caemmerer, S. and Quick, P. (2000). Rubisco: physiology in vivo. In: Advances in Photosynthesis: Physiology and Metabolism (eds Leegood, R.C., Sharkey, T.D. and von Caemmerer, S.), Kluwer Academic Publishers, Dodrecht, Boston, London, pp. 86–107.Google Scholar
von Caemmerer, S., Evans, J.R., Hudson, G.S., et al. (1994). The kinetics of ribulose-1,5-bisphosphate carboxylase/oxygenase in-vivo inferred from measurements of photosynthesis in leaves of transgenic tobacco. Planta, 195, 88–97.CrossRefGoogle Scholar
von Caemmerer, S., Lawson, T., Oxborough, K., et al. (2004). Stomatal conductance does not correlate with photosynthetic capacity in transgenic tobacco with reduced amounts of Rubisco. Journal of Experimental Botany, 55, 1157–1166.CrossRefGoogle Scholar
von Caemmerer, S.V. and Farquhar, G.D. (1981). Some relationships between the biochemistry of photosynthesis and the gas-exchange of leaves. Planta, 153, 376–387.CrossRefGoogle ScholarPubMed
von Willert, D.J., Armbruster, N.Drees, T., et al. (2005). Welwitschia mirabilis: CAM or not CAM – what is the answer?Functional Plant Biology, 32, 389–395.CrossRefGoogle Scholar
Voronin, P.Y., Ivanova, L.A., Ronzhina, D.A., et al. (2003). Structural and functional changes in the leaves of plants from steppe communities as affected by aridization of the eurasian climate. Russian Journal of Plant Physiology, 50, 604–611.CrossRefGoogle Scholar
Vountas, M., Rozanov, V.V. and Burrows, J.P. (1998). Filling in of Fraunhofer and gas-asorption lines in sky spectra as caused rotational Raman scattering. Journal of Quantitative Spectroscopy Radiative Transfer, 60, 943–961.CrossRefGoogle Scholar
Voznesenskaya, E.V., Chuong, S.D.X., Kiirats, O., et al. (2005a). Evidence that C4 species in genus Stipagrostis, family Poaceae, are NADP-malic enzyme subtype with nonclassical type of Kranz anatomy (Stipagrostoid). Plant Science, 168 (3), 731–739.CrossRefGoogle Scholar
Voznesenskaya, E.V., Chuong, S.D.X., Koteyeva, N.K., et al. (2005b). Functional compartmentation of C4 photosynthesis in the triple-layered chlorenchyma of Aristida (Poaceae). Functional Plant Biology, 32, 67–77.CrossRefGoogle Scholar
Voznesenskaya, E.V., Franceschi, V.R., Kiirats, O., et al. (2001). Kranz anatomy is not essential for terrestrial C4 plant photosynthesis. Nature, 414, 543–546.CrossRefGoogle Scholar
Vrábl, D., Vašková, M., Hronková, M., et al. (2009). Mesophyll conductance to CO2 transport estimated by two independent methods: effect of ambient CO2 concentration and abscisic acid. Journal of Experimental Botany, 60, 2315–2323.CrossRefGoogle Scholar
Vrba, E.S., Denton, G.H., Partridge, T.C., et al. (eds) (1995). Paleoclimate and evolution, with emphasis on human origins. Yale University Press, New Haven, CT. 547 p.Google Scholar
Vu, J.C.V. and Yelenosky, G. (1991). Photosynthesis responses of citrus trees to soil flooding. Physiologia Plantarum, 81, 7–14.CrossRefGoogle Scholar
Vu, J.C.V., Allen, L.H., Jr., Boote, K.J., et al. (1997). Effects of elevated CO2 and temperature on photosynthesis and Rubisco in rice and soybean. Plant Cell and Environment, 20, 68–76.CrossRefGoogle Scholar
Vyas, P., Bisht, M.S., Miyazawa, S.I., et al. (2007). Effects of polyploidy on photosynthetic properties and anatomy in leaves of Phlox drummondii. Functional Plant Biology, 34, 673–682.CrossRefGoogle Scholar
Wada, M., Kagawa, T. and Sato, Y. (2003). Chloroplast movement. Annual Review of Plant Physiology and Plant Molecular Biology, 54, 455–468.CrossRefGoogle ScholarPubMed
Wagner, J. and Larcher, W. (1981). Dependence of CO2 gas exchange and acid metabolism of the alpine CAM plant Sempervivum montanum on temperature and light. Oecologia, 50, 88–93.CrossRefGoogle ScholarPubMed
Wainwright, C.M. (1977). Sun-tracking and related leaf movements in a desert lupine (Lupinus arizonicus). American Journal of Botany, 64, 1032–1041.CrossRefGoogle Scholar
Walcroft, A.S., Whitehead, D., Silvester, W.B., et al.(1997). The response of photosynthetic model parameters to temperature and nitrogen concentration in Pinus radiata D. Don. Plant, Cell and Environment, 20, 1338–1348.CrossRefGoogle Scholar
Walker, B.H. and Noy Meir, I. (1982). Aspects of the stability and resilience of savanna ecosystems. In: Ecology of Tropical Savanna Ecosystems (ed. Huntley, B.J. and Walker, B.H.), Springer-Verlag, New York, USA, pp. 556–590.Google Scholar
Walker, D.A. (1993). Polarographic measurement of oxygen. In: Photosynthesis and Production in a Changing Environment. A Field and Laboratory Manual (eds Hall, D.O., Scurlock, J.M.O, Bolhar-Nordenkampf, H.R., Leegood, R.C. and Long, S.P.), Chapman and Hall, London, UK, pp. 168–180.
Walker, D.A. and Robinson, S.P. (1978). Chloroplast and cell. A contemporary view of photosynthetic carbon assimilation. Berichte Der Deutschen Botanischen Gesellschaft, 91, 513–526.Google Scholar
Walker, D.A. and Sivak, M.N. (1985). Can phosphate limit photosynthetic carbon assimilation in vivo?Physiologie Vegetale, 23, 829–841.Google Scholar
Walker, D.A. and Sivak, M.N. (1986). Photosynthesis and phosphate: a cellular affair?Trends in Biochemical Sciences, 11, 176–179.CrossRefGoogle Scholar
Wallin, G., Linder, S., Lindroth, A., et al. (2001). Carbon dioxide exchange in Norway spruce at the shoot, tree and ecosystem scale. Tree Physiology, 21, 969–976.CrossRefGoogle ScholarPubMed
Walter, A. and Schurr, U. (1999). The modular character of growth in Nicotiana tabacum plants under steady state nutrition. Journal of Experimental Botany, 50, 1169–1177.CrossRefGoogle Scholar
Walter, A. and Schurr, U. (2000). Spatial variability of leaf development, growth and function. In: Leaf Development and Canopy Growth (eds Marshall, B. and Roberts, J.), Sheffield Academic Press, Sheffield, UK.
Walter, A. and Schurr, U. (2005). Dynamics of leaf and root growth – endogenous control versus environmental impact. Annals Botany, 95, 891–900.CrossRefGoogle ScholarPubMed
Walter, A., Scharr, H., Gilmer, F., et al. (2007). Dynamics of seedling growth acclimation towards altered light conditions can be quantified via GROWSCREEN: a setup and procedure designed for rapid optical phenotyping of different plant species. New Phytologist, 174, 447–455.CrossRefGoogle ScholarPubMed
Walters, M.B., Kruger, E.L. and Reich, P.B. (1993a). Growth, biomass distribution and CO2 exchange of northern hardwood seedlings in high and low light: relationships with successional status and shade tolerance. Oecologia, 94, 7–16.CrossRefGoogle ScholarPubMed
Walters, M.B., Kruger, E.L. and Reich, P.B. (1993b). Relative growth rate in relation to physiological and morphological traits for northern hardwood tree seedlings: species, light environment and ontogenetic considerations. Oecologia, 96, 219–231.CrossRefGoogle ScholarPubMed
Walters, R.G. and Horton, P. (1991). Resolution of components of non-photochemical chlorophyll fluorescence quenching in barley leaves. Photosynthesis Research, 27, 121–133.CrossRefGoogle ScholarPubMed
Wand, S.J.E., Esler, K.J. and Bowie, M.R. (2001). Seasonal photosynthetic temperature responses and changes in δ13C under varying temperature regimes in leaf-succulent and drought-deciduous shrubs from the Succulent Karoo, South Africa. South African Journal of Botany, 67, 235–243.CrossRefGoogle Scholar
Wand, S.J.E., Esler, K.J., Rundel, P.W., et al. (1999). A preliminary study of the responsiveness to seasonal atmospheric and rainfall patterns of wash woodland species in the arid Richtersveld. Plant Ecology, 142, 149–160.CrossRefGoogle Scholar
Wanek, W., Heintel, S. and Richter, A. (2001). Preparation of starch and other carbon fractions from higher plant leaves for stable carbon isotope analysis. Rapid Communications in Mass Spectrometry, 15, 1136–1140.CrossRefGoogle ScholarPubMed
Wang, D. and Luthe, D. S. (2003). Heat sensitivity in a bentgrass variant. Failure to accumulate a chloroplast heat shock protein isoform implicated in heat tolerance. Plant Physiology, 133, 319–327.CrossRefGoogle Scholar
Wang, D., Heckathorn, S.A., Barua, D., et al. (2008). Effects of elevated CO2 on the tolerance of photosynthesis to acute heat stress in C3, C4, and CAM species. American Journal of Botany, 95, 165–176.CrossRefGoogle ScholarPubMed
Wang, D.S., Naidu, S.L., Portis, A.R., et al. (2008). Can the cold tolerance of C-4 photosynthesis in Miscanthusxgiganteus relative to Zea mays be explained by differences in activities and thermal properties of Rubisco?Journal of Experimental Botany, 59, 1779–1787.CrossRefGoogle Scholar
Wang, F.Z., Wang, Q.B. and Kwon, S.Y. (2005). Enhanced drought tolerance of transgenic O. sativa plants expressing a pea manganese superoxide dismutase. Journal of Plant Physiology, 162, 465–472.CrossRefGoogle Scholar
Wang, H. and Jin, J.Y. (2005). Photosynthetic rate, chlorophyll fluorescence parameters, and lipid peroxidation of maize leaves as affected by zinc deficiency. Photosynthetica, 43, 591–596.CrossRefGoogle Scholar
Wang, N. and Nobel, P.S. (1996). Doubling the CO2 concentration enhanced the activity of carbohydrate-metabolism enzymes, source carbohydrate production, photoassimilate transport, and sink strength for Opuntia ficus-indica. Plant Physiology, 110, 893–902.CrossRefGoogle ScholarPubMed
Wang, P. and Song, C.P. (2008). Guard-cell signalling for hydrogen peroxide and abscisic acid. New Phytologist, 178, 703–718.CrossRefGoogle ScholarPubMed
Wang, X., Manning, W., Feng, Z., et al. (2007). Ground-level ozone in China: Distribution and effects on crop yields. Environmental Pollution, 147, 394–400.CrossRefGoogle ScholarPubMed
Wang, X.F. and Yakir, D. (1995). Temporal and spatial variation in the oxygen-18 content of leaf water in different plant species. Plant, Cell and Environment, 18, 1377–1385.CrossRefGoogle Scholar
Wang, X.Q., Wu, W.H. and Assmann, S.M. (1998). Differential responses of adaxial and abaxial guard cell of broad bean to abscisic acid and calcium. Plant Physiology, 118, 1421–1429.CrossRefGoogle ScholarPubMed
Wang, Y. and Deng, T. (2005). A 25 m.y. isotopic record of paleodiet and environmental change from fossil mammals and paleosols from the NE margin of the Tibetan Plateau. Earth and Planetary Science Letters, 236, 322–338.CrossRefGoogle Scholar
Wang, Y. and Jacob, D.J. (1998). Anthropogenic forcing on tropospheric ozone and OH since preindustrial times. Journal of Geophysical Research, 103, D23.CrossRefGoogle Scholar
Wang, Y., Cerling, T.E. and MacFadden, B.J. (1994). Fossil horses and carbon isotopes – New evidence for Cenozoic dietary habit and ecosystem changes in North America. Palaeogeography, Palaeoclimatology and Palaeoecology, 107, 269–279.CrossRefGoogle Scholar
Wang, Y., Guignard, G., Thévenard, F., et al. (2005). Cuticular anatomy of Sphenobaiera huangii (Ginkgoales) from the Lower Jurassic of Hubei, China. American Journal of Botany, 92, 709–721.CrossRefGoogle ScholarPubMed
Wang, Z.Q. (1996). Recovery of vegetation from the terminal Permian mass extinction in north China. Review of Palaeobotany and Palynology, 91, 121–142.Google Scholar
Ward, J., Baker, H. and Beale, M.H. (2007). Recent applications of NMR spectroscopy in plant metabolomics. FEBS Journal, 274, 1126–1131.CrossRefGoogle Scholar
Wardle, D.A., Hornberg, G., Zackrisson, O., et al. (2003). Long-term effects of wildfire on ecosystem properties across an island area gradient. Science, 300, 972–976.CrossRefGoogle ScholarPubMed
Waring, R.H. and Running, S.W. (1998). Forest Ecosystems. Analysis at multiple scales. Academic Press, San Diego, USA, pp. 1–370.Google Scholar
Warner, D.A. and Edwards, G.E. (1989). Effects of polyploidy on photosynthetic rates, photosynthetic enzymes, contents of DNA, chlorophyll, and sizes and numbers of photosynthetic cells in the C4 dicot Atriplex confertifolia. Plant Physiology, 91, 1143–1451.CrossRefGoogle Scholar
Warner, D.A. and Edwards, G.E. (1993). Effects of polyploidy on photosynthesis. Photosynthesis Research, 35, 135–147.CrossRefGoogle ScholarPubMed
Warner, D.A., Ku, M.S.B. and Edwards, G.E. (1987). Photosynthesis, leaf anatomy, and cellular constituents in the polyploid C4 grass Panicum virgatum. Plant Physiology, 84, 461–466.CrossRefGoogle Scholar
Warren, C.R. (2004). The photosynthetic limitation posed by internal conductance to CO2 movement is increased by nutrient supply. Journal of Experimental Botany, 55, 2313–2321.CrossRefGoogle ScholarPubMed
Warren, C.R. (2006a). Why does photosynthesis decrease with needle age in Pinus pinaster?Trees-Structure and Function, 20, 157–164.CrossRefGoogle Scholar
Warren, C.R. (2006b). Estimating the internal conductance to CO2 movement. Functional Plant Biology, 33, 431–442.CrossRefGoogle Scholar
Warren, C.R. (2007). Does growth temperature affect the temperature response of photosynthesis and internal conductance to CO2? A test with Eucalyptus regnans. Tree Physiology, 28, 11–19.CrossRefGoogle Scholar
Warren, C.R. (2008a). Stand aside stomata, another actor deserves centre stage: the forgotten role of the internal conductance to CO2 transfer. Journal Experimental Botany, 59, 1475–1487.CrossRefGoogle ScholarPubMed
Warren, C.R. (2008b). Soil water deficits decrease the internal conductance to CO2 transfer but atmospheric water deficits do not. Journal of Experimental Botany, 59, 327–334.CrossRefGoogle ScholarPubMed
Warren, C.R. and Adams, M.A. (2000). Trade-offs between the persistence of foliage and productivity in two Pinus species. Oecologia, 124, 487–494.CrossRefGoogle ScholarPubMed
Warren, C.R. and Adams, M.A. (2001). Distribution of N, Rubisco and photosynthesis in Pinus pinaster and acclimation to light. Plant, Cell and Environment, 24, 597–609.CrossRefGoogle Scholar
Warren, C.R. and Adams, M.A. (2002). Phosphorus affects growth and partitioning of nitrogen to Rubisco in Pinus pinaster. Tree Physiology, 22, 11–19.CrossRefGoogle ScholarPubMed
Warren, C.R. and Adams, M.A. (2004). Evergreen trees do not maximize instantaneous photosynthesis. Trends in Plant Science, 9, 270–274.CrossRefGoogle Scholar
Warren, C.R. and Adams, M.A. (2006). Internal conductance does not scale with photosynthetic capacity: implications for carbon isotope discrimination and the economics of water and N use in photosynthesis. Plant, Cell and Environment, 29, 192–201.CrossRefGoogle ScholarPubMed
Warren, C.R. and Dreyer, E. (2006). Temperature response of photosynthesis and internal conductance to CO2: results from two independent approaches. Journal of Experimental Botany, 57, 3057–3067.CrossRefGoogle ScholarPubMed
Warren, C.R., Adams, M.A. and Chen, Z.L. (2000). Is photosynthesis related to concentrations of nitrogen and Rubisco in leaves of Australian native plants?Australian Journal of Plant Physiology, 27, 407–416.Google Scholar
Warren, C.R., Aranda, I. and Cano, F.J. (2011). Responses to water stress of gas exchange and metabolites in Eucalyptus and Acacia spp. Plant, Cell and Environment, 34, 1609–1629.CrossRef
Warren, C.R., Dreyer, E. and Adams, M.A. (2003b). Photosynthesis-Rubisco relationships in foliage of Pinus sylvestris in response to nitrogen supply and the proposed role of Rubisco and amino acids as nitrogen stores. Trees: Structure and Function, 17, 359–366.Google Scholar
Warren, C.R., Dreyer, E., Tausz, M., et al. (2006). Ecotype adaptation and acclimation of leaf traits to rainfall in 29 species of 16-year-old Eucalyptus at two common gardens. Functional Ecology, 20, 929–940.CrossRefGoogle Scholar
Warren, C.R., Ethier, G.H., Livingston, N.J., et al. (2003a). Transfer conductance in second growth Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco) canopies. Plant, Cell and Environment, 26, 1215–1227.CrossRefGoogle Scholar
Warren, C.R., Hovenden, M.J., Davidson, N.J., et al. (1998). Cold hardening reduces photoinhibition of Eucalypts nitens and E. pauciflora at frost temperatures. Oecologia, 113, 350–359.CrossRefGoogle Scholar
Warren, C.R., Livingston, N.J. and Turpin, D.H. (2004). Water stress decreases the transfer conductance of Douglas fir (Pseudotsuga menziesii) seedlings. Tree Physiology, 24, 971–979.CrossRefGoogle ScholarPubMed
Warren, C.R., Low, M., Matyssek, R., et al. (2007). Internal conductance to CO2 transfer of adult Fagus sylvatica: Variation between sun and shade leaves and due to free-air ozone fumigation. Environmental and Experimental Botany, 59, 130–138.CrossRefGoogle Scholar
Warren, C.R., McGrath, J.F. and Adams, M.A. (2001). Water availability and carbon isotope discrimination in conifers. Oecologia, 127, 476–486.CrossRefGoogle ScholarPubMed
Wartinger, A., Heilmeier, H., Hartung, W., et al. (1990). Daily and seasonal courses of leaf conductance and abscisic acid in the xylem sap of almond trees [Prunus dulcis (Miller) D.A. Webb] under desert conditions. The New Phytologist, 116, 581–587.CrossRefGoogle Scholar
Waters, E.R., Lee, G.J. and Vierling, E. (1996). Evolution, structure, and function of the small heat shock proteins in plants. Journal of Experimental Botany, 47, 325–338.CrossRefGoogle Scholar
Watkins, J.E., Rundel, P.W. and Cardelus, C.L. (2007). The influence of life form on carbon and nitrogen relationships in tropical rainforest ferns. Oecologia, 153, 225–232.CrossRefGoogle ScholarPubMed
Watkins, Jr., J.E., Mack, M.C., Sinclair, T.R., et al. (2007). Ecological and evolutionary consequences of desiccation tolerance in tropical fern gametophytes. The New Phytologist, 176, 708–717.CrossRefGoogle ScholarPubMed
Watkinson, J.I., Sioson, A.A., Vasquez-Robinet, C., et al. (2003). Photosynthetic acclimation is reflected in specific patterns of gene expression in drought-stressed loblolly pine. Plant Physiology, 133, 1702–1716.CrossRefGoogle ScholarPubMed
Watling, J.R. and Press, M.C. (2001). Impacts of infection by parasitic angiosperms on host photosynthesis. Plant Biology, 3, 244–250.CrossRefGoogle Scholar
Watling, J.R., Robinson, S.A., Woodrow, I.E., et al. (1997). Responses of rainforest understory plants to excess light during sunflecks. Australian Journal of Plant Physiology, 24, 17–25.CrossRefGoogle Scholar
Watson, D.J. (1947). Comparative physiological studies on the growth of field crops. Annals of Botany, 11, 41–76.CrossRefGoogle Scholar
Watson, D.J. (1952). The physiological basis of variation in yield. Advances in Agronomy, 4, 101–145.CrossRefGoogle Scholar
Watson, M.A. and Caspar, B.B. (1984). Morphogenetic constraints on patterns of carbon distribution in plants. Annual Review of Ecology and Systematics, 15, 233–258.CrossRefGoogle Scholar
Weaver, J.C., Powell, K.T., Mintzer, R.A., et al. (1984). The diffusive permeability of bilayer membranes the contribution of transient aqueous pores. Bioelectrochemistry and Bioenergetics, 12, 405–412.CrossRefGoogle Scholar
Weaver, L.M. and Amasino, R.M. (2001). Senescence is induced in individually darkened Arabidopsis leaves, but inhibited in whole darkened plants. Plant Physiology, 127, 876–886.CrossRefGoogle ScholarPubMed
Webb, E.K., Pearman, G.I. and Leuning, R. (1980). Correction of flux measurements for density effects due to heat and water vapor transfer. Quarterly Journal of the Royal Meteorological Society, 106, 85–100.CrossRefGoogle Scholar
Wedin, D.A. and Tilman, D. (1993). Competition among grasses along a nitrogen gradient: initial conditions and mechanisms of competition. Ecological Monographs, 63, 199–229.CrossRefGoogle Scholar
Wedin, D.A. and Tilman, D. (1996). Influence of nitrogen loading and species composition on the carbon balance of grasslands. Science, 274, 1720–1723.CrossRefGoogle ScholarPubMed
Wedler, M., Köstner, B. and Tenhunen, J.D. (1996). Understory contribution to stand total water loss at an old Norway spruce forest. In: Verhandlungen der Gessellschaft für Ökologie, vol 26 (eds Pfadenhauer, J., Kappen, L., Mahn, E.G., Otte, A. and Plachter, H.) Gustav Fischer Verlag, Stuttgart, Germany, pp. 69–77.
Weigel, H.J. (1985). Inhibition of photosynthetic reactions of isolated chloroplasts by cadmium. Journal of Plant Physiology, 119, 179–189.CrossRefGoogle Scholar
Weih, M. and Karlsson, P.S. (2001). Growth response of Mountain birch to air and soil temperature: is increasing leaf-nitrogen content an acclimation to lower air temperature?The New Phytologist, 150, 147–155.CrossRefGoogle Scholar
Weikert, R.M., Wedler, M., Lippert, M., et al. (1989). Photosynthetic performance, chloroplast pigments, and mineral content of various needle age classes of spruce (Picea abies) with and without the new flush: an experimental approach for analysing forest decline phenomena. Trees, 3, 161–172.CrossRefGoogle Scholar
Weinbaum, S.A., Muraoka, T.T. and Plant, R.E. (1994). Intracanopy variation in nitrogen cycling through leaves is influenced by irradiance and proximity to developing fruit in mature walnut trees. Trees, 9, 6–11.CrossRefGoogle Scholar
Weis, E. (1982). Influence of light on the heat sensitivity of the photosynthetic apparatus in isolated spinach chloroplasts. Plant Physiology, 70, 1530–1534.CrossRefGoogle ScholarPubMed
Weis, E. (1985a). Chlorophyll fluorescence at 77K in intact leaves: characterization of a technique to eliminate artifacts related to self-absorption. Photosynthesis Research, 6, 73–86.CrossRefGoogle Scholar
Weis, E. (1985b). Light- and temperature-induced changes in the distribution of excitation energy between photosystem I and photosystem II in spinach leaves. Biochimica et Biophysica Acta, 807, 118–126.CrossRefGoogle Scholar
Weis, E., Meng, Q., Siebke, K., et al. (1998). Topography of leaf carbon metabolism as analyzed by chlorophyll a-fluorescence. In: Photosynthesis: Mechanisms and Effects, vol 5 (ed. Garab, G.), Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 4259–4264.Google Scholar
Weise, S.E., Schrader, S.M., Kleinbeck, K.R., et al. (2006). Carbon balance and circadian regulation of hydrolytic and phosphorolytic breakdown of transitory starch. Plant Physiology, 141, 879–886.CrossRefGoogle ScholarPubMed
Weiss, I., Mizrahi, Y. and Raveh, E. (2009). Chamber response time: a neglected issue in gas exchange measurements. Photosynthetica, 47, 121–124.CrossRefGoogle Scholar
Welling, A. and Palva, E.T. (2008). Involvement of CBF transcription factors in winter hardiness in birch. Plant Physiology, 147, 1199–1211.CrossRefGoogle ScholarPubMed
Wells, R., Schulze, L.L., Ashley, D.A., et al. (1982). Cultivar differences in canopy apparent photosynthesis and their relationship to seed yield in soybeans. Crop Science, 22, 886–890.CrossRefGoogle Scholar
Welp, L.R., Randerson, T.R. and Liu, H.P. (2007). The sensitivity of carbon fluxes to spring warming and summer drought depends on plant functional type in boreal forest ecosystems. Agricultural and Forest Meteorology, 147, 172–185.CrossRefGoogle Scholar
Welsch, R., Beyer, P., Hugueney, P., et al. (2000). Regulation and activation of phytoene synthase, a key enzyme in carotenoid biosynthesis, during photomorphogenesis. Planta, 211, 846–854.CrossRefGoogle ScholarPubMed
Welter, S.C. (1989). Arthropod impact on plant gas exchange. In: Insect-Plant Interactions (ed. Bernays, E.A.), CRC Press, Boca Raton, FL, USA, pp. 135–151.Google Scholar
Wen, X.F., Sun, X.M., Zhang, S.C., et al. (2008). Continuous measurements of water vapour D/H and 18O/16O isotope ratios in the atmosphere. Journal of Hydrology, 349, 489–500.CrossRefGoogle Scholar
Weng, X.Y., Zheng, C.J., Xu, H.X., et al. (2007). Characteristics of photosynthesis and functions of the water-water cycle in rice (Oryza sativa) leaves in response to potassium deficiency. Physiologia Plantarum, 131, 614–621.CrossRefGoogle ScholarPubMed
Werk, K.S., Ehleringer, J., Forseth, I.N., et al. (1983). Photosynthetic characteristics of Sonoran Desert winter annuals. Oecologia, 59, 101–105.CrossRefGoogle ScholarPubMed
Werner, C. and Máguas, C. (2010). Carbon isotope discrimination as a tracer of functional traits in a Mediterranean macchia plant community. Functional Plant Biology, 37, 467–477.CrossRefGoogle Scholar
Werner, C., Correia, O. and Beyschlag, W. (1999). Two different strategies of Mediterranean macchia plants to avoid photoinhibitory damage by excessive radiation levels during summer drought. Acta Oecologica, 20, 15–23.CrossRefGoogle Scholar
Werner, C., Correia, O. and Beyschlag, W. (2002). Characteristic patterns of chronic and dynamic photoinhibition of different functional groups in a Mediterranean ecosystem. Functional Plant Biology, 29, 999–1011.CrossRefGoogle Scholar
Werner, C., Ryel, R.J., Correia, O., et al. (2001a). Structural and functional variability within the canopy and its relevance for carbon gain and stress avoidance. Acta Oecologica, 22, 129–138.CrossRefGoogle Scholar
Werner, C., Ryel, R.J., Correia, O., et al. (2001b). Effects of photoinhibition on whole-plant carbon gain assessed with a photosynthesis model. Plant, Cell and Environment, 24, 27–40.CrossRefGoogle Scholar
West, J.B., Espeleta, J.F. and Donovan, L.A. (2003). Root longevity and phenology differences between two co-occurring savanna bunchgrasses with different leaf habits. Functional Ecology, 17, 20–28.CrossRefGoogle Scholar
West, J.D., Peak, D., Peterson, J.Q., et al. (2005). Dynamics of stomatal patches for a single surface of Xanthium strumarium L. leaves observed with fluorescence and thermal images. Plant, Cell and Environment, 28, 633–641.CrossRefGoogle Scholar
Westbeek, M.H.M, Pons, T.L., Cambridge, M.L., et al. (1999). Analysis of differences in photosynthetic nitrogen use eficiency of alpine and lowland Poa species. Oecologia, 19, 120–126.Google Scholar
Westhoff, P. and Gowik, U. (2004). Evolution of C4 phosphoenolpyruvate carboxylase. Genes and proteins: a case study with the genus Flaveria. Annals of Botany, 93, 13–23.CrossRefGoogle ScholarPubMed
Weston, D.J., Bauerle, W.L., Swire-Clark, G.A., et al. (2007). Characterization of Rubisco activase from thermally contrasting genotypes pf Acer rubrum (Aceraceae). American Journal of Botany, 94, 926–934.CrossRefGoogle Scholar
Wheeler, R.M. (1992). Gas-exchange measurements using a large, closed plant growth chamber. HortScience, 27, 777–780.Google ScholarPubMed
Wheeler, R.M., Corey, K.A., Sager, J.C., et al. (1993). Gas exchange characteristics of wheat stands grown in a closed, controlled environment. Crop Science, 33, 161–168.CrossRefGoogle Scholar
Whiley, A.W., Schaffer, B. and Lara, S.P. (1992). Carbon dioxide exchange of developing avocado (Persea americana Mill.) fruit. Tree Physiology, 11, 85–94.CrossRefGoogle ScholarPubMed
White, J.G. and Zasoski, R.H. (1999). Mapping soil micronutrients. Field Crop Research, 60, 11–26.CrossRefGoogle Scholar
White, T.A., Campbell, B.D., Kemp, P.D., et al. (2000). Sensitivity of three grassland communities to simulated extreme temperature and rainfall events. Global Change Biology, 6, 671–684.CrossRefGoogle Scholar
Whitehead, D. (1998). Regulation of stomatal conductance and transpiration in forest canopies. Tree Physiology, 18, 633–644.CrossRefGoogle ScholarPubMed
Whitehead, D. and Beadle, C.L. (2004). Physiological regulation of productivity and water use in Eucalyptus: a review. Forest Ecology and Management, 193, 113–140.CrossRefGoogle Scholar
Whitehead, D., Barbour, M.M., Griffin, K.L. et al. (2011). Effects of leaf age and tree size on stomatal and mesophyll limitations to photosynthesis in mountain beech (Nothofagus solandrii var. cliffortiodes). Tree Physiology, 31, 985–996.CrossRef
Whitehead, D., Livingston, N.J., Kelliher, P.M., et al. (1996). Response of transpiration and photosynthesis to a transient change in illuminated foliage area for a Pinus radiata D. Don tree. Plant, Cell and Environment, 19, 949–957.CrossRefGoogle Scholar
Whitelam, G. (1995). Plant photomorphogenesis – a green light for cryptochrome research. Current Biology, 5, 1351–1353.CrossRefGoogle Scholar
Whitham, S., Quan, S., Chang, H.S., et al. (2003). Diverse RNA viruses elicit the expression of common sets of genes in susceptible Arabidopsis thaliana plants. Plant Journal, 32, 271–283.CrossRefGoogle Scholar
Whiting, M.D. and Lang, G.A. (2001). Canopy architecture and cuvette flow patterns influence whole-canopy net CO2 exchange and temperature in sweet cherry. HortScience, 36, 691–698.Google Scholar
Whitmore, T.C. (1990). An introduction to tropical rain forests. Oxford University Press, Oxford, UK.Google Scholar
Whitney, S.M. and Andrews, T.J. (2001). The gene for the ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) small subunit relocated to the plastid genome of tobacco directs the synthesis of small subunits that assemble into Rubisco. Plant Cell, 13, 193–205.CrossRefGoogle ScholarPubMed
Whitney, S.M. and Sharwood, R.E. (2008). Construction of a tobacco master line to improve Rubisco engineering in chloroplasts. Journal of Experimental Botany, 59, 1909–1921.CrossRefGoogle ScholarPubMed
Wi, S.J., Kim, W.T. and Park, K.Y. (2006). Overexpression of carnation S-adenosylmethionine decarboxylase gene generates a broad-spectrum tolerance to abiotic stresses in transgenic tobacco plants. Plant Cell Report, 25, 1111–1121.CrossRefGoogle ScholarPubMed
Wickman, F.E. (1952). Variations in the relative abundance of the carbon isotopes in plants. Geochimica et Cosmochimica Acta, 2, 243–252.CrossRefGoogle Scholar
Wild, M., Gilgen, H., Roesch, A., et al. (2005). From dimming to brightening: decadal changes in solar radiation at Earth’s surface. Science, 308, 847–850.CrossRefGoogle ScholarPubMed
Wildi, B. and Lütz, C. (1996). Antioxidant composition of selected high alpine plant species from different altitudes. Plant, Cell and Environment, 19, 138–146.CrossRefGoogle Scholar
Wilhelmová, N., Procházková, D., Sindelárová, M., et al. (2005). Photosynthesis in leaves of Nicotiana tabacum L. infected with tobacco mosaic virus. Photosynthetica, 43, 597–602.CrossRefGoogle Scholar
Wilkins, M.B. (1992). Circadian rhythms: their origin and control. New Phytologist, 121, 347–375.CrossRefGoogle Scholar
Wilkinson, S. and Davies, W.J. (1997). Xylem sap pH increase: a drought signal received at the apoplastic face of the guard cells that involves the suppression of saturable abscisic acid uptake by the epidermal symplast. Plant Physiology, 113, 559–573.CrossRefGoogle Scholar
Wilkinson, S. and Davies, W.J. (2002). ABA-based chemical signalling: the co-ordination of responses to stress in plants. Plant, Cell and Environment, 25, 195–210.CrossRefGoogle ScholarPubMed
Wilkinson, S. and Davies, W.J. (2008). Manipulation of the apoplastic pH of intact plants mimics stomatal and growth responses to water availability and microclimatic variation. Journal of Experimental Botany, 59, 619–631.CrossRefGoogle ScholarPubMed
Wilkinson, S. and Davies, W.J. (2009). Ozone suppresses soil drying- and abscisic acid (ABA)- induced stomatal closure via an ethylene-dependent mechanism. Plant, Cell and Environment, 32, 949–959.CrossRefGoogle ScholarPubMed
Wilkinson, S. and Davies, W.J. (2010). Drought, ozone, ABA and ethylene: new insights from cell to plant to community. Plant, Cell and Environment, 33, 510–525.CrossRefGoogle Scholar
Wilkinson, S., Clephan, A.L. and Davies, W.J. (2001). Rapid low temperature-induced stomatal closure occurs in cold-tolerant Commelina communis leaves but not in cold-sensitive tobacco leaves, via a mechanism that involves apoplastic calcium but not abscisic acid. Plant Physiology, 126(4), 1566–1578.CrossRefGoogle Scholar
Williams, E.L., Hovenden, M.J. and Close, D.C. (2003). Strategies of light energy utilisation, dissipation and attenuation in six co-occurring alpine heath species in Tasmania. Functional Plant Biology, 30, 1205–1218.CrossRefGoogle Scholar
Williams, G.J. III (1974). Photosynthetic adaptation to temperature in C3 and C4 grasses: possible ecological role in shortgrass prarie. Plant Physiology, 54, 709–711.CrossRefGoogle Scholar
Williams, K., Field, C.B. and Mooney, H.A. (1989). Relationships among leaf construction costs, leaf longevity, and light environment in rainforest species of the genus Piper. American Naturalist, 133, 198–211.CrossRefGoogle Scholar
Williams, M., Rastetter, E.B., Fernandes, D.N., et al. (1996). Modelling the soil-plant-atmosphere continuum in a Quercus-Acer stand at Harvard Forest: the regulation of stomatal conductance by light, nitrogen and soil/plant hydraulic properties. Plant, Cell and Environment, 19, 911–927.CrossRefGoogle Scholar
Williams, M., Woodward, F.I., Baldocchi, D.D., et al. (2004). CO2 capture from the leaf to the landscape. In: Photosynthetic Adaptation: Chloroplast to the Landscape (eds Smith, W.K., Vogelmann, T.C. and Critchley, C.), Springer, New York, USA, pp. 133–168.Google Scholar
Williams, T.G., Flanagan, L.B. and Coleman, J.R. (1996). Photosynthetic gas exchange and discrimination against 13CO2 and C18O16O in tobacco plants modified by an antisense construct to have low chloroplastic carbonic anhydrase. Plant Physiology, 112, 319–326.CrossRefGoogle ScholarPubMed
Williams, W.P., Brain, A.P.R. and Dominy, P.J. (1992). Induction of non-bilayer lipid phase separation in chloroplast thylakoid membranes by compatible co-solutes and its relation to the thermal stability of photosystem II. Biochimica et Biophysica Acta, 1099, 137–144.Google Scholar
Willmer, C. and Fricker, M. (1996). Stomata, 2nd edn. London: Chapman and Hall.CrossRefGoogle Scholar
Willms, J.R., Salon, C. and Layzell, D.B. (1999). Evidence for light-stimulated fatty acid synthesis in soybean fruit. Plant Physiology, 120, 1117–1128.CrossRefGoogle ScholarPubMed
Wilson, J.A., OgunKanmi, A.B. and Mansfield, T.A. (1978). Effects of external potassium supply on stomatal closure induced by abscisic acid. Plant, Cell and Environment, 1, 199–201.CrossRefGoogle Scholar
Wilson, K., Goldstein, A., Falge, E., et al. (2002). Energy balance closure at FLUXNET sites. Agricultural and Forest Meteorology, 113, 223–243.CrossRefGoogle Scholar
Wilson, K.B., Baldocchi, D.D. and Hanson, P.J. (2001). Leaf age affects the seasonal pattern of photosynthetic capacity and net ecosystem exchange of carbon in a deciduous forest. Plant, Cell and Environment, 24, 571–583.CrossRefGoogle Scholar
Wilson, K.B., Baldocchi, D.D. and Hanson, P.J. (2000a). Quantifying stomatal and non-stomatal limitations to carbon assimilation resulting from leaf aging and drought in mature deciduous tree species. Tree Physiology, 20, 787–797.CrossRefGoogle ScholarPubMed
Wilson, K.B., Baldocchi, D.D. and Hanson, P.J. (2000b). Spatial and seasonal variability of photosynthetic parameters and their relationship to leaf nitrogen in a deciduous forest. Tree Physiology, 20, 565–578.CrossRefGoogle Scholar
Wilson, K.E., Krol, M. and Huner, N.P.A. (2003). Temperature-induced greening of Chlorella vulgaris. The role of the cellular energy balance and zeaxanthin-dependent nonphotochemical quenching. Planta, 217, 616–627.CrossRefGoogle ScholarPubMed
Winder, T.L. and Nishio, J. (1995). Early iron deficiency stress response in leaves of sugar beet. Plant Physiology, 108, 1487–1494.CrossRefGoogle ScholarPubMed
Wingler, A., Marés, M. and Pourtau, N. (2004). Spatial and metabolic regulation of photosynthetic parameters during leaf senescence. New Phytologist, 161, 781–789.CrossRefGoogle Scholar
Wingler, A., Masclaux-Daubresse, C. and Fischer, A.M. (2009). Sugars, senescence, and ageing in plants and heterotrophic organisms. Journal of Experimental Botany, 60, 1063–1066.CrossRefGoogle ScholarPubMed
Winkel, A. and Borum, J. (2009). Use of sediment CO2 by submersed rooted plants. Annals of Botany, 103, 1015–1023.CrossRefGoogle ScholarPubMed
Winkel, T., Méthy, M. and Thénot, F. (2002). Radiation use efficiency, chlorophyll fluorescence, and reflectance indices associated with ontogenic changes in water-limited Chenopodium quinoa leaves. Photosynthetica, 40, 227–232.CrossRefGoogle Scholar
Winner, W.E., Lefohn, A.S., Cotter, I.S., et al. (1989). Plant responses to elevational gradients of O3 exposures in Virginia. Proceedings of the National Academy of Sciences of the USA, 86, 8828–8832.CrossRefGoogle Scholar
Winner, W.E., Mooney, H.A., Williams, K., et al. (1985). Measuring and assessing SO2 effects on photosynthesis and plant growth. In: Sulfur Dioxide and Vegetation-Physiology, Ecology and Policy Issues (eds Winner, W.E., Mooney, H.A. and Goldstein, R.A.), Stanford University Press, Stanford, USA, pp. 118–132.Google Scholar
Winsemius, H.C., Savenije, H.H.G. and Bastiaanssen, W.G.M. (2008). Constraining model parameters on remotely sensed evaporation: justification for distribution in ungauged basins?Hydrology and Earth System Sciences, 12, 1403–1413.CrossRefGoogle Scholar
Winslow, J.C., Hunt, E.R. and Piper, S.C. (2003). The influence of seasonal water availability on global C3 versus C4 grassland biomass and its implications for climate change research. Ecological Modelling, 163, 153–173.CrossRefGoogle Scholar
Winter, K. (1973). Zum Problem der Ausbildung des Crassulaceen-Säurestoffwechsels bei Mesembryanthemum crystallinum unter NaCl-Einfluss. Planta, 109, 135–145.CrossRefGoogle Scholar
Winter, K. and Smith, J.A.C. (ed). (1996). Crassulacean Acid Metabolism: Biochemistry, Ecophysiology, and Evolution. Springer, Berlin, Germany.CrossRefGoogle Scholar
Winter, K., Aranda, J. and Holtum, J.A.M. (2005). Carbon isotope composition and water-use efficiency in plants with crassulacean acid metabolism. Functional Plant Biology, 32, 381–388.CrossRefGoogle Scholar
Winter, K., Garcia, M. and Holtum, J.A.M. (2009). Canopy CO2 exchange of two neotropical tree species exhibiting constitutive and facultative CAM photosynthesis, Clusia rosea and Clusia cylindrica. Journal of Experimental Botany, 60, 3167–3177.CrossRefGoogle ScholarPubMed
Winter, K., Osmond, C.B. and Pate, J.S. (1981). Coping with salinity. In: The Biology of Australian Plants (eds Pate, J.S. and McComb, A.J.), University of Western Australia Press, Netlands, Australia, pp. 88–113.Google Scholar
Winter, K., Richter, A., Engelbrecht, B., et al. (1997). Effect of elevated CO2 on growth and crassulacean acid metabolism activity of Kalanchoë pinnata under tropical conditions. Planta, 201, 389–396.CrossRefGoogle Scholar
Winter, K., Wallace, B.J., Stocker, G.C., et al. (1983). Crassulacean acid metabolism in Australian vascular epiphytes and some related species. Oecologia, 57, 129–141.CrossRefGoogle ScholarPubMed
Wintermans, J. and De Mots, A. (1965). Spectrophotometric of chlorophyll a and b and their pheophytins in ethanol. Biochemistry and Biophysics Acta, 109, 448–453.Google Scholar
Wise, R.R. (1995). Chilling-enhanced photooxidation – the production, action and study of reactive oxygen species produced during chilling in the light. Photosynthesis Research, 45, 79–97.CrossRefGoogle Scholar
Wise, R.R. and Ort, D.R. (1989). Photophosphorylation after chilling in the light: Effects on membrane energization and coupling factor activity. Plant Physiology, 90, 657–664.CrossRefGoogle ScholarPubMed
Wise, R.R., Olson, A.J., Schrader, S.M., et al. (2004). Electron transport is the functional limitation of photosynthesis in field-grown Pima cotton plants at high temperature. Plant, Cell and Environment, 27, 717–724.CrossRefGoogle Scholar
Wisniewski, M., Bassett, C. and Gusta, L.V. (2003). An overview of cold hardiness in woody plants: Seeing the forest through the trees. Hortscience, 38, 952–959.Google Scholar
Wittig, V.E., Ainsworth, E.A. and Long, S.P. (2007). To what extent do current and projected increases in surface ozone affect photosynthesis and stomatal conductance of trees? A meta-analytic review of the last 3 decades of experiments. Plant, Cell and Environment, 30, 1150–1162.CrossRefGoogle Scholar
Wittmann, C. and Pfanz, H. (2007). Temperature dependency of bark photosynthesis in beech (Fagus sylvatica L.) and birch (Betula pendula Roth.) trees. Journal of Experimental Botany, 58, 4293–4306.CrossRefGoogle ScholarPubMed
Wittmann, C., Pfanz, H., Loreto, F., et al. (2006). Stem CO2 release under illumination: corticular photosynthesis, photorespiration or inhibition of mitochondrial respiration?Plant, Cell and Environment, 29, 1149–1158.CrossRefGoogle ScholarPubMed
Wohlfahrt, G., Anderson-Dunn, M., Bahn, M., et al. (2008b). Biotic, abiotic and management controls on the net ecosystem CO2 exchange of European mountain grassland ecosystems. Ecosystems, 11, 1338–1351.CrossRefGoogle Scholar
Wohlfahrt, G., Bahn, M., Tappeiner, U., et al. (2000). A model of whole plant gas exchange for herbaceous species from mountain grassland sites differing in land use. Ecological Modelling, 125, 173–201.CrossRefGoogle Scholar
Wohlfahrt, G., Fenstermaker, L.F. and Arnone, J.A. (2008a). Large annual net ecosystem CO2 uptake of a Mojave Desert ecosystem. Global Change Biology, 14, 1475–1487.CrossRefGoogle Scholar
Wolfe, K.H., Morden, C.W. and Palmer, J.D. (1992). Function and evolution of a minimal plastid genome from a nonphotosynthetic parasitic plant. Proc Natl Acad Sci USA, 89, 10648–10652.CrossRefGoogle ScholarPubMed
Wollman, F.A. (2001). State transitions reveal the dynamics and flexibility of the photosynthetic apparatus. EMBO Journal, 20, 3623–3630.CrossRefGoogle ScholarPubMed
Wong, C.E., Li, Y., Labbe, A., et al. (2006). Transcriptional profiling implicates novel interactions between abiotic stress and hormonal responses in Thellungiella, a close relative of Arabidopsis. Plant Physiology, 140, 1437–1450.CrossRefGoogle ScholarPubMed
Wong, S.C. and Osmond, C.B. (1991). Elevated atmospheric partial pressure of CO2 and plant growth. III. Interactions between Triticum aestivum (C3) and Echinochloa frumentacea (C4) during growth in mixed culture under different CO2, N nutrition and irradiance treatments with emphasis on belowground responses estimated using the delta 13C value of root biomass. Australian Journal of Plant Physiology, 18, 137–152.CrossRefGoogle Scholar
Wong, S.C., Cowan, I.R. and Farquhar, G.D. (1979). Stomatal conductance correlates with photosynthetic capacity. Nature, 282, 424–426.CrossRefGoogle Scholar
Wong, S.C., Cowan, I.R. and Farquhar, G.D. (1985). Leaf conductance relation to rate of CO2. 2. Effects of short-term exposures to different photon-flux densities. Plant Physiology, 78, 826–829.CrossRefGoogle Scholar
Wong, S.C., Cowan, I.R. and Farquhar, G.D. (1985). Leaf conductance in relation to the rate of CO2 assimilation. I. Influence of nitrogen nutrition, phosphorus nutrition, photon-flux density, and ambient partial pressure of CO2 during ontogeny. Plant Physiology, 78, 821–825.CrossRefGoogle Scholar
Wongwises, S., Pornsee, A. and Siroratsakul, E. (1999). Gas-wall shear stress distribution in horizontal stratified two-phase flow. International Communications in Heat and Mass Transfer, 26, 849–860.CrossRefGoogle Scholar
Woo, N.S., Badger, M.R., Pogson, B.J. (2008). A rapid, non-invasive procedure for quantitative assessment of drought survival using chlorophyll fluorescence. Plant Methods, 4, 27.CrossRefGoogle ScholarPubMed
Woodall, G.S., Dodd, I.C. and Stewart, G.R. (1998). Contrasting leaf development within the genus Syzygium. Journal of Experimental Botany, 49, 79–87.Google Scholar
Woodroffe, C.D. (1992). Mangrove sediments and geomorphology. Tropical Mangrove Ecosystems (eds Robertson, A.I. and Alongi, D.M.), American Geophysical Union, Washington, DC, USA, pp. 7–41.Google Scholar
Woodrow, L., Jiao, J., Tsujita, M.J., et al. (1989). Whole plant and leaf steady state gas exchange during ethylene exposure in Xanthium strumarium L. Plant Physiology, 90, 85–90.CrossRefGoogle ScholarPubMed
Woodruff, D.R., Meinzer, F.C., Lachenbruch, B., et al. (2009). Coordination of leaf structure and gas exchange along a height gradient in a tall conifer. Tree Physiology, 29, 261–272.CrossRefGoogle Scholar
Woods, H.A. and Smith, J.N. (2010). Universal model for water costs of gas exchange by animals and plants. Proceedings of the National Academy of Sciences USA, 107, 8469–8474.CrossRefGoogle ScholarPubMed
Woodward, F.I. (1979). The differential temperature responses of the growth of certain plant species from different altitudes. I. Growth analysis of Phleum alpinum L., P. bertolonii D.C., Sesleria albicans Kit. and Dactylis glomerata L. New Phytologist, 82, 385–395.CrossRefGoogle Scholar
Woodward, F.I. (1987a). Stomatal numbers are sensitive to increases in CO2 from pre-industrial levels. Nature, 327, 617–618.CrossRefGoogle Scholar
Woodward, F.I. (1987b). Climate and Plant Distribution. Cambridge University Press, Cambridge, UK.Google Scholar
Woodward, F.I. (1993). Plant responses to pasts concentrations of CO2. Vegetatio, 104, 105–155.Google Scholar
Woodward, F.I. and Kelly, C.K. (1995). The influence of CO2 concentration on stomatal density. The New Phytologist, 131, 311–327.CrossRefGoogle Scholar
Woodward, F.I., Körner, C. and Crabtree, R.C. (1986). The dynamics of leaf extension in plants with diverse altitudinal ranges. I. Field observations on temperature responses at one altitude. Oecologia, 70, 222–226.CrossRefGoogle ScholarPubMed
Wosfy, S.C., Goulden, M.L., Munger, J.W., et al. (1993). Net exchange of CO2 in a mid-latitude forest. Science, 260, 1314–1317.CrossRef
WRI (2005). World Resources Institute: Freshwater resources 2005. .
Wright, D.P., Baldwin, B.C., Shephard, M.C., et al. (1995). Source–sink relationships in wheat leaves infected with powdery mildew. I. Alterations in carbohydrate metabolism. Physiological and Molecular Plant Pathology, 47, 237–253.CrossRefGoogle Scholar
Wright, I.J., and Cannon, K. (2001). Relationships between leaf lifespan and structural defences in a low nutrient, sclerophyll flora. Functional Ecology, 15, 351–359.CrossRefGoogle Scholar
Wright, I.J., Reich, P.B. and Westoby, M. (2003). Least-cost input mixtures of water and nitrogen for photosynthesis. The American Naturalist, 161, 98–111.Google ScholarPubMed
Wright, I.J., Reich, P.B., Cornelissen, J.H.C., et al. (2005). Modulation of leaf economic traits and trait relationships by climate. Global Ecology and Biogeography, 14, 411–421.CrossRefGoogle Scholar
Wright, I.J., Reich, P.B., Westoby, M., et al. (2004). The world-wide leaf economics spectrum. Nature, 428, 821–827.CrossRefGoogle Scholar
Wright, S.I., Ness, R.W., Foxe, J.P., et al. (2008). Genomic consequences of outcrossing and selfing in plants. International Journal of Plant Sciences, 169, 105–118.CrossRefGoogle Scholar
Wu, X., Luo, Y., Weng, E., et al. (2009). Conditional inversion to estimate parameters from eddy flux observations. Journal of Plant Ecology, 2, 55–68.CrossRefGoogle Scholar
Wullschleger, S.D. (1993). Biochemical limitations to carbon assimilation in C3 plants-a retrospective analysis of the A/Ci curves from 109 species. Journal of Experimental Botany, 44, 907–920.CrossRefGoogle Scholar
Wullschleger, S.D. and Norby, R.J. (2001). Sap velocity and canopy transpiration in a sweetgum stand exposed to free-air CO2 enrichment (FACE). New Phytologist, 25, 489–498.CrossRefGoogle Scholar
Wullschleger, S.D., Tschaplinski, T.J. and Norby, R.J. (2002). Plant water relations at elevated CO2 – implications for water-limited environments. Plant, Cell and Environment, 25, 319–331.CrossRefGoogle ScholarPubMed
Wünsche, J.N. and Palmer, J.W. (1997). Portable through-flow cuvette system for measuring whole-canopy gas exchange of apple trees in the field. HortScience, 32, 653–658.Google Scholar
Wyrich, R., Dressen, U., Brockman, S., et al. (1998). The molecular basis of C4 photosynthesis in sorghum: isolation, characterization and RFLP mapping of mesophyll- and bundle-sheath-specific cDNAs as obtained by differential screening. Plant Molecular Biology, 37, 319–335.CrossRefGoogle ScholarPubMed
Xie, J., Li, Y., Zhai, C., et al. (2008). CO2 absorption by alkaline soils and its implication to the global carbon cycle. Environmental Geology, 56, 953–61.CrossRefGoogle Scholar
Xin, Z. and Browse, J. (2000). Cold comfort farm: the acclimation of plants to freezing temperatures. Plant Cell Environment, 23, 893–902.CrossRefGoogle Scholar
Xiong, F.S., Mueller, E.C. and Day, T.A. (2000). Photosynthetic and respiratory acclimation and growth response of Antarctic vascular plants to contrasting temperature regimes. American Journal of Botany, 87, 700–710.CrossRefGoogle ScholarPubMed
Xiong, F.S., Ruhland, C.T. and Day, T.A. (1999). Photosynthetic temperature response of the Antarctic vascular plants Colobanthus quitensis and Deschampsia antarctica. Physiologia Plantarum, 106, 276–286.CrossRefGoogle Scholar
Xiong, J. and Bauer, C.E. (2002). Complex evolution of photosynthesis. Annual Review of Plant Biology, 53, 503–521.CrossRefGoogle ScholarPubMed
Xu, C., Fan, J., Riekhof, W., et al. (2003). A permease-like protein involved in ER to thylakoid lipid transfer in Arabidopsis. EMBO Journal, 22, 2370–2379.CrossRefGoogle ScholarPubMed
Xu, H.L., Gauthier, L., Desjardins, Y., et al. (1997). Photosynthesis in leaves, stem and petioles of greenhouse-grown tomato plants. Photosynthetica, 33, 113–123.CrossRefGoogle Scholar
Xu, H.X., Weng, X.Y. and Yang, Y. (2007). Effect of phosphorus deficiency on the photosynthetic characteristics of rice plants. Russian Journal of Plant Physiology, 54, 741–748.CrossRefGoogle Scholar
Xu, L. and Baldocchi, D.D. (2004). Seasonal variation in carbon dioxide exchange over a Mediterranean annual grassland in California. Agricultural and Forest Meteorology, 123, 79–96.CrossRefGoogle Scholar
Xu, L.K. and Baldocchi, D.D. (2003). Seasonal trends in photosynthetic parameters and stomatal conductance of blue oak (Quercus douglasii) under prolonged summer drought and high temperature. Tree physiology, 23, 865–877.CrossRefGoogle ScholarPubMed
Xu, L.K., Matista, A.A. and Hsiao, T.C. (1999). A technique for measuring CO2 and water vapour profiles within and above plant canopies over short periods. Agricultural and Forest Meteorology, 94, 1–12.CrossRefGoogle Scholar
Xu, S. (2002). QTL analysis in plants. In: Quantitative Trait Loci: Methods and Protocols (eds Camp, N. and Cox, A.), Humana Press, Totowa, USA, pp. 283–310.Google Scholar
Xu, Z. and Zhou, G. (2008). Responses of leaf stomatal density to water status and its relationship with photosynthesis in a grass. Journal of Experimental Botany, 59, 3317–3325.CrossRefGoogle Scholar
Xu, Z., Zhou, G. and Shimizu, H. (2009). Are plant growth and photosynthesis limited by pre-drought following rewatering in grass?Journal of Experimental Botany, 60, 3737–3749.CrossRefGoogle ScholarPubMed
Yabuta, Y., Mieda, T., Rapolu, M., et al. (2007). Light regulation of ascorbate biosynthesis is dependent on the photosynthetic electron transport chain but independent of sugars in Arabidopsis. Journal of Experimental Botany, 58, 2661–2671.CrossRefGoogle ScholarPubMed
Yakir, D. (1998). Oxygen-18 of leaf water: a crossroad for plant-associated isotopic signals. In: Stable Isotopes. Integration of Biological, Ecological and Geochemical Processes (ed. Griffiths, H.), BIOS Scientific Publishers, Oxford, UK, pp. 147–168.Google Scholar
Yakir, D. (2003). The stable isotopic composition of atmospheric CO2. Treatise on Geochemistry, 4, 175–212.CrossRefGoogle Scholar
Yakir, D. and Wang, X.F. (1996). Fluxes of CO2and water between terrestrial vegetation and the atmosphere estimated from isotope measurements. Nature, 380, 515–517.CrossRefGoogle Scholar
Yakir, D., De Niro, M.J. and Rundel, P.W. (1989). Isotopic inhomogeneity of leaf water:evidence and implications for the use of isotopic signal transduced by plants. Geochimica et Cosmochimica Acta, 53, 2769–2773.CrossRefGoogle Scholar
Yamaguchi-Shinozaki, K. and Shinozaki, K. (2006). Responses and tolerance to dehydration and cold stresses. Annual Review of Plant Biology, 57, 781–803.CrossRefGoogle ScholarPubMed
Yamamoto, H.Y., Kamite, L. and Wang, Y.Y. (1972). An ascorbate-induced absorbance change in chloroplasts from violaxanthin de-epoxidation. Plant Physiology, 49, 224–228.CrossRefGoogle ScholarPubMed
Yamane, Y., Kashino, Y., Koike, H., et al. (1997). Increases in the fluorescence F0 level and reversible inhibition of photosystem II reaction center by high-temperature treatments in higher plants. Photosynthesis Research, 52, 57–64.CrossRefGoogle Scholar
Yamane, Y., Kashino, Y., Koike, H., et al. (1998). Effects of high temperatures on the photosynthetic systems in spinach: Oxygen-evolving activities, fluorescence characteristics and the denaturation process. Photosynthesis Research, 57, 51–59.CrossRefGoogle Scholar
Yamasaki, H. (1997). A function of colour. Trends in Plant Science, 2, 7–9.CrossRefGoogle Scholar
Yamashita, N., Ishida, A., Kushima, H., et al. (2000). Acclimation to sudden increase in light favouring an invasive over native trees in subtropical islands, Japan. Oecologia, 125, 412–419.CrossRefGoogle Scholar
Yamashita, N., Koike, N. and Ishida, A. (2002). Leaf ontogenetic dependence of light acclimation in invasive and native subtropical trees of different successional status. Plant Cell Environment, 25, 1341–1356.CrossRefGoogle Scholar
Yamori, W., Nagai, T. and Makino, A. (2011). The rate-limiting step for CO2 assimilation at different temperatures is influenced by the leaf nitrogen content in several C3 crop species. Plant, Cell and Environment, 34, 764–777.CrossRefGoogle Scholar
Yamori, W., Noguchi, K., Hanba, Y.T., et al. (2006). Effects of internal conductance on the temperature dependence of the photosynthetic rate in spinach leaves from contrasting growth temperatures. Plant and Cell Physiology, 47, 1069–1080.CrossRefGoogle ScholarPubMed
Yamori, W., Noguchi, K.O. and Terashima, I. (2005). Temperature acclimation of photosynthesis in spinach leaves: analyses of photosynthetic components and temperature dependencies of photosynthetic partial reactions. Plant, Cell and Environment, 28, 536–547.CrossRefGoogle Scholar
Yan, K., Chen, N., Qu, Y.Y., et al. (2008). Overexpression of sweet pepper glycerol-3-phosphate acyltransferase gene enhanced thermotolerance of photosynthetic apparatus in transgenic tobacco. Journal of Integrative Plant Biology, 50, 613–621.CrossRefGoogle ScholarPubMed
Yan, Y., Stolz, S., Chetelat, A., et al. (2007). A downstream mediator in the growth repression limb of the jasmonate pathway. The Plant Cell, 19, 2470–2483.CrossRefGoogle ScholarPubMed
Yang, C., Guo, R., Jie, F., et al. (2007). Spatial analysis of Arabidopsis thaliana gene expression in response to Turnip mosaic virus infection. Molecular Plant-Microbe Interactions, 20, 358–370.CrossRefGoogle ScholarPubMed
Yang, C.W., Xu, H.H., Wang, L.L., et al. (2009). Comparative effects of salt-stress and alkali-stress on the growth, photosynthesis, solute accumulation, and ion balance of barley plants. Photosynthetica, 47, 79–86.CrossRefGoogle Scholar
Yang, X., Liang, Z., Wen, X., et al. (2008). Genetic engineering of the biosynthesis of glycinebetaine leads to increased tolerance of photosynthesis to salt stress in transgenic tobacco plants. Plant Molecular Biology, 66, 73–86.CrossRefGoogle ScholarPubMed
Yang, X., Wen, X., Gong, H., et al. (2007). Genetic engineering of the biosynthesis of glycinebetaine enhances thermotolerance of photosystem II in tobacco plants. Planta, 225, 719–733.CrossRefGoogle ScholarPubMed
Yang, X.J., Guignard, G., Thévenard, F., et al. (2009). Leaf cuticle ultrastructure of Pseudofrenelpsis dalatzenis (Chow et Tsao) Cao ex Zhou (Cheirolepidaceae) from the Lower Cretaceous Dalazi formation of Jilin, China. Review of Paleobotany and Palynology, 153, 8–18.Google Scholar
Yano, S. and Terashima, I. (2001). Separate localization of light signal perception for sun or shade type chloroplast and palisade tissue differentiation in Chenopodium album. Plant and Cell Physiology, 42, 1303–1310.CrossRefGoogle ScholarPubMed
Yano, S. and Terashima, I. (2004). Developmental processes of sun and shade leaves in Chenopodium album L. Plant Cell Environment, 27, 781–793.CrossRefGoogle Scholar
Yasumura, Y., Hikosaka, K. and Hirose, T. (2007). Nitrogen resorption and protein degradation during leaf senescence in Chenopodium album grown in different light and nitrogen conditions. Functional Plant Biology, 34, 409–417.CrossRefGoogle Scholar
Ye, Z.P. and Yu, Q. (2008). A coupled model of stomatal conductance and photosynthesis for winter wheat. Photosynthetica, 46, 637–640.CrossRefGoogle Scholar
Yemm, E.W. and Willis, A.J. (1954). Chlorophyll and photosynthesis in stomatal guard cells. Nature, 173, 726.CrossRefGoogle ScholarPubMed
Yensen, N.P., Fontes, M.R., Glenn, E.P., et al. (1981). New salt tolerant crops for the Sonoran Desert. Desert Plants, 3, 111–118.Google Scholar
Yeoh, H.H., Badger, M.R. and Watson, L. (1980). Variations in Km(CO2) of ribulose-1,5-bisphosphate carboxylase among grasses. Plant Physiology, 66, 1110–1112.CrossRefGoogle Scholar
Yin, X. and Struik, P.C. (2009). Theoretical reconsiderations when estimating the mesophyll conductance to CO2 diffusion in leaves of C3 plants by analysis of combined gas exchange and chlorophyll fluorescence measurements. Plant, Cell and Environment, 32, 1513–1524.CrossRefGoogle Scholar
Yin, X., Goudriaan, J., Lantinga, E.A., et al. (2003). A flexible sigmoid function of determinate growth. Annals of Botany, 91, 361–371.CrossRefGoogle ScholarPubMed
Yin, X., Harbinson, J. and Struik, P.C. (2006). Mathematical review of literature to assess alternative electron transports and interphotosystem excitation partitioning of steady state C3 photosynthesis under limiting light. Plant, Cell and Environment, 29, 1771–1782.CrossRefGoogle ScholarPubMed
Yin, X., Sun, Z., Struik, P.C. et al. (2011). Evaluating a new method to estimate the rate of leaf respiration in the light by analysis of combined gas exchange and chlorophyll fluorescence measurements. Journal of Experimental Botany, 62, 3489–3499.CrossRefGoogle Scholar
Yin, X.Y., Struik, P.C., Romero, P., et al. (2009). Using combined measurements of gas exchange and chlorophyll fluorescence to estimate parameters of a biochemical C3 photosynthesis model: a critical appraisal and a new integrated approach applied to leaves in a wheat (Triticum aestivum) canopy. Plant, Cell and Environment, 32, 448–464.CrossRefGoogle Scholar
Ying, J., Lee, E.A. and Tollenaar, M. (2000). Response of maize leaf photosynthesis to low temperature during the grain-filling period. Field Crops Research, 68, 87–96.CrossRefGoogle Scholar
Yiotis, C., Manetas, Y. and Psaras, K.G. (2006). Leaf and green stem anatomy of the drought deciduous Mediterranean shrub Calicotome villosa (Poiret) Link. (Leguminosae). Flora, 201, 102–107.CrossRefGoogle Scholar
Yoder, B.J., Ryan, M.G., Waring, R. H, et al. (1994). Evidence of reduced photosynthetic rates in old trees. Forest Science, 40, 513–527.Google Scholar
Yokota, A. and Calvin, D.T. (1985). Ribulose bisphosphate carboxylase/oxygenase content determined with [14C]carboxypentitol bisphosphate in plants and algae. Plant Physiology, 77, 735–739.CrossRefGoogle Scholar
Yokthongwattana, K. and Melis, A. (2006). Photoinhibition and recovery in oxygenic photosynthesis: mechanism of photosystem II damage and repair cycle. In: Photoprotection, Photoinhibtion, Gene Regulation, and Environment (eds Demmig-Adams, B., Adams, WW. and Mattoo, A.K.), Springer, Dordrecht, Germany, pp. 175–191.Google Scholar
Yoo, S.D., Greer, D.H., Laing, W.A., et al. (2003). Changes in photosynthetic efficiency and carotenoid composition in leaves of white clover at different developmental stages. Plant Physiology and Biochemestry, 41, 887–893.CrossRefGoogle Scholar
Yoshida, K., Terashima, I.. and Noguchi, K. (2007). Up-regulation of mitochondrial alternative oxidase concomitant with chloroplast over-reduction by excess light. Plant and Cell Physiology, 48, 606–614.CrossRefGoogle ScholarPubMed
Yoshida, K., Watanabe, C., Kato, Y., et al. (2008). Influence of chloroplastic photo-oxidative stress on mitochondrial alternative oxidase capacity and respiratory properties: a case study with Arabidopsis yellow variegated 2. Plant and Cell Physiology, 49, 592–603.CrossRefGoogle ScholarPubMed
Yoshida, S. (2003). Molecular regulation of leaf senescence. Current Opinion in Plant Biology, 6, 79–84.CrossRefGoogle ScholarPubMed
Yoshimoto, M., Oue, H. and Kobayashi, K. (2005). Energy balance and water use efficiency of rice canopies under free-air CO2 enrichment. Agricultural and Forest Meteorology, 133, 226–246.CrossRefGoogle Scholar
Yoshinaga, K. and Kugita, M. (2004). Evolution of early land plants : insights from chloroplast genomic sequences and RNA editing. Endocytobiosis Cell Research, 15, 256–271.Google Scholar
Yruela, I., Alfonso, M., Ortiz de Zarate, I., et al. (1993). Precise location of the Cu(II)-inhibitory binding site in higher plant and bacterial photosynthetic reaction centers as probed by light-induced absorption changes. Journal of Biological Chemistry, 268, 1684–1689.Google ScholarPubMed
Ytterberg, A.J., Peltier, J.B. and van Wijk, K.J. (2006). Protein profiling of plastoglobules in chloroplasts and chromoplasts. A surprising site for differential accumulation of metabolic enzymes. Plant Physiology, 140, 984–997.CrossRefGoogle ScholarPubMed
Yu, M., Xie, Y. and Zhang, X. (2005). Quantification of intrinsic water use efficiency along a moisture gradient in northeastern China. Journal of Environmental Quality, 34, 1311–1318.CrossRefGoogle ScholarPubMed
Yu, Q., Osborne, L. and Rengel, Z. (1998). Micronutrient deficiency changes activities of superoxide dismutase and ascorbate peroxidase in tobacco plants. Journal of Plant Nutrition, 21, 1427–1437.CrossRefGoogle Scholar
Yukawa, J. and Tsuda, K. (1986). Leaf longevity of Quercus glauc Thunb., with reference to the influence of gall formation by Contarinia sp. (Diptera: Cecidomyiidae) on the early mortality of fresh leaves. Memoirs of the Faculty of Agriculture, Kagoshima University, 22, 73–77.
Zabel, B., Hawes, P., Stuart, H., et al. (1999). Construction and engineering of a created environment: overview of the Biosphere 2 closed system. Ecological Engineering, 13, 43–63.CrossRefGoogle Scholar
Zachos, J., Pagani, M., Sloan, L., et al. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693.CrossRefGoogle ScholarPubMed
Zaitlin, M. and Jagendorf, A.T. (1960). Photosynthetic phosphorylation and Hill reaction activities of chloroplasts isolated from plants infected with tobacco mosaic virus. Virology, 12, 477–486.CrossRefGoogle ScholarPubMed
Zaluar, H.L.T. and Scarano, F.R. (2000). Facilitação em restingas de moitas: um sécolo de buscas por espécies focais. In: Ecologia de Restingas e Lagoas Costeiras (eds Esteves, F.A. and Lacereda, L.D.), NUPEM-UFRJ, Rio de Janeiro, Brazil, pp. 3–23.
Zamolodchikov, D.G., Karelin, D.V., Ivaschenko, A. I., et al. (2003). CO2 flux measurements in Russian Far East tundra using eddy covariance and closed chamber techniques. Tellus, 55B, 879–892.CrossRefGoogle Scholar
Zandonadi, R.S., Pinto, F.A.C., Sena, D.G., et al. (2005). Identification of lesser cornstalk borer-attacked maize plants using infrared images. Biosystems engineering, 91, 433–439.CrossRefGoogle Scholar
Zangerl, A.R., Berenbaum, M.R., DeLucia, E.H., et al. (2003). Fathers, fruits and photosynthesis: pollen donor effects on fruit photosynthesis in wild parsnip. Ecology Letters, 6, 966–970.CrossRefGoogle Scholar
Zangerl, A.R., Hamilton, J.G., Miller, T.J., et al. (2002). Impact of folivory on photosynthesis is greater than the sum of its holes. Proceeding of the National Academy of Science USA, 99, 1088–1091.CrossRefGoogle Scholar
Zankel, K. L. (1973). Rapid fluorescence changes observed in chloroplasts: their relationship to the O2 evolving system. Biochimica et Biophysica Acta, 325, 138–148.CrossRefGoogle ScholarPubMed
Zaragoza-Castells, J., Sánchez-Gómez, D., Hartley, I.P., et al. (2008). Climate-dependent variations in leaf respiration in a dry land, low productivity Mediterranean forest: the importance of acclimation in both high-light and shaded habitats. Functional Ecology, 22, 172–184.Google Scholar
Zaragoza-Castells, J., Sánchez-Gómez, D., Valladares, F., et al. (2007). Does growth irradiance affect temperature dependence and thermal acclimation of leaf respiration? Insights from a Mediterranean tree with long-lived leaves. Plant, Cell and Environment, 30, 820–833.CrossRefGoogle Scholar
Zarco-Tejada, P.J.Miller, J.R.Mohammed, G.H., et al. (2000b). Chlorophyll fluorescence effects on vegetation apparent relectance. II. Laboratory and airborne canopy level measurements with hyperspectral data. Remote Sensing of Environment, 74, 596–608.CrossRefGoogle Scholar
Zarco-Tejada, P.J.Miller, J.R.Mohammed, G.H., et al. (2000a). Chlorophyll fluorescence effects on vegetation apparent relectance. I. Leaf-level measurements and model simulation. Remote Sensing of Environment, 74, 582–595.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Berjón, A., López-Lozano, R., et al. (2005a). Assessing vineyard condition with hyperspectral indices: Leaf and canopy reflectance simulation in a row-structured discontinuous canopy. Remote Sensing of the Environment, 99, 271–287.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Miller, J.R., Mohammed, G.H., et al. (2001). Scaling-up and model inversion methods with narrow-band optical indices for chlorophyll content estimation in closed forest canopies with hyperspectral data. IEEE Transactions on Geoscience and Remote Sensing, 39, 1491–1507.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Pushnik, J.C., Dobrowski, S., et al. (2003b). Steady state chlorophyll a fluorescence detection from canopy derivative reflectance and double-peak red-edge effects. Remote Sensing of Environment, 84, 283–294.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Rueda, C.A. and Ustin, S.L. (2003a). Water content estimation in vegetation with MODIS reflectance data and model inversion methods. Remote Sensing of the Environment, 85, 109–124.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Ustin, S.L. and Whiting, M.L. (2005b). Temporal and spatial relationships between within-field yield variability in cotton and high-spatial hyperspectral remote sensing imagery. Agronomy Journal, 97, 641–653.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Berni, J.A.J., Suárez, L., et al. (2009). Imaging chlorophyll fluorescence with an airborne narrow-bannd multispectral camera for vegetation detection. Remote Sensing of the Environment, 113, 1262–1275.CrossRefGoogle Scholar
Zarco-Tejada, P.J., Miller, J.R., Pedros, R., et al. (2006). FluorMODgui V3.0 : A graphic user interface for the spectral simulation of leaf and canopy chlorophyll fluorescence. Computers and Geosciences, 32, 5, 577–591.CrossRefGoogle Scholar
Zarter, C.R., Demming-Adams, B., Ebbert, V., et al. (2006). Photosynthetic capacity and light harvesting efficiency during the winter-to-spring transition in subalpine conifers. New Phytologist, 172, 283–292.CrossRefGoogle Scholar
Zavaleta, E.S. (2001). Influences of climate and atmospheric changes on plant diversity and ecosystem function in a California grassland. Ph.D. thesis, Stanford University, Stanford, California.
Zawada, D.G. (2003). Image processing of underwater multispectral imagery IEEE. Journal of Oceanic Engineering, 28, 583–594.CrossRefGoogle Scholar
Zeeman, S.C., Smith, S.M. and Smith, A.M. (2004). The breakdown of starch in leaves. New Phytologist, 163, 247–261.CrossRefGoogle Scholar
Zeiger, E., Talbott, L.D., Frechilla, S., et al. (2002). The guard cell chloroplast: a perspective for the twenty first century. The New Phytologist, 153, 415–424.CrossRefGoogle Scholar
Zeliou, K., Manetas, Y. and Petropoulou, Y. (2009). Transient winter leaf reddening in Cistus creticus characterizes weak (stress-sensitive) individuals, yet anthocyanins cannot alleviate the adverse effects on photosynthesis. Journal of Experimental Botany, 60, 3031–3042.CrossRefGoogle ScholarPubMed
Zelitch, I. (1982). The close relationship between net photosynthesis and crop yield. BioScience, 32, 792–802.CrossRefGoogle Scholar
Zellnig, G., Perktold, A. and Zechmann, B. (2010). Fine structural quantification of drought-stressed Picea abies (L.) organelles based on 3D reconstructions. Protoplasma, 243, 129–136.CrossRefGoogle ScholarPubMed
Zeuthen, J., Mikkelsen, J.N., Paluda-Müller, G., et al. (1997) .Effects of increased UV-B radiation and elevated levels of tropospheric ozone on physiological processes in European beech (Fagus sylvatica). Physiologia Plantarum, 100, 281–290.CrossRefGoogle Scholar
Zha, T.S., Xing, Z.S., Wang, K.Y., et al. (2007). Total and component carbon fluxes of a Scots pine ecosystem from chamber measurements and eddy covariance. Annals of Botany, 99, 345–353.CrossRefGoogle ScholarPubMed
Zhang, B.Z., Kang, S.Z., Li, F.S., et al. (2008a). Comparison of three evapotranspiration models to Bowen ratio-energy balance method for a vineyard in an and desert region of northwest China. Agricultural and Forest Meteorology, 148, 1629–1640.CrossRefGoogle Scholar
Zhang, J., Griffis, T.J. and Baker, J.M. (2006). Using continuous stable isotope measurements to partition net ecosystem CO2 exchange. Plant, Cell and Environment, 29, 483–496.CrossRefGoogle ScholarPubMed
Zhang, J.L., Meng, L.Z. and Cao, K.F. (2009d). Sustained diurnal photosynthetic depression in uppermost-canopy leaves of four dipterocarp species in the rainy and dry seasons: does photorespiration play a role in photoprotection?Tree Physiology, 29, 217–228.CrossRefGoogle ScholarPubMed
Zhang, L.M., Lohmann, C., Prändl, R., et al. (2003). Heat stress-dependent DNA binding of Arabidopsis heat shock transcription factor HSF1 to heat shock gene promoters in Arabidopsis suspension culture cells in vivo. Biological Chemistry, 384, 959–963.CrossRefGoogle ScholarPubMed
Zhang, N., Zhao, Y.S. and Yu, G.R. (2009b). Simulated annual carbon fluxes of grassland ecosystems in extremely arid conditions. Ecological Research, 24, 185–206.CrossRefGoogle Scholar
Zhang, R. and Sharkey, T.D. (2009). Photosynthetic electron transport and proton flux under moderate heat stress. Photosynthesis Research, 100, 29–43.CrossRefGoogle ScholarPubMed
Zhang, R., Cruz, J.A., Kramer, D.M., et al. (2009c). Moderate heat stress reduces the pH component of the transthylakoid proton motive force in light-adapted intact tobacco leaves. Plant, Cell and Environment, 32, 1538–1547.CrossRefGoogle Scholar
Zhang, S.B., Hu, H. and Li, Z.R. (2008b). Variation of photosynthetic capacity with leaf age in an alpine orchid, Cypripedium flavum. Acta Physiologiae Plantarum, 30, 381–388.CrossRefGoogle Scholar
Zhang, X., Wollenweber, B., Jiang, D., et al. (2008c). Water deficits and heat shock effects on photosynthesis of a transgenic Arabidopsis thaliana constitutively expressing ABP9, a bZIP transcription factor. Journal of Experimental Botany, 59, 839–848.CrossRefGoogle ScholarPubMed
Zhang, Y., Primavesi, L.F., Jhurreea, D., et al. (2009a). Inhibition of Snf1-related protein kinase (SnRK1) activity and regulation of metabolic pathways by trehalose 6-phosphate. Plant Physiology, 149, 1860–1871.CrossRefGoogle ScholarPubMed
Zhao, M. and Running, S.W. (2010). Drought-induced reduction in global terrestrial net primary production from 2000 through 2009. Science, 329, 940–943.CrossRefGoogle ScholarPubMed
Zhao, S. and Blumwald, E. (1998). Changes in oxidation-reduction state and antioxidant enzymes in the roots of jack pine seedlings during cold acclimation. Physiologia Plantarum, 104, 134–142.CrossRefGoogle Scholar
Zheng, C., Jiang, D., Liu, F., et al. (2009). Effects of salt and waterlogging stresses and their combination on leaf photosynthesis, chloroplast ATP synthesis, and antioxidant capacity in wheat. Plant Science, 176, 575–582.CrossRefGoogle ScholarPubMed
Zhou, J., Wang, X., Jiao, Y., et al. (2007b). Global genome expression analysis of rice in response to drought and high-salinity stresses in shoot, flag leaf, and panicle. Plant Molecular Biology, 63, 591–608.CrossRefGoogle ScholarPubMed
Zhou, X.L., Peng, C.H., Dang, Q.L., et al. (2008). Simulating carbon exchange in Canadian Boreal forests I. Model structure, validation, and sensitivity analysis. Ecological Modelling, 219, 287–299.CrossRefGoogle Scholar
Zhou, Y., Lam, H.M. and Zhang, J. (2007a). Inhibition of photosynthesis and energy dissipation induced by water and high light stresses in rice. Journal of Experimental Botany, 58, 1207–1217.CrossRefGoogle Scholar
Zhu, J., Goldstein, G. and Bartholomew, D.P. (1999). Gas exchange and carbon isotope composition of Ananas comosus in response to elevated CO2 and temperature. Plant, Cell and Environment, 22, 999–1007.CrossRefGoogle Scholar
Zhu, J., Talbott, L.D., Jin, X., et al. (1998). The stomatal response to CO2 is linked to changes in guard cell zeaxanthin. Plant, Cell and Environment, 21, 813–820.CrossRefGoogle Scholar
Zhu, X.G., de Sturler, E. and Long, S.P. (2007). Optimizing the distribution of resources between enzymes of carbon metabolism can dramatically increase photosynthetic rate: A numerical simulation using an evolutionary algorithm. Plant Physiology, 145, 513–526.CrossRefGoogle ScholarPubMed
Zhu, X.G., Long, S.P. and Ort, D.R. (2008). What is the maximum efficiency with which photosynthesis can convert solar energy into biomass?Current opinion in biotechnology, 19, 153–159.CrossRefGoogle ScholarPubMed
Zhu, X.G., Long, S.P. and Ort, D.R. (2010). Improving photosynthetic efficiency for greater yield. Annual Review of Plant Biology, 61, 235–261.CrossRefGoogle ScholarPubMed
Zhu, X.G., Portis, A.R. and Long, S.P. (2004). Would transformation of C3 crop plants with foreign Rubisco increase productivity? A computational analysis extrapolating from kinetic properties to canopy photosynthesis. Plant, Cell and Environment, 27, 155–165.CrossRefGoogle Scholar
Zielinski, R.E., Werneke, J.M. and Jenkins, M.E. (1989). Coordinate expression of rubisco activase and rubisco during barley leaf cell development. Plant Physiology, 90, 516–521.CrossRefGoogle ScholarPubMed
Ziska, L.H. and Bunce, J.A. (1994). Direct and indirect inhibition of single leaf respiration by elevated CO2 concentrations: interaction with temperature. Physiologia Plantarum, 90, 130–138.CrossRefGoogle Scholar
Ziska, L.H. and Bunce, J.A. (1995). Growth and photosynthetic response of three soybean cultivars to simultaneous increases in growth temperature and CO2. Physiologia Plantarum, 94, 575–584.CrossRefGoogle Scholar
Ziska, L.H. and Bunce, J.A. (1998). The influence of increasing growth temperatue and CO2 concentration on the ratio of respiration to photosynthesis in soybean seedlings. Global Change Biology, 4, 637–643.CrossRefGoogle Scholar
Zotz, G. (2007). Johansson revisited: the spatial structure of epiphyte assemblages. Journal of Vegetation Science, 18, 123–130.CrossRefGoogle Scholar
Zotz, G. and Winter, K. (1993a). Short-term regulation of Crassulacean acid metabolism activity in a tropical hemiepiphyte, Clusia uvitana. Plant Physiology, 102, 835–841.CrossRefGoogle Scholar
Zotz, G. and Winter, K. (1993b). Short-term photosynthesis measurements predict leaf carbon balance in tropical rain-forest canopy plants. Planta, 191, 409–412.CrossRefGoogle Scholar
Zotz, G. and Winter, K. (1994a). A one-year study on carbon, water, and nutrient relationships in a tropical C3-CAM hemiepiphyte, Clusia uvitana. New Phytologist, 127, 45–60.CrossRefGoogle Scholar
Zotz, G. and Winter, K. (1994b). Annual carbon balance and nitogen use efficiency in tropical C3 and CAM epiphytes. New Phytologist, 126, 481–492.CrossRefGoogle Scholar
Zotz, G., Patiño, S. and Tyree, M.T. (1997). CO2 gas exchange and the occurrence of CAM in tropical woody hemiepiphytes. Flora, 192, 143–150.CrossRefGoogle Scholar
Zotz, G., Vollrath, B. and Schmidt, G. (2003). Carbon relations of fruits of epiphytic orchids. Flora, 198, 98–105.CrossRefGoogle Scholar
Zou, J., Rodriguez-Sas, S., Aldea, M., et al. (2005). Expression profiling soybean response to Pseudomonas syringae reveals new defense-related genes and rapid HR-Specific downregulation of photosynthesis. Molecular Plant-Microbe Interactions, 18, 1161–1174.CrossRefGoogle ScholarPubMed

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 Jaume Flexas, Universitat de les Illes Balears, Palma de Mallorca, Francesco Loreto, Hipólito Medrano, Universitat de les Illes Balears, Palma de Mallorca
  • Book: Terrestrial Photosynthesis in a Changing Environment
  • Online publication: 05 March 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9781139051477.042
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 Jaume Flexas, Universitat de les Illes Balears, Palma de Mallorca, Francesco Loreto, Hipólito Medrano, Universitat de les Illes Balears, Palma de Mallorca
  • Book: Terrestrial Photosynthesis in a Changing Environment
  • Online publication: 05 March 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9781139051477.042
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 Jaume Flexas, Universitat de les Illes Balears, Palma de Mallorca, Francesco Loreto, Hipólito Medrano, Universitat de les Illes Balears, Palma de Mallorca
  • Book: Terrestrial Photosynthesis in a Changing Environment
  • Online publication: 05 March 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9781139051477.042
Available formats
×