Journal articles on the topic 'Desiccation toleranc'

To see the other types of publications on this topic, follow the link: Desiccation toleranc.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Desiccation toleranc.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Englert, John M., Keith Warren, Leslie H. Fuchigami, and Tony H. H. Chen. "Antidesiccant Compounds Improve the Survival of Bare-root Deciduous Nursery Trees." Journal of the American Society for Horticultural Science 118, no. 2 (March 1993): 228–35. http://dx.doi.org/10.21273/jashs.118.2.228.

Full text
Abstract:
Desiccation stress during the postharvest handling of bare-root deciduous trees can account for dieback and poor regrowth after transplanting. Desiccation tolerance of three bare-root deciduous hardwood species was determined at monthly harvest intervals from Sept. 1990 through Apr. 1991. Among the three species tested red oak (Quercus rubra L.) was most tolerant to desiccation, followed by Norway maple (Acer platanoides L.) and Washington hawthorn (Crataegus phaenopyrum Medic.). Maximum desiccation tolerance of all three species occurred during the January and February harvests. Of 20 film-forming compounds tested, the antidesiccant Moisturin was the most effective in reducing water loss from bare-root trees during desiccation stress and in improving survival and plant performance during re-establishment in the laboratory, greenhouse, and field. Moisturin-treated plants lost up to 80% less water than untreated plants. Washington hawthorn seedlings treated with Moisturin before severe desiccating conditions had the highest survival, lowest dieback/plant, and highest root growth ratings. The results indicate that Moisturin is an effective means of overcoming postharvest desiccation stress in desiccation sensitive plants, such as Washington hawthorn.
APA, Harvard, Vancouver, ISO, and other styles
2

Wolkers, Willem F., and Folkert A. Hoekstra. "In situFTIR Assessment of Desiccation-Tolerant Tissues." Spectroscopy 17, no. 2-3 (2003): 297–313. http://dx.doi.org/10.1155/2003/831681.

Full text
Abstract:
This essay shows how Fourier transform infrared (FTIR) microspectroscopy can be applied to study thermodynamic parameters and conformation of endogenous biomolecules in desiccation-tolerant biological tissues. Desiccation tolerance is the remarkable ability of some organisms to survive complete dehydration. Seed and pollen of higher plants are well known examples of desiccation-tolerant tissues. FTIR studies on the overall protein secondary structure indicate that during the acquisition of desiccation tolerance, plant embryos exhibit proportional increases inα-helical structures and thatµ-sheet structures dominate upon drying of desiccation sensitive-embryos. During ageing of pollen and seeds, the overall protein secondary structure remains stable, whereas drastic changes in the thermotropic response of membranes occur, which coincide with a complete loss of viability. Properties of the cytoplasmic glassy matrix in desiccation-tolerant plant organs can be studied by monitoring the position of the OH-stretching vibration band of endogenous carbohydrates and proteins as a function of temperature. By applying these FTIR techniques to maturation-defective mutant seeds ofArabidopsis thalianawe were able to establish a correlation between macromolecular stability and desiccation tolerance. Taken together,in situFTIR studies can give unique information on conformation and stability of endogenous biomolecules in desiccation-tolerant tissues.
APA, Harvard, Vancouver, ISO, and other styles
3

Pardo, Jeremy, Ching Man Wai, Hannah Chay, Christine F. Madden, Henk W. M. Hilhorst, Jill M. Farrant, and Robert VanBuren. "Intertwined signatures of desiccation and drought tolerance in grasses." Proceedings of the National Academy of Sciences 117, no. 18 (April 23, 2020): 10079–88. http://dx.doi.org/10.1073/pnas.2001928117.

Full text
Abstract:
Grasses are among the most resilient plants, and some can survive prolonged desiccation in semiarid regions with seasonal rainfall. However, the genetic elements that distinguish grasses that are sensitive versus tolerant to extreme drying are largely unknown. Here, we leveraged comparative genomic approaches with the desiccation-tolerant grass Eragrostis nindensis and the related desiccation-sensitive cereal Eragrostis tef to identify changes underlying desiccation tolerance. These analyses were extended across C4 grasses and cereals to identify broader evolutionary conservation and divergence. Across diverse genomic datasets, we identified changes in chromatin architecture, methylation, gene duplications, and expression dynamics related to desiccation in E. nindensis. It was previously hypothesized that transcriptional rewiring of seed desiccation pathways confers vegetative desiccation tolerance. Here, we demonstrate that the majority of seed-dehydration–related genes showed similar expression patterns in leaves of both desiccation-tolerant and -sensitive species. However, we identified a small set of seed-related orthologs with expression specific to desiccation-tolerant species. This supports a broad role for seed-related genes, where many are involved in typical drought responses, with only a small subset of crucial genes specifically induced in desiccation-tolerant plants.
APA, Harvard, Vancouver, ISO, and other styles
4

Tapia, Hugo, Lindsey Young, Douglas Fox, Carolyn R. Bertozzi, and Douglas Koshland. "Increasing intracellular trehalose is sufficient to confer desiccation tolerance toSaccharomyces cerevisiae." Proceedings of the National Academy of Sciences 112, no. 19 (April 27, 2015): 6122–27. http://dx.doi.org/10.1073/pnas.1506415112.

Full text
Abstract:
Diverse organisms capable of surviving desiccation, termed anhydrobiotes, include species from bacteria, yeast, plants, and invertebrates. However, most organisms are sensitive to desiccation, likely due to an assortment of different stresses such as protein misfolding and aggregation, hyperosmotic stress, membrane fracturing, and changes in cell volume and shape leading to an overcrowded cytoplasm and metabolic arrest. The exact stress(es) that cause lethality in desiccation-sensitive organisms and how the lethal stresses are mitigated in desiccation-tolerant organisms remain poorly understood. The presence of trehalose in anhydrobiotes has been strongly correlated with desiccation tolerance. In the yeastSaccharomyces cerevisiae, trehalose is essential for survival after long-term desiccation. Here, we establish that the elevation of intracellular trehalose in dividing yeast by its import from the media converts yeast from extreme desiccation sensitivity to a high level of desiccation tolerance. This trehalose-induced tolerance is independent of utilization of trehalose as an energy source, de novo synthesis of other stress effectors, or the metabolic effects of trehalose biosynthetic intermediates, indicating that a chemical property of trehalose is directly responsible for desiccation tolerance. Finally, we demonstrate that elevated intracellular maltose can also make dividing yeast tolerant to short-term desiccation, indicating that other disaccharides have stress effector activity. However, trehalose is much more effective than maltose at conferring tolerance to long-term desiccation. The effectiveness and sufficiency of trehalose as an antagonizer of desiccation-induced damage in yeast emphasizes its potential to confer desiccation tolerance to otherwise sensitive organisms.
APA, Harvard, Vancouver, ISO, and other styles
5

Oliver, Melvin J., Jill M. Farrant, Henk W. M. Hilhorst, Sagadevan Mundree, Brett Williams, and J. Derek Bewley. "Desiccation Tolerance: Avoiding Cellular Damage During Drying and Rehydration." Annual Review of Plant Biology 71, no. 1 (April 29, 2020): 435–60. http://dx.doi.org/10.1146/annurev-arplant-071219-105542.

Full text
Abstract:
Desiccation of plants is often lethal but is tolerated by the majority of seeds and by vegetative tissues of only a small number of land plants. Desiccation tolerance is an ancient trait, lost from vegetative tissues following the appearance of tracheids but reappearing in several lineages when selection pressures favored its evolution. Cells of all desiccation-tolerant plants and seeds must possess a core set of mechanisms to protect them from desiccation- and rehydration-induced damage. This review explores how desiccation generates cell damage and how tolerant cells assuage the complex array of mechanical, structural, metabolic, and chemical stresses and survive.Likewise, the stress of rehydration requires appropriate mitigating cellular responses. We also explore what comparative genomics, both structural and responsive, have added to our understanding of cellular protection mechanisms induced by desiccation, and how vegetative desiccation tolerance circumvents destructive, stress-induced cell senescence.
APA, Harvard, Vancouver, ISO, and other styles
6

Fu, J. R., J. P. Jin, Y. F. Peng, and Q. H. Xia. "Desiccation tolerance in two species with recalcitrant seeds: Clausena lansium (Lour.) and Litchi chinensis (Sonn.)." Seed Science Research 4, no. 2 (June 1994): 257–61. http://dx.doi.org/10.1017/s0960258500002245.

Full text
Abstract:
AbstractSeeds were collected at weekly intervals from mid-maturation to the fully ripened stage. As seed development progressed, desiccation tolerance increased. Desiccation tolerance of C. lansium seeds was greatest at 67 days after anthesis (DAA), when they tolerated air drying for 9 days; 74 DAA was considered as physiological maturity, and their full viability was only maintained for up to 3 days of drying; overripened seeds (88 DAA) had the lowest desiccation tolerance. In L. chinensis, the desiccation sensitivity of seeds at 98 DAA (fully mature) was higher than that at 84 and 91 DAA (less mature); among the excised embryonic axes at different developmental stages, the less mature ones were less sensitive to desiccation than the fully mature ones; excised embryonic axes of the same stage were more tolerant of desiccation than whole seeds.
APA, Harvard, Vancouver, ISO, and other styles
7

Greggains, Valerie, William E. Finch-Savage, W. Paul Quick, and Neil M. Atherton. "Putative desiccation tolerance mechanisms in orthodox and recalcitrant seeds of the genusAcer." Seed Science Research 10, no. 3 (September 2000): 317–27. http://dx.doi.org/10.1017/s0960258500000362.

Full text
Abstract:
AbstractRecalcitrant seeds are shed moist from the plant and do not survive desiccation to the low moisture contents required for prolonged storage. It has been widely hypothesised that during desiccation of these seeds a stress induced metabolic imbalance develops that leads to free radical mediated damage and viability loss. We investigated this hypothesis in a comparison of two sympatric species ofAcerduring late seed development and post-harvest desiccation:A. platanoides(Norway maple) has orthodox seeds andA. pseudoplatanus(sycamore) has recalcitrant seeds. In both species, respiration rates declined to similar levels at shedding, and the extent of defences against free radicals appears no less in sycamore than that in Norway maple. During drying there was no evidence for the accumulation of a stable free radical, increased lipid peroxidation or decline in free radical scavenging enzymes in either species. In addition, there was a very similar, large increase in total tocopherol in both species. This increase in sycamore was largely of alpha-tocopherol, whereas in Norway maple the increase was largely from its precursor, gamma-tocopherol. Arguably this suggests a similar mechanism in both species, but increased oxidative stress in sycamore. In general, the results suggest that, although damage resulting in viability loss was clearly taking place, the limitation to desiccation tolerance did not result from inadequate free radical scavenging. Soluble carbohydrates and dehydrin-like proteins were also measured during late seed development and drying in sycamore and Norway maple. The greater concentrations of sucrose, raffinose and stachyose and amounts of dehydrins in the radicles and cotyledons of Norway maple compared with those in sycamore was consistent with greater desiccation tolerance in the former. Sycamore seeds are dormant and at the tolerant end of the continuum of desiccation sensitivity among recalcitrant species, and this may account for their different response to that of the seeds of other more sensitive recalcitrant species studied.
APA, Harvard, Vancouver, ISO, and other styles
8

Sinzar-Sekulic, Jasmina, Marko Sabovljevic, and Branka Stevanovic. "Comparison of desiccation tolerance among mosses from different habitats." Archives of Biological Sciences 57, no. 3 (2005): 189–92. http://dx.doi.org/10.2298/abs0503189s.

Full text
Abstract:
Three moss species from the karst region were compared to establish their respective patterns of desiccation tolerance. Different life forms of bryophytes were chosen to obtain evidence of their life strategies during drought conditions. Comparative analyses of electrolyte leakage were performed to screen for tolerance of the membrane to water stress and for signs of damage to the fine structure of the protoplasm. The experiments were carried out by exposing the plants to water stress caused by PEG 600. The results show that the most desiccation tolerant species is Thamnobryum alopecurum, less but fairly tolerant is Anomodon viticulosus, while the aquatic Rhynchostegium riparioides is intolerant of desiccation.
APA, Harvard, Vancouver, ISO, and other styles
9

Marks, Rose A., Mpho Mbobe, Marilize Greyling, Jennie Pretorius, David Nicholas McLetchie, Robert VanBuren, and Jill M. Farrant. "Variability in Functional Traits along an Environmental Gradient in the South African Resurrection Plant Myrothamnus flabellifolia." Plants 11, no. 10 (May 18, 2022): 1332. http://dx.doi.org/10.3390/plants11101332.

Full text
Abstract:
Many desiccation-tolerant plants are widely distributed and exposed to substantial environmental variation across their native range. These environmental differences generate site-specific selective pressures that could drive natural variation in desiccation tolerance across populations. If identified, such natural variation can be used to target tolerance-enhancing characteristics and identify trait associations within a common genetic background. Here, we tested for natural variation in desiccation tolerance across wild populations of the South African resurrection plant Myrothamnus flabellifolia. We surveyed a suite of functional traits related to desiccation tolerance, leaf economics, and reproductive allocation in M. flabellifolia to test for trait associations and tradeoffs. Despite considerable environmental variation across the study area, M. flabellifolia plants were extremely desiccation tolerant at all sites, suggesting that tolerance is either maintained by selection or fixed in these populations. However, we detected notable associations between environmental variation, population characteristics, and fitness traits. Relative to mesic sites, plants in xeric sites were more abundant and larger, but were slower growing and less reproductive. The negative association between growth and reproduction with plant size and abundance pointed towards a potential growth–abundance tradeoff. The finding that M. flabellifolia is more common in xeric sites despite reductions in growth rate and reproduction suggests that these plants thrive in extreme aridity.
APA, Harvard, Vancouver, ISO, and other styles
10

WRIGHT, JONATHAN C. "Desiccation Tolerance and Water-Retentive Mechanisms in Tardigrades." Journal of Experimental Biology 142, no. 1 (March 1, 1989): 267–92. http://dx.doi.org/10.1242/jeb.142.1.267.

Full text
Abstract:
Tardigrades entering a state of anhydrobiosis (cryptobiosis) show considerable interspecific variation in desiccation tolerance, lower lethal humidities for initial desiccation ranging from 78 to 53 %. Species most tolerant of rapid initial drying also show the most rapid acquisition of tolerance to low humidities (25–31 %) following drying in high humidity. Surface area reduction during tun formation shows a significant positive regression against desiccation tolerance in the Eutardigrada. The most desiccation-tolerant species thus infold the largest areas of cuticle. By comparison, the heterotardigrade Echiniscus testudo shows a very poor capacity for surface area reduction. The thick dorsal plates may restrict cuticle intucking in this species. When hydrated tardigrades are desiccated in 80 % relative humidity they show a characteristic profile of mass loss, dehydrating rapidly in the first few minutes and then showing an abrupt reduction in transpiration (the ‘permeability slump'). This applies to living animals, which form tuns when desiccated, and to dead animals remaining extended. The permeability slump is not, therefore, a metabolic phenomenon and is not related to tun formation. Subsequent transpiration rates are very low and decline exponentially. The permeability slump allows animals to retain considerable amounts of internal water when desiccated, although less water is retained if the desiccation rate is increased. This may determine upper tolerable desiccation rates if a certain minimum quantity of water is required for the metabolic synthesis of membrane protectants. A significant positive regression between water retention and desiccation tolerance supports this hypothesis.
APA, Harvard, Vancouver, ISO, and other styles
11

O'LEARY, S. A., A. M. BURNELL, and J. R. KUSEL. "Biophysical properties of the surface of desiccation-tolerant mutants and parental strain of the entomopathogenic nematode Heterorhabditis megidis (strain UK211)." Parasitology 117, no. 4 (October 1998): 337–45. http://dx.doi.org/10.1017/s0031182098003151.

Full text
Abstract:
Entomopathogenic nematodes (EPN) are useful biological control agents of insect pests. However, the infective juvenile (IJ) stage which is the only stage to occur outside the host is susceptible to environmental extremes such as desiccation. We have isolated desiccation-tolerant strains of the EPN Heterorhabditis megidis. In this paper we describe the surface properties of these desiccation-tolerant mutants. Heterorhabditid IJs retain the sheath of the previous larval stage. The mutant lines possess alterations in the surface properties of the sheath. Differences were observed in fluorescent lipid analogue insertion into the surface of the sheath. Furthermore, cationized ferritin-binding studies demonstrated that the mutant lines possessed an increase in net negative surface charge. Removal of the surface layer of the sheath resulted in the loss of the mutant phenotype and in a reduction in the desiccation tolerance of the parental strain. Therefore, the negatively charged ‘surface coat’ appears to play an important role in the desiccation tolerance of Heterorhabditis species.
APA, Harvard, Vancouver, ISO, and other styles
12

Aldridge, C. D., and R. J. Probert. "Seed development, the accumulation of abscisic acid and desiccation tolerance in the aquatic grasses Porteresia coarctata (Roxb.) Tateoka and Oryza sativa L." Seed Science Research 3, no. 2 (June 1993): 97–103. http://dx.doi.org/10.1017/s0960258500001641.

Full text
Abstract:
AbstractSeeds of Oryza sativa L. (desiccation tolerant, orthodox) completed each of five distinct stages of development approximately 7 days earlier than seeds of Porteresia coarctata (Roxb.) Tateoka (desiccation intolerant, recalcitrant), despite the fact that O. sativa plants matured under cooler conditions. Isolated embryos of O. sativa were more sensitive to rapid desiccation at 6 days post anthesis (DPA) compared with naked caryopses. More than 90% of the latter were desiccation tolerant at 8 DPA and at all stages tested the germination capacity and/or rate of germination was greater following drying. In contrast, drying resulted in a complete loss of viability in seeds of P. coarctata at all stages tested. In both species, abscisic acid (ABA) began to accumulate in embryos mid-way through development. At the respective peaks, the concentration of ABA in P. coarctata was twice that recorded in O. sativa confirming that ABA levels per se do not determine desiccation tolerance in seeds. The possibility that desiccation tolerance is linked to sensitivity to ABA is discussed.
APA, Harvard, Vancouver, ISO, and other styles
13

Daws, Matthew I., Sheina Bolton, David F. R. P. Burslem, Nancy C. Garwood, and Christopher E. Mullins. "Loss of desiccation tolerance during germination in neo-tropical pioneer seeds: implications for seed mortality and germination characteristics." Seed Science Research 17, no. 4 (December 2007): 273–81. http://dx.doi.org/10.1017/s0960258507837755.

Full text
Abstract:
AbstractOrthodox, desiccation-tolerant seeds lose desiccation tolerance during germination. Here, we quantify the timing of the loss of desiccation tolerance, and explore the implications of this event for seed mortality and the shape of germination progress curves for pioneer tree species. For the nine species studied, all seeds in a seedlot lost desiccation tolerance after the same fixed proportion of their time to germination, and this proportion was fairly constant across the species (0.63–0.70). The loss of desiccation tolerance after a fixed proportion of the time to germination has the implication that the maximum number of seeds in a seedlot that can be killed by a single drying event during germination (Mmax) increases with an increasing time to 50% germination (t50) and an increasing slope of the germination progress curve. Consequently, to prevent the seed population from becoming highly vulnerable to desiccation-induced mortality, species with a greater t50 would be expected to have a shallower germination progress curve. In conclusion, these data suggest that the loss of desiccation tolerance during germination may constitute a significant, but previously unexplored, source of mortality for seeds in seasonal environments with unpredictable rainfall.
APA, Harvard, Vancouver, ISO, and other styles
14

Davidson, JK. "Nonparallel Geographic Patterns for Tolerance to Cold and Desiccation in Drosophila-Melanogaster and Drosophila-Simulans." Australian Journal of Zoology 38, no. 2 (1990): 155. http://dx.doi.org/10.1071/zo9900155.

Full text
Abstract:
D. melanogaster populations from the fluctuating temperate climate of Melbourne (38�S) and in the tropics at Townsville (19�S) were investigated for differentiation in cold tolerance and desiccation tolerance, and were found to differ as predicted a priori from climatic considerations. Flies from the former locality were more tolerant to both of these environmental stresses. In comparable D. simulans populations, there was no significant differentiation between populations for cold tolerance or desiccation tolerance. In both species, there was genetic variation within each population. It is hypothesised that the non-parallel patterns in these sibling species may be due to different genetic strategies in temporal variation for tolerance to the stresses associated with climatic extremes. For D. melanogaster and D. simulans, there was no strong association between cold tolerance and desiccation tolerance in either the Melbourne and Townsville populations. Correlations between the stresses over the 15 strains were calculated for the sexes, generations, localities and species separately and were all non-significant. Cold tolerance and desiccation tolerance thus involve different physiological mechanisms.
APA, Harvard, Vancouver, ISO, and other styles
15

Górecki, R. J., A. I. Piotrowicz-Cieślak, L. B. Lahuta, and R. L. Obendorf. "Soluble carbohydrates in desiccation tolerance of yellow lupin seeds during maturation and germination." Seed Science Research 7, no. 2 (June 1997): 107–16. http://dx.doi.org/10.1017/s0960258500003445.

Full text
Abstract:
AbstractMaturing yellow lupin seeds were desiccation tolerant. Glucose, sucrose and cyclitols (mainly D-pinitol, D-chiro-inositol and myo-inositol) were predominant at the early stages of seed growth. Accumulation of the raffinose family oligosaccharides (RFOs) and the galactosyl cyclitols including galactinol, digalactosyl myo-inositol, galactopinitol A, galactopinitol B, trigalactopinitol A, ciceritol, fagopyritol B1 and fagopyritol B2 appeared during seed maturation; their increase correlated with seed germinability after desiccation. The loss of desiccation tolerance after seed germination was also studied. For the desiccation tolerance test, intact seedlings were dried rapidly or slowly followed by rehydration. Soluble carbohydrates were assayed before and after drying. Root tissues were more sensitive to desiccation than hypocotyl tissues and completely lost desiccation tolerance within 36 h of imbibition after both fast and slow-drying treatments. Survival of hypocotyls decreased gradually up to 96 h after imbibition. Loss of RFOs and galactosyl cyclitols in axis tissues preceded visible germination. Loss of desiccation tolerance was accompanied by loss of RFOs and galactosyl cyclitols and an increase in reducing sugars in cotyledon, hypocotyl and radicle tissues. Drying did not induce the accumulation of RFOs and galactosyl cyclitols in seedling tissues.
APA, Harvard, Vancouver, ISO, and other styles
16

Blomstedt, Cecilia, Cara Griffiths, Donald Gaff, John Hamill, and Alan Neale. "Plant Desiccation Tolerance and its Regulation in the Foliage of Resurrection “Flowering-Plant” Species." Agronomy 8, no. 8 (August 14, 2018): 146. http://dx.doi.org/10.3390/agronomy8080146.

Full text
Abstract:
The majority of flowering-plant species can survive complete air-dryness in their seed and/or pollen. Relatively few species (‘resurrection plants’) express this desiccation tolerance in their foliage. Knowledge of the regulation of desiccation tolerance in resurrection plant foliage is reviewed. Elucidation of the regulatory mechanism in resurrection grasses may lead to identification of genes that can improve stress tolerance and yield of major crop species. Well-hydrated leaves of resurrection plants are desiccation-sensitive and the leaves become desiccation tolerant as they are drying. Such drought-induction of desiccation tolerance involves changes in gene-expression causing extensive changes in the complement of proteins and the transition to a highly-stable quiescent state lasting months to years. These changes in gene-expression are regulated by several interacting phytohormones, of which drought-induced abscisic acid (ABA) is particularly important in some species. Treatment with only ABA induces desiccation tolerance in vegetative tissue of Borya constricta Churchill. and Craterostigma plantagineum Hochstetter. but not in the resurrection grass Sporobolus stapfianus Gandoger. Suppression of drought-induced senescence is also important for survival of drying. Further research is needed on the triggering of the induction of desiccation tolerance, on the transition between phases of protein synthesis and on the role of the phytohormone, strigolactone and other potential xylem-messengers during drying and rehydration.
APA, Harvard, Vancouver, ISO, and other styles
17

Osborne, D. J., and I. I. Boubriak. "DNA and desiccation tolerance." Seed Science Research 4, no. 2 (June 1994): 175–85. http://dx.doi.org/10.1017/s0960258500002166.

Full text
Abstract:
AbstractThis article reviews mechanisms by which specialized cells of different life forms have overcome the lethal effects of dehydration and considers how the maintenance of genetic information is central to survival. As a dynamic and hydrated molecule in vivo, DNA can assume different conformational structures depending upon the water activity, the base sequence and the presence of specific binding proteins. The attainment of stable secondary structures that are resistant to degradation in vivo at low water potentials is proposed as a likely accompaniment to desiccation tolerance. In addition, chemical modification of bases in DNA, the extent of methylation and conformational changes could determine the expression of different gene sequences as cells pass from desiccation-tolerant to desiccation-intolerant states. We monitored the integrity of extracted DNA in embryos of seeds and in wind-dispersed pollen during transition from their desiccation tolerance to desiccation intolerance on hydration and germination. We present evidence to show that the DNA of these two stages is different and that it is the DNA from desiccation-tolerant cells only that retains integrity when the cells are subjected to desiccation regimes. We discuss these findings in relation to certain hydration-sensitive DNA structures and to other relevant biological systems.
APA, Harvard, Vancouver, ISO, and other styles
18

Peredo, Elena L., and Zoe G. Cardon. "Shared up-regulation and contrasting down-regulation of gene expression distinguish desiccation-tolerant from intolerant green algae." Proceedings of the National Academy of Sciences 117, no. 29 (July 7, 2020): 17438–45. http://dx.doi.org/10.1073/pnas.1906904117.

Full text
Abstract:
Among green plants, desiccation tolerance is common in seeds and spores but rare in leaves and other vegetative green tissues. Over the last two decades, genes have been identified whose expression is induced by desiccation in diverse, desiccation-tolerant (DT) taxa, including, e.g., late embryogenesis abundant proteins (LEA) and reactive oxygen species scavengers. This up-regulation is observed in DT resurrection plants, mosses, and green algae most closely related to these Embryophytes. Here we test whether this same suite of protective genes is up-regulated during desiccation in even more distantly related DT green algae, and, importantly, whether that up-regulation is unique to DT algae or also occurs in a desiccation-intolerant relative. We used three closely related aquatic and desert-derived green microalgae in the family Scenedesmaceae and capitalized on extraordinary desiccation tolerance in two of the species, contrasting with desiccation intolerance in the third. We found that during desiccation, all three species increased expression of common protective genes. The feature distinguishing gene expression in DT algae, however, was extensive down-regulation of gene expression associated with diverse metabolic processes during the desiccation time course, suggesting a switch from active growth to energy-saving metabolism. This widespread downshift did not occur in the desiccation-intolerant taxon. These results show that desiccation-induced up-regulation of expression of protective genes may be necessary but is not sufficient to confer desiccation tolerance. The data also suggest that desiccation tolerance may require induced protective mechanisms operating in concert with massive down-regulation of gene expression controlling numerous other aspects of metabolism.
APA, Harvard, Vancouver, ISO, and other styles
19

Mukuka, John, Olaf Strauch, and Ralf-Udo Ehlers. "Variability in desiccation tolerance among different strains of the entomopathogenic nematode Heterorhabditis bacteriophora." Nematology 12, no. 5 (2010): 711–20. http://dx.doi.org/10.1163/138855409x12607871174454.

Full text
Abstract:
Abstract The shelf life of biological control products based on the entomopathogenic nematode Heterorhabditis bacteriophora is rather short. In order to prolong shelf life, the metabolism of nematodes during storage must be reduced. This can be achieved by means of desiccation of the infective third-stage dauer juveniles (DJ). The tolerance can be increased by an adaptation to moderate desiccation conditions. Previous investigations indicate that the heritability of the desiccation tolerance is high, justifying a genetic selection for enhanced tolerance. This investigation screened the desiccation tolerance of 43 strains of Heterorhabditis spp. and 18 hybrid/inbred strains of H. bacteriophora. Dehydrating conditions measured as water activity (aw values) were produced by treating DJ with different concentrations of the non-ionic polymer poly(ethylene glycol) 600. Significant inter-specific variation was recorded between nematode strains and species. The mean tolerated aw value (MW50) ranged from 0.90 to 0.95 for non-adapted and 0.67 to 0.99 for adapted nematode populations. For selective breeding, only the 10% most tolerant individuals would be used. The lowest aw value tolerated by 10% of a population (MW10) ranged from of 0.845 to 0.932 for non-adapted nematode populations and 0.603 to 0.950 for adapted nematode populations. Adaptation significantly increased the desiccation tolerance and a weak correlation was recorded for tolerance with and without adaptations. The most tolerant nematode strains will form the basis for the foundation of a parental stock produced by cross-breeding and following genetic selection for enhanced tolerance. Thus, this investigation is another milestone on the road to domestication of H. bacteriophora for commercial use in sustainable pest management.
APA, Harvard, Vancouver, ISO, and other styles
20

Zhang, Zhaojie, and Gracie R. Zhang. "Chromosome-condensed G1 phase yeast cells are tolerant to desiccation stress." Microbial Cell 9, no. 2 (February 7, 2022): 42–51. http://dx.doi.org/10.15698/mic2022.02.770.

Full text
Abstract:
The budding yeast Saccharomyces cerevisiae is capable of surviving extreme water loss for a long time. However, less is known about the mechanism of its desiccation tolerance. In this study, we revealed that in an exponential culture, all desiccation tolerant yeast cells were in G1 phase and had condensed chromosomes. These cells share certain features of stationary G0 cells, such as low metabolic level. They were also replicatively young, compared to the desiccation sensitive G1 cells. A similar percentage of chromosome-condensed cells were observed in stationary phase but the condensation level was much higher than that of the log-phase cells. These chromosome-condensed stationary cells were also tolerant to desiccation. However, the majority of the desiccation tolerant cells in stationary phase do not have condensed chromosomes. We speculate that the log-phase cells with condensed chromosome might be a unique feature developed through evolution to survive unpredicted sudden changes of the environment.
APA, Harvard, Vancouver, ISO, and other styles
21

Aigner, Siegfried, Erwann Arc, Michael Schletter, Ulf Karsten, Andreas Holzinger, and Ilse Kranner. "Metabolite Profiling in Green Microalgae with Varying Degrees of Desiccation Tolerance." Microorganisms 10, no. 5 (April 30, 2022): 946. http://dx.doi.org/10.3390/microorganisms10050946.

Full text
Abstract:
Trebouxiophyceae are microalgae occupying even extreme environments such as polar regions or deserts, terrestrial or aquatic, and can occur free-living or as lichen photobionts. Yet, it is poorly understood how environmental factors shape their metabolism. Here, we report on responses to light and temperature, and metabolic adjustments to desiccation in Diplosphaera epiphytica, isolated from a lichen, and Edaphochlorella mirabilis, isolated from Tundra soil, assessed via growth and photosynthetic performance parameters. Metabolite profiling was conducted by GC–MS. A meta-analysis together with data from a terrestrial and an aquatic Chlorella vulgaris strain reflected elements of phylogenetic relationship, lifestyle, and relative desiccation tolerance of the four algal strains. For example, compatible solutes associated with desiccation tolerance were up-accumulated in D. epiphytica, but also sugars and sugar alcohols typically produced by lichen photobionts. The aquatic C. vulgaris, the most desiccation-sensitive strain, showed the greatest variation in metabolite accumulation after desiccation and rehydration, whereas the most desiccation-tolerant strain, D. epiphytica, showed the least, suggesting that it has a more efficient constitutive protection from desiccation and/or that desiccation disturbed the metabolic steady-state less than in the other three strains. The authors hope that this study will stimulate more research into desiccation tolerance mechanisms in these under-investigated microorganisms.
APA, Harvard, Vancouver, ISO, and other styles
22

Fernández-Marín, Beatriz, Miren Irati Arzac, Marina López-Pozo, José Manuel Laza, Thomas Roach, Matthias Stegner, Gilbert Neuner, and José I. García-Plazaola. "Frozen in the dark: interplay of night-time activity of xanthophyll cycle, xylem attributes, and desiccation tolerance in fern resistance to winter." Journal of Experimental Botany 72, no. 8 (February 22, 2021): 3168–84. http://dx.doi.org/10.1093/jxb/erab071.

Full text
Abstract:
Abstract While most ferns avoid freezing as they have a tropical distribution or shed their fronds, wintergreen species in temperate and boreoalpine ecosystems have to deal with sub-zero temperatures. Increasing evidence has revealed overlapping mechanisms of desiccation and freezing tolerance in angiosperms, but the physiological mechanisms behind freezing tolerance in ferns are far from clear. We evaluated photochemical and hydraulic parameters in five wintergreen fern species differing in their ability to tolerate desiccation. We assessed frond freezing tolerance, ice nucleation temperature and propagation pattern, and xylem anatomical traits. Dynamics of photochemical performance and xanthophyll cycle were evaluated during freeze–thaw events under controlled conditions and, in selected species, in the field. Only desiccation-tolerant species, which possessed a greater fraction of narrow tracheids (<18 μm) than sensitive species, tolerated freezing. Frond freezing occurred in the field at –3.4 ± 0.9 °C (SD) irrespective of freezing tolerance, freezable water content, or tracheid properties. Even in complete darkness, maximal photochemical efficiency of photosystem II was down-regulated concomitantly with zeaxanthin accumulation in response to freezing. This was reversible upon re-warming only in tolerant species. Our results suggest that adaptation for freezing tolerance is associated with desiccation tolerance through complementary xylem properties (which may prevent risk of irreversible cavitation) and effective photoprotection mechanisms. The latter includes de-epoxidation of xanthophylls in darkness, a process evidenced for the first time directly in the field.
APA, Harvard, Vancouver, ISO, and other styles
23

Beardmore, Tannis, and Pierre J. Charest. "Black spruce somatic embryo germination and desiccation tolerance. I. Effects of abscisic acid, cold, and heat treatments on the germinability of mature black spruce somatic embryos." Canadian Journal of Forest Research 25, no. 11 (November 1, 1995): 1763–72. http://dx.doi.org/10.1139/x95-191.

Full text
Abstract:
The effect of cold, abscisic acid (ABA), and heat treatments on the germination of mature black spruce somatic embryos (SEs) was examined. Specifically, the quality of mature black spruce SEs germinants (as assessed by the presence of hypocotyl vitrification), germination time (synchrony of root and shoot growth), and desiccation tolerance was evaluated following the treatments. Germination of black spruce SEs without any treatments was high (i.e., 89%), but 43% of germinants exhibited a vitrified hypocotyl and root growth lagged behind shoot growth. Mature SEs were exposed to cold (2 °C for 2 days), ABA (20 μM ABA for 7 days), and heat (40 °C for 10 min, 30 °C for 1 h, 25 °C for 12 h) treatments and then either germinated or exposed to fast or slow desiccation. The ABA, cold and heat treatments resulted in a slightly earlier, more uniform germination, due to increased root growth, but did not increase the quality of germinants. Sixty four percent of mature SEs were tolerant to slow desiccation, but only 18% were tolerant to fast desiccation to approximately a 20% water content (on a fresh mass basis). Tolerance to slow and fast desiccation significantly increased and the percentage of germinants with vitrified hypocotyls decreased significantly when the SEs were first exposed to the ABA and cold treatments compared with the mature SEs. None of the treatments resulted in an increase in the number of germinants compared with that of the mature SEs. These results show that the ABA and cold treatments are beneficial for increasing SE tolerance to slow and fast desiccation, decreasing germination time, and increasing the quality of germinants. The ABA treatment was superior with respect to desiccation tolerance and to the quality of germinants.
APA, Harvard, Vancouver, ISO, and other styles
24

Rabert, Claudia, Karla Inostroza, Silvana Bravo, Néstor Sepúlveda, and León A. Bravo. "Exploratory Study of Fatty Acid Profile in Two Filmy Ferns with Contrasting Desiccation Tolerance Reveal the Production of Very Long Chain Polyunsaturated Omega-3 Fatty Acids." Plants 9, no. 11 (October 24, 2020): 1431. http://dx.doi.org/10.3390/plants9111431.

Full text
Abstract:
Lipids are fundamental components of cell membranes and play a significant role in their integrity and fluidity. Alteration in lipid composition of membranes has been reported to be a major response to abiotic environmental stresses. This work was focused on the characterization of frond lipid composition and membrane integrity during a desiccation–rehydration cycle of two filmy fern species with contrasting desiccation tolerance: Hymenophyllum caudiculatum (less tolerant) and Hymenophyllum plicatum (more tolerant). The relative water content decreased without differences between species when both filmy ferns were subjected to desiccation. However, H. plicatum reached a higher relative water content than H. caudiculatum after rehydration. Fatty acids profiles showed the presence of a very long chain polyunsaturated fatty acid during the desiccation–rehydration cycle, with eicosatrienoic acid being the most abundant. Additionally, propidium iodide permeation staining and confocal microscopy demonstrated that, following the desiccation–rehydration cycle, H. plicatum exhibited a greater membrane integrity than H. caudiculatum. The lack of some very long chain fatty acids such as C22:1n9 and C24:1n9 in this species contrasting with H. plicatum may be associated with its lower membrane stability during the desiccation–rehydration cycle. This report provides the first insight into the fatty acid composition and dynamics of the membrane integrity of filmy ferns during a desiccation–rehydration cycle. This could potentially play a role in determining the different levels of desiccation tolerance and microhabitat preferences exhibited by Hymenophyllaceae species.
APA, Harvard, Vancouver, ISO, and other styles
25

Adegbola, Yai Ulrich, and Héctor E. Pérez. "Extensive Desiccation and Aging Stress Tolerance Characterize Gaillardia pulchella (Asteraceae) Seeds." HortScience 51, no. 2 (February 2016): 159–63. http://dx.doi.org/10.21273/hortsci.51.2.159.

Full text
Abstract:
We investigated the response of Gaillardia pulchella seeds to desiccation and aging stress to gain some perspective on the germplasm storage potential and seed vigor of this species. Seed–water relations of mature, freshly harvested G. pulchella seeds were characteristic of desiccation-tolerant species. For example, initial seed water potential (−53 MPa) was well below the lethal water potential limit (−15 MPa) for desiccation-sensitive seeds. Desiccation tolerance was confirmed by high (>70%), rapid (t50 range 4–7 days), and uniform germination following equilibration drying. Likewise, post–saturated salt accelerated aging (SSAA) germination tests indicated a high degree of vigor of fresh seeds. The substantial level of desiccation and aging-related stress tolerance in G. pulchella seeds suggests that these organs potentially display orthodox storage physiology and an ability to endure variable seed bed conditions.
APA, Harvard, Vancouver, ISO, and other styles
26

Alejo-Jacuinde, Gerardo, and Luis Herrera-Estrella. "Exploring the High Variability of Vegetative Desiccation Tolerance in Pteridophytes." Plants 11, no. 9 (April 30, 2022): 1222. http://dx.doi.org/10.3390/plants11091222.

Full text
Abstract:
In the context of plant evolution, pteridophytes, which is comprised of lycophytes and ferns, occupy an intermediate position between bryophytes and seed plants, sharing characteristics with both groups. Pteridophytes is a highly diverse group of plant species that occupy a wide range of habitats including ecosystems with extreme climatic conditions. There is a significant number of pteridophytes that can tolerate desiccation by temporarily arresting their metabolism in the dry state and reactivating it upon rehydration. Desiccation-tolerant pteridophytes exhibit a strategy that appears to be intermediate between the constitutive and inducible desiccation tolerance (DT) mechanisms observed in bryophytes and angiosperms, respectively. In this review, we first describe the incidence and anatomical diversity of desiccation-tolerant pteridophytes and discuss recent advances on the origin of DT in vascular plants. Then, we summarize the highly diverse adaptations and mechanisms exhibited by this group and describe how some of these plants could exhibit tolerance to multiple types of abiotic stress. Research on the evolution and regulation of DT in different lineages is crucial to understand how plants have adapted to extreme environments. Thus, in the current scenario of climate change, the knowledge of the whole landscape of DT strategies is of vital importance as a potential basis to improve plant abiotic stress tolerance.
APA, Harvard, Vancouver, ISO, and other styles
27

Gaff, D. F., D. Bartels, and J. L. Gaff. "Changes in Gene Expression during Drying in a Desiccation-Tolerant Grass Sporobolus stapfianus and a Desiccation-Sensitive Grass Sporobolus pyramidalis." Functional Plant Biology 24, no. 5 (1997): 617. http://dx.doi.org/10.1071/pp96073.

Full text
Abstract:
For the first time in the grasses, a desiccation-tolerant species (Sporobolus stapfianus) was examined for evidence of drought-induced changes in gene transcription. Desiccation tolerance (the ability of this species to recover from a water potential of –540 MPa) is induced in the resurrection grass during the drying process itself. Specific mRNA was compared in extracts of air-dry, drying and fully hydrated leaves by comparisons of the encoded proteins translated in vitro and partitioned by 2- dimensional electrophoresis. Forty-one genes, that were not expressed in hydrated leaves, were transcribed during drying, whereas only 25 novel polypeptides (translated in vitro) were detected; this suggests that gene expression was controlled mainly at the transcriptional level, but possibly also at the translational level. Leaves of S. stapfianus become desiccation tolerant as they dry on intact plants with mechanically undisturbed roots, whereas leaves on plants whose roots have been disturbed die during drying. Complements of mRNA from live S. stapfianus leaves changed markedly from full hydration to 70% RWC and to air-dryness; they also differed markedly from drought-sensitive leaves (on plants with disturbed roots) at 70% RWC and dead air-dry S. stapfianus leaves and from leaves of the desiccation sensitive grass S. pyramidalis at the same water contents. Drought-induced injury could not be attributed to low abundance of mRNA in either species. Five criteria which might be involved in desiccation tolerance were applied to specific in vitro proteins of S. stapfianus; 12 novel proteins correlated with desiccation tolerance in a least four of the five criteria.
APA, Harvard, Vancouver, ISO, and other styles
28

Hoekstra, F. A., A. M. Haigh, F. A. A. Tetteroo, and T. van Roekel. "Changes in soluble sugars in relation to desiccation tolerance in cauliflower seeds." Seed Science Research 4, no. 2 (June 1994): 143–47. http://dx.doi.org/10.1017/s0960258500002142.

Full text
Abstract:
AbstractChanges in soluble sugars in cauliflower seeds were followed during 50 h of imbibition in relation to desiccation tolerance. Sucrose and stachyose contents decreased, and glucose and fructose accumulated. This occurred in radicles first and subsequently in hypocotyls and cotyledons. Loss of desiccation tolerance in the various seed parts coincided with an increase in glucose and fructose and the complete loss of stachyose, but sucrose content, the major sugar, was still high. Drying imbibed seeds over silica gel did not evoke resynthesis of stachyose, but did increase sucrose and decrease glucose and fructose contents. Seeds primed in solutions of 30% polyethylene glycol for 10 days showed a loss of stachyose, while sucrose remained high and glucose and fructose contents were still very low. Redrying of primed seeds did not change the sugar contents. The primed seeds were still tolerant of desiccation. We conclude that stachyose is not a prerequisite for desiccation tolerance, but that sucrose may be. We suggest that glucose and fructose may be involved in desiccation damage.
APA, Harvard, Vancouver, ISO, and other styles
29

Marques, Alexandre, Gonda Buijs, Wilco Ligterink, and Henk Hilhorst. "Evolutionary ecophysiology of seed desiccation sensitivity." Functional Plant Biology 45, no. 11 (2018): 1083. http://dx.doi.org/10.1071/fp18022.

Full text
Abstract:
Desiccation sensitive (DS) seeds do not survive dry storage due to their lack of desiccation tolerance. Almost half of the plant species in tropical rainforests produce DS seeds and therefore the desiccation sensitivity of these seeds represents a problem for and long-term biodiversity conservation. This phenomenon raises questions as to how, where and why DS (desiccation sensitive)-seeded species appeared during evolution. These species evolved probably independently from desiccation tolerant (DT) seeded ancestors. They adapted to environments where the conditions are conducive to immediate germination after shedding, e.g. constant and abundant rainy seasons. These very predictable conditions offered a relaxed selection for desiccation tolerance that eventually got lost in DS seeds. These species are highly dependent on their environment to survive and they are seriously threatened by deforestation and climate change. Understanding of the ecology, evolution and molecular mechanisms associated with seed desiccation tolerance can shed light on the resilience of DS-seeded species and guide conservation efforts. In this review, we survey the available literature for ecological and physiological aspects of DS-seeded species and combine it with recent knowledge obtained from DT model species. This enables us to generate hypotheses concerning the evolution of DS-seeded species and their associated genetic alterations.
APA, Harvard, Vancouver, ISO, and other styles
30

Gee, O. H., R. J. Probert, and S. A. Coomber. "‘Dehydrin-like’ proteins and desiccation tolerance in seeds." Seed Science Research 4, no. 2 (June 1994): 135–41. http://dx.doi.org/10.1017/s0960258500002130.

Full text
Abstract:
AbstractThe relationship between tolerance of seeds to extreme desiccation and the presence of ‘dehydrinlike’ proteins was investigated in groups of related taxa from the unrelated plant families Aceraceae and Gramineae. Dehydrin-like proteins were identified by Western blot analysis using an antibody raised against a synthetic oligopeptide representing the 23-amino acid consensus sequence common to all group 2 late-embryogenesis-abundant (LEA) proteins.Evidence is presented that seeds of Acer pseudoplatanus and A. saccharinum are desiccation intolerant (recalcitrant) whereas seeds of A. platanoides and A. rubrum are desiccation tolerant (orthodox). Despite these differences, dehydrinlike proteins at 60 and 20 kDa were detected in all four species.Dehydrins at 20 kDa were also detected in seed samples of two aquatic grasses, Porteresia coarctata and Oryza sativa from the tribe Oryzeae, despite seeds of the former rapidly losing viability on drying, whereas O. sativa is one of the best-known examples of desiccation-tolerant seeds. In O. sativa, there was a correlation between contents of dehydrins detected and the proportion of individuals capable of withstanding extreme drying. However, the possibility of a causal link between these parameters is equivocal. Dehydrin-like proteins were also detected in desiccation-sensitive seeds of Zizania palustris, Z. latifolia and Z. texana and desiccation-intolerant seeds of Spartina anglica, all from the Gramineae.The presence of group 2 LEAs is clearly not diagnostic of desiccation tolerance in seeds. However, a more direct correlation with the expression of other groups of LEAs cannot be discounted.
APA, Harvard, Vancouver, ISO, and other styles
31

Ivanchina, Ludmila A., and Sergei V. Zalesov. "The effect of spruce plantation density on resilience of mixed forests in the Perm Krai." Journal of Forest Science 65, No. 7 (July 31, 2019): 263–71. http://dx.doi.org/10.17221/14/2019-jfs.

Full text
Abstract:
Over the course of the last few decades, many countries across the globe have experienced mass desiccation of spruce plantations. The subject of our research was the spruce forests of the Russian Perm Krai’s mixed forest zone. Spruce is a shade–tolerant tree species and low plantation density may adversely affect the spruce health. The aim of this research is to establish how influential the spruce stand density is on causing desiccation in mixed zones in the Perm Krai. The results of an on-site survey which had recorded spruce desiccation in 2017 were analysed. Within the boundaries of the aforementioned forest areas, 2017 saw the desiccation of spruce trees in 301 forest allotments covering an area of 5,343.7 ha. The value of the weighted average category of spruce forest health in Prikamye varies from 2.7 (severely weakened) to 4.2 (desiccating), and the percentage of the volume of spruce deadwood varies from 17% to 59.5%. When the spruce stand density rises from 0.4 to 0.8, spruce stand resilience to desiccation increases.
APA, Harvard, Vancouver, ISO, and other styles
32

Oren, Nadav, Stefan Timm, Marcus Frank, Oliver Mantovani, Omer Murik, and Martin Hagemann. "Red/far-red light signals regulate the activity of the carbon-concentrating mechanism in cyanobacteria." Science Advances 7, no. 34 (August 2021): eabg0435. http://dx.doi.org/10.1126/sciadv.abg0435.

Full text
Abstract:
Desiccation-tolerant cyanobacteria can survive frequent hydration/dehydration cycles likely affecting inorganic carbon (Ci) levels. It was recently shown that red/far-red light serves as signal-preparing cells toward dehydration. Here, the effects of desiccation on Ci assimilation by Leptolyngbya ohadii isolated from Israel’s Negev desert were investigated. Metabolomic investigations indicated a decline in ribulose-1,5-bisphosphate carboxylase/oxygenase carboxylation activity, and this was accelerated by far-red light. Far-red light negatively affected the Ci affinity of L. ohadii during desiccation and in liquid cultures. Similar effects were evident in the non–desiccation-tolerant cyanobacterium Synechocystis. The Synechocystis Δcph1 mutant lacking the major phytochrome exhibited reduced photosynthetic Ci affinity when exposed to far-red light, whereas the mutant ΔsbtB lacking a Ci uptake inhibitory protein lost the far-red light inhibition. Collectively, these results suggest that red/far-red light perception likely via phytochromes regulates Ci uptake by cyanobacteria and that this mechanism contributes to desiccation tolerance in strains such as L. ohadii.
APA, Harvard, Vancouver, ISO, and other styles
33

Peng, Yifang, Tianyi Ma, Xin Wang, Meijuan Zhang, Yingxu Xu, Jie Wei, Wei Sha, and Jing Li. "Proteomic and Transcriptomic Responses of the Desiccation-Tolerant Moss Racomitrium canescens in the Rapid Rehydration Processes." Genes 14, no. 2 (February 2, 2023): 390. http://dx.doi.org/10.3390/genes14020390.

Full text
Abstract:
The moss Racomitrium canescens (R. canescens) has strong desiccation tolerance. It can remain desiccated for years and yet recover within minutes of rehydration. Understanding the responses and mechanisms underlying this rapid rehydration capacity in bryophytes could identify candidate genes that improve crop drought tolerance. We explored these responses using physiology, proteomics, and transcriptomics. Label-free quantitative proteomics comparing desiccated plants and samples rehydrated for 1 min or 6 h suggesting that damage to chromatin and the cytoskeleton had occurred during desiccation, and pointing to the large-scale degradation of proteins, the production of mannose and xylose, and the degradation of trehalose immediately after rehydration. The assembly and quantification of transcriptomes from R. canescens across different stages of rehydration established that desiccation was physiologically stressful for the plants; however, the plants recovered rapidly once rehydrated. According to the transcriptomics data, vacuoles appear to play a crucial role in the early stages of R. canescens recovery. Mitochondria and cell reproduction might recover before photosynthesis; most biological functions potentially restarted after ~6 h. Furthermore, we identified novel genes and proteins related to desiccation tolerance in bryophytes. Overall, this study provides new strategies for analyzing desiccation-tolerant bryophytes and identifying candidate genes for improving plant drought tolerance.
APA, Harvard, Vancouver, ISO, and other styles
34

Leduc, Simone Nadur Motta, João Paulo Naldi Silva, Maríia Gaspar, Claudio José Barbedo, and Rita de Cássia Leone Figueiredo-Ribeiro. "Non-structural carbohydrates of immature seeds of Caesalpinia echinata (Leguminosae) are involved in the induction of desiccation tolerance." Australian Journal of Botany 60, no. 1 (2012): 42. http://dx.doi.org/10.1071/bt11236.

Full text
Abstract:
Seeds of Caesalpinia echinata fill up to physiological maturation phase ~60 days after anthesis (DAA) in the field. These seeds are desiccation tolerant to 0.08 gH2O gDW–1 and can be stored for 2 years under freezing temperatures without losing germinability. Starch (40–50%), soluble carbohydrates (10–15%, mainly sucrose and cyclitols), in addition to traces of raffinose and stachyose detected early at maturation, are supposed to be related to the acquisition of desiccation tolerance. In the present work we demonstrate that desiccation-intolerant immature seeds (45 DAA) of C. echinata can be dried until 0.14 gH2O gDW–1 when previously soaked in polyethylene glycol (PEG) solution, maintaining high germination percentage. In contrast, seeds of 55 DAA tolerated drying until 0.14 gH2O gDW–1 without previous PEG treatment, indicating that they have already reached desiccation tolerance at this developmental stage. High-performance anion exchange chromatography analysis revealed that cyclitols and sucrose increased markedly in the embryonic axes at 45 DAA after PEG treatment, reaching levels found in embryos at 55 DAA. These results suggest that PEG treatment mimics the natural maturation drying of C. echinata seeds, changing carbohydrate metabolism and triggering processes involved in desiccation tolerance.
APA, Harvard, Vancouver, ISO, and other styles
35

Le, Tuan Ngoc, Cecilia K. Blomstedt, Jianbo Kuang, Jennifer Tenlen, Donald F. Gaff, John D. Hamill, and Alan D. Neale. "Desiccation-tolerance specific gene expression in leaf tissue of the resurrection plant Sporobolus stapfianus." Functional Plant Biology 34, no. 7 (2007): 589. http://dx.doi.org/10.1071/fp06231.

Full text
Abstract:
The desiccation tolerant grass Sporobolus stapfianus Gandoger can modulate cellular processes to prevent the imposition of irreversible damage to cellular components by water deficit. The cellular processes conferring this ability are rapidly attenuated by increased water availability. This resurrection plant can quickly restore normal metabolism. Even after loss of more than 95% of its total water content, full rehydration and growth resumption can occur within 24 h. To study the molecular mechanisms of desiccation tolerance in S. stapfianus, a cDNA library constructed from dehydration-stressed leaf tissue, was differentially screened in a manner designed to identify genes with an adaptive role in desiccation tolerance. Further characterisation of four of the genes isolated revealed they are strongly up-regulated by severe dehydration stress and only in desiccation-tolerant tissue, with three of these genes not being expressed at detectable levels in hydrated or dehydrating desiccation-sensitive tissue. The nature of the putative proteins encoded by these genes are suggestive of molecular processes associated with protecting the plant against damage caused by desiccation and include a novel LEA-like protein, and a pore-like protein that may play an important role in peroxisome function during drought stress. A third gene product has similarity to a nuclear-localised protein implicated in chromatin remodelling. In addition, a UDPglucose glucosyltransferase gene has been identified that may play a role in controlling the bioactivity of plant hormones or secondary metabolites during drought stress.
APA, Harvard, Vancouver, ISO, and other styles
36

Ali, Natasha, Robin Probert, Fiona Hay, Hannah Davies, and Wolfgang Stuppy. "Post-dispersal embryo growth and acquisition of desiccation tolerance inAnemone nemorosaL. seeds." Seed Science Research 17, no. 3 (September 2007): 155–63. http://dx.doi.org/10.1017/s0960258507783149.

Full text
Abstract:
AbstractA UK seed conservation collection ofAnemone nemorosaL. seeds held at the Millennium Seed Bank (MSB) showed low viability in its first post-storage test. Because achenes ofA. nemorosaare naturally dispersed when they are green, we tested the hypothesis that seeds may not be fully desiccation tolerant and storable at the time of natural dispersal, and that a post-harvest treatment could increase the proportion of desiccation-tolerant seeds. Achenes harvested at the point of natural dispersal in late May in 2003 and 2004 were either placed immediately on 1% water agar at 20°C (‘laboratory’ treatment), or placed in nylon sachets and buried in leaf litter among plants growing in the wild (‘field’ treatment). Samples were withdrawn at intervals over a period of 168 d and tested for desiccation tolerance (drying to 0.059 g H2O (g DW)− 1) and longevity (controlled ageing at 60% relative humidity and 45°C). An initial increase, followed by a decline, in the proportion of seeds surviving desiccation and in the longevity of both laboratory- and field-treated samples coincided with the development of embryos from globular to heart- and then torpedo-shaped. Developmental arrest was not required for the acquisition of desiccation tolerance, and continued growth and development of the embryo resulted in a loss of desiccation tolerance, similar to that seen in orthodox seeds upon radicle emergence. Furthermore, whileA. nemorosaseeds, like many from theRanunculaceaefamily, might be described as having morphological or morphophysiological dormancy, this lack of developmental arrest does not fit with the usual concept of dormancy. The implications of these results for the classification systems of seed-storage behaviour and dormancy, and for the long-term conservation of seeds ofA. nemorosa, are discussed.
APA, Harvard, Vancouver, ISO, and other styles
37

Khanam, Salma, Kimie Atsuzawa, and Yasuko Kaneko. "Localization of Lipid Droplets in Embryonic Axis Radicle Cells of Soybean Seeds under Various Imbibition Regimes Indicates Their Role in Desiccation Tolerance." Plants 12, no. 4 (February 10, 2023): 799. http://dx.doi.org/10.3390/plants12040799.

Full text
Abstract:
Desiccation tolerance allows plant seeds to remain viable during desiccation and subsequent re-hydration. In this study, we tried to develop an experimental system to understand the difference between desiccation tolerant and desiccation sensitive radicle cells by examining excised embryonic axes after re-desiccation and subsequent imbibition under various regimes. Embryonic axes excised from soybean (Glycine max (L.) Merr.) seeds imbibed for 3 h to 15 h which remained attached to the cotyledons during imbibition would grow normally after 24 h of desiccation and re-imbibition on wet filter paper. By contrast, when the embryonic axes excised after 3 h imbibition of seeds were kept on wet filter paper for 12 h to 16 h, their growth was significantly retarded after 24 h of desiccation and subsequent re-imbibition. Numerous lipid droplets were observed lining the plasma membrane and tonoplasts in radicle cells of desiccation tolerant embryonic axes before and after desiccation treatment. By contrast, the lipid droplets lining the plasma membrane and tonoplasts became very sparse in radicle cells that were placed for longer times on wet filter paper before desiccation. We observed a clear correlation between the amount of lipid droplets lining plasma membranes and the ability to grow after desiccation and re-imbibition of the excised embryonic axes. In addition to the reduction of lipid droplets in the cells, a gradual increase in starch grains was observed. Large starch grains accumulated in the radicle cells of those axes that failed to grow further.
APA, Harvard, Vancouver, ISO, and other styles
38

Mladenov, Petko, Diana Zasheva, Sébastien Planchon, Céline C. Leclercq, Denis Falconet, Lucas Moyet, Sabine Brugière, et al. "Proteomics Evidence of a Systemic Response to Desiccation in the Resurrection Plant Haberlea rhodopensis." International Journal of Molecular Sciences 23, no. 15 (July 31, 2022): 8520. http://dx.doi.org/10.3390/ijms23158520.

Full text
Abstract:
Global warming and drought stress are expected to have a negative impact on agricultural productivity. Desiccation-tolerant species, which are able to tolerate the almost complete desiccation of their vegetative tissues, are appropriate models to study extreme drought tolerance and identify novel approaches to improve the resistance of crops to drought stress. In the present study, to better understand what makes resurrection plants extremely tolerant to drought, we performed transmission electron microscopy and integrative large-scale proteomics, including organellar and phosphorylation proteomics, and combined these investigations with previously published transcriptomic and metabolomics data from the resurrection plant Haberlea rhodopensis. The results revealed new evidence about organelle and cell preservation, posttranscriptional and posttranslational regulation, photosynthesis, primary metabolism, autophagy, and cell death in response to desiccation in H. rhodopensis. Different protective intrinsically disordered proteins, such as late embryogenesis abundant (LEA) proteins, thaumatin-like proteins (TLPs), and heat shock proteins (HSPs), were detected. We also found a constitutively abundant dehydrin in H. rhodopensis whose phosphorylation levels increased under stress in the chloroplast fraction. This integrative multi-omics analysis revealed a systemic response to desiccation in H. rhodopensis and certain targets for further genomic and evolutionary studies on DT mechanisms and genetic engineering towards the improvement of drought tolerance in crops.
APA, Harvard, Vancouver, ISO, and other styles
39

Shamrock, Vanessa J., and George G. Lindsey. "A compensatory increase in trehalose synthesis in response to desiccation stress in Saccharomyces cerevisiae cells lacking the heat shock protein Hsp12p." Canadian Journal of Microbiology 54, no. 7 (July 2008): 559–68. http://dx.doi.org/10.1139/w08-044.

Full text
Abstract:
The effect of HSP12 deletion on the response of yeast to desiccation was investigated. The Δhsp12 strain was found to be more desiccation tolerant than the wild-type strain. Furthermore, the increased intracellular trehalose levels in the Δhsp12 strain suggested that this strain compensated for the lack of Hsp12p synthesis by increasing trehalose synthesis, which facilitated increased desiccation tolerance. Results obtained from flow cytometry using the membrane exclusion dye propidium iodide suggested that Hsp12p helped maintain plasma membrane integrity during desiccation. Analysis of the oxidative loads experienced by the wild-type and Δhsp12 strains showed that during mid-exponential phase, the increased trehalose levels present in the Δhsp12 cells resulted in increased protection of these cells against reactive oxygen species compared with wild-type cells. During stationary phase, lower levels of reactive oxygen species reduction by reduced glutathione was enhanced in the wild-type strain, which displayed lower intracellular trehalose concentrations. Comparison of the tolerance of the wild-type and Δhsp12 strains with applied oxidative stress showed that the Δhsp12 strain was more tolerant to exogenously applied H2O2, which we attributed to the higher intracellular trehalose concentration. Flow cytometry demonstrated that Hsp12p played a role in maintaining plasma membrane integrity during applied oxidative stress.
APA, Harvard, Vancouver, ISO, and other styles
40

Nimkingrat, Prakaijan, Felix Uhlmann, Olaf Strauch, and Ralf-Udo Ehlers. "Desiccation tolerance of dauers of entomopathogenic nematodes of the genus Steinernema." Nematology 15, no. 4 (2013): 451–58. http://dx.doi.org/10.1163/15685411-00002692.

Full text
Abstract:
For transport of entomopathogenic nematodes to the user, developmentally arrested dauer juveniles (DJ) are mixed with inert carriers at high density. If quiescence is not induced, DJ will quickly lose energy reserves and die. To induce quiescence DJ can be moderately desiccated. This study investigated the desiccation tolerance by measurement of water activity (-value) tolerated by 50% of populations (WA50) of different Steinernema species and strains. DJ were tested with or without prior adaptation to desiccation stress. Stress conditions were produced by exposure to various concentrations of poly(ethylene glycol) 600. Significant differences in desiccation tolerance were recorded between strains and species, but, except for Steinernema abbasi, not within strains of one species. Without adaptation to stress conditions, the most tolerant species was S. carpocapsae (WA50 = 0.836) followed by S. abbasi (0.86). Adaption to stress by exposure to an -value of 0.95 for 48 h increased tolerance to 0.68 and 0.66, respectively. The least tolerant species were S. kraussei, S. glaseri and S. ethiopiense. Tolerance recorded would allow storage at a water activity that would inhibit growth of bacteria but not of fungi. Analysis of water activity tolerated by only 10% of the population indicated potential for genetic improvement by selective breeding for S. carpocapsae, S. abbasi and S. arenarium to reach water activity levels that would also inhibit fungal growth.
APA, Harvard, Vancouver, ISO, and other styles
41

Gaff, Donald F., Cecilia K. Blomstedt, Alan D. Neale, Tuan N. Le, John D. Hamill, and Hamid R. Ghasempour. "Sporobolus stapfianus, a model desiccation-tolerant grass." Functional Plant Biology 36, no. 7 (2009): 589. http://dx.doi.org/10.1071/fp08166.

Full text
Abstract:
Sporobolus stapfianus Gandoger, one of ~40 known ‘anabiotic’grass species (i.e. ‘able to regain vital activity from a state of latent life’), is the most versatile tool for research into desiccation tolerance in vegetative grass tissue. Current knowledge on this species is presented, including the features that suit it for investigations into the plant’s ability to survive dehydration of its leaf protoplasm. The main contributors to desiccation tolerance in S. stapfianus leaves appear to be: accumulation during dehydration of protectants of membranes and proteins; mechanisms limiting oxidative damage; a retention of protein synthetic activity in late stages of drying that is linked with changes in gene expression and in the proteomic array; and an ability to retain net synthesis of ATP during drying. S. stapfianus exemplifies an advanced stage of an evolutionary trend in desiccation tolerant plants towards increased importance of the dehydration phase (for induction of tolerance, for synthesis of protectants and for proteomic changes).
APA, Harvard, Vancouver, ISO, and other styles
42

Tamaru, Yoshiyuki, Yayoi Takani, Takayuki Yoshida, and Toshio Sakamoto. "Crucial Role of Extracellular Polysaccharides in Desiccation and Freezing Tolerance in the Terrestrial Cyanobacterium Nostoc commune." Applied and Environmental Microbiology 71, no. 11 (November 2005): 7327–33. http://dx.doi.org/10.1128/aem.71.11.7327-7333.2005.

Full text
Abstract:
ABSTRACT The cyanobacterium Nostoc commune is adapted to the terrestrial environment and has a cosmopolitan distribution. In this study, the role of extracellular polysaccharides (EPS) in the desiccation tolerance of photosynthesis in N. commune was examined. Although photosynthetic O2 evolution was not detected in desiccated colonies, the ability of the cells to evolve O2 rapidly recovered after rehydration. The air-dried colonies contained approximately 10% (wt/wt) water, and field-isolated, natural colonies with EPS were highly water absorbent and were rapidly hydrated by atmospheric moisture. The cells embedded in EPS in Nostoc colonies were highly desiccation tolerant, and O2 evolution was not damaged by air drying. Although N. commune was determined to be a mesophilic cyanobacterium, the cells with EPS were heat tolerant in a desiccated state. EPS could be removed from cells by homogenizing colonies with a blender and filtering with coarse filter paper. This treatment to remove EPS did not damage Nostoc cells or their ability to evolve O2, but O2 evolution was significantly damaged by desiccation treatment of the EPS-depleted cells. Similar to the EPS-depleted cells, the laboratory culture strain KU002 had only small amount of EPS and was highly sensitive to desiccation. In the EPS-depleted cells, O2 evolution was also sensitive to freeze-thaw treatment. These results strongly suggest that EPS of N. commune is crucial for the stress tolerance of photosynthesis during desiccation and during freezing and thawing.
APA, Harvard, Vancouver, ISO, and other styles
43

Andersson, L., A. Yahya, S. Johansson, and J. Liew. "Seed desiccation tolerance and storage behaviour in Cordeauxia edulis." Seed Science and Technology 35, no. 3 (October 1, 2007): 660–73. http://dx.doi.org/10.15258/sst.2007.35.3.13.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Martyn, A. J., L. U. Seed, and C. A. Offord. "Seed desiccation tolerance of threatened Australian species Myrsine richmondensis." Seed Science and Technology 36, no. 1 (April 1, 2008): 206–9. http://dx.doi.org/10.15258/sst.2008.36.1.23.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Veiga-Barbosa, L., S. Mira, M. E. González-Benito, M. M. Souza, L. M. M. Meletti, and F. Pérez-García. "Seed germination, desiccation tolerance and cryopreservation of Passiflora species." Seed Science and Technology 41, no. 1 (April 1, 2013): 89–97. http://dx.doi.org/10.15258/sst.2013.41.1.08.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Chen, Tony H. H., Paul Murakami, Porter Lombard, and Leslie H. Fuchigami. "Desiccation Tolerance in Bare-rooted Apple Trees Prior to Transplanting." Journal of Environmental Horticulture 9, no. 1 (March 1, 1991): 13–17. http://dx.doi.org/10.24266/0738-2898-9.1.13.

Full text
Abstract:
Abstract Ten types of apple (Malus domestica L.) trees (six different scions on M.7 rootstock and four ‘Red Delicious’ scions on M.7, MM. 111, MM.106 and seedling rootstocks) were subjected to air drying for periods of 0 to 48 hr with or without 3 months of cold storage at 0°C (32°F). The kinetics of water loss during drying treatment and the transplanting survival and regrowth vigor were recorded. Both the scions and rootstocks influenced the tolerance of apple trees to desiccation stress. Among the plant materials tested, ‘Red Delicious’ on MM. 111 rootstock had the highest level of tolerance to desiccation. Three months of cold storage at 0°C (32°F) resulted in the considerable loss of tissue water, but the grafted trees survived if no further desiccation occurred prior to planting. Only ‘Red Delicious’/MM. 111 tolerated desiccation from the combination of three months of cold storage followed by a 48 hr exposure to air drying, while other scion/rootstock compound systems seldom survived. The analyses of water loss from apple trees indicated that the loss did not follow a simple first order reaction. However, there were two distinct first order water loss slopes suggesting that the loss was from two fractions of water inside plant tissues. One fraction of tissue lost water at a faster rate than the following second fraction which was slowly released from the plant tissues. There was no difference in the critical water content and rate of water loss between the tolerant trees, (i.e., ‘Red Delicious’ on MM. 111) and the others. Therefore it is suggested that trees on MM. 111 are more tolerant to desiccation because of the tolerance to water loss in the tissue.
APA, Harvard, Vancouver, ISO, and other styles
47

Aberlenc-Bertossi, Frédérique, Nathalie Chabrillange, Françoise Corbineau, and Yves Duval. "Acquisition of desiccation tolerance in developing oil palm (Elaeis guineensis Jacq.) embryos in planta and in vitro in relation to sugar content." Seed Science Research 13, no. 2 (June 2003): 179–86. http://dx.doi.org/10.1079/ssr2003135.

Full text
Abstract:
AbstractRelationships between desiccation tolerance and dry matter, water and sugar contents were studied throughout the development of oil palm (Elaeis guineensis Jacq.) zygotic embryos and in immature embryos cultured on a sucrose-enriched medium. Embryo dry weight during in planta development increased between 80 and 140 d after pollination (DAP) and was then stable until maturity. Embryos underwent dehydration until 120 DAP, but their moisture content remained high at maturity (c. 2 g H2O g-1 DW). Desiccation tolerance was acquired between 83 and 104 DAP, and was positively correlated with embryo age and dry weight, and negatively correlated with initial water content during this period. Sucrose, the main soluble sugar present throughout embryo development, accounted for an average of 24% of the dry weight. Glucose and fructose contents decreased to less than 1 mg g-1 DW in embryos at maturity. At 117 DAP, as embryos became tolerant to desiccation, the monosaccharides/sucrose ratio fell to 0.015 and raffinose was detected. Stachyose appeared later in 147-day-old embryos and accumulated until shedding. In vitro culture of immature embryos in the presence of high sucrose concentrations (350 and 700 mM) resulted in an increase in their dry weight and a decrease in their water content, and induced the acquisition of desiccation tolerance. Under these conditions, sucrose accumulated in embryos to 30–40% on a dry weight basis, but neither raffinose nor stachyose was detected. Acquisition of desiccation tolerance by oil palm immature embryos was associated both in planta and in vitro with an accumulation of dry matter, a reduction of moisture content, and a fall in the monosaccharides/sucrose ratio. In planta, survival to dehydration was also related with the deposition of oligosaccharides whereas in vitro, it was related with high sucrose accumulation. The role of sugars in the acquisition of desiccation tolerance in oil palm embryos is discussed.
APA, Harvard, Vancouver, ISO, and other styles
48

Kijowska-Oberc, Joanna, Aleksandra M. Staszak, Mikołaj K. Wawrzyniak, and Ewelina Ratajczak. "Changes in Proline Levels during Seed Development of Orthodox and Recalcitrant Seeds of Genus Acer in a Climate Change Scenario." Forests 11, no. 12 (December 18, 2020): 1362. http://dx.doi.org/10.3390/f11121362.

Full text
Abstract:
In the present study, we examined the utility of proline usage as a biochemical indicator of metabolic changes caused by climate change (mean temperature and precipitation) during seed development of two Acer species differing in desiccation tolerance: Norway maple (Acer platanoides L.—desiccation tolerant—orthodox) and sycamore (Acer pseudoplatanus L.—desiccation sensitive—recalcitrant). In plants, proline is an element of the antioxidant system, which has a role in response to water loss and high temperatures. Our study considered whether proline could be treated as an indicator of tree seed viability, crucial for genetic resources conservation. Proline content was measured biweekly in developing seeds (between 11 and 23 weeks after flowering) collected in consecutive years (2017, 2018, and 2019). We showed that proline concentrations in recalcitrant seeds were positively correlated with mean two-week temperature. In contrast, in orthodox seeds no such relationship was found. Proline content proved to be sensitive to thermal-moisture conditions changes, which makes it a promising biochemical marker of seed desiccation tolerance in different climatic conditions.
APA, Harvard, Vancouver, ISO, and other styles
49

Legardón, Ane, and José Ignacio García-Plazaola. "Gesneriads, a Source of Resurrection and Double-Tolerant Species: Proposal of New Desiccation- and Freezing-Tolerant Plants and Their Physiological Adaptations." Biology 12, no. 1 (January 10, 2023): 107. http://dx.doi.org/10.3390/biology12010107.

Full text
Abstract:
Gesneriaceae is a pantropical family of plants that, thanks to their lithophytic and epiphytic growth forms, have developed different strategies for overcoming water scarcity. Desiccation tolerance or “resurrection” ability is one of them: a rare phenomenon among angiosperms that involves surviving with very little relative water content in their tissues until water is again available. Physiological responses of desiccation tolerance are also activated during freezing temperatures, a stress that many of the resurrection gesneriads suffer due to their mountainous habitat. Therefore, research on desiccation- and freezing-tolerant gesneriads is a great opportunity for crop improvement, and some of them have become reference resurrection angiosperms (Dorcoceras hygrometrica, Haberlea rhodopensis and Ramonda myconi). However, their difficult indoor cultivation and outdoor accessibility are major obstacles for their study. Therefore, this review aims to identify phylogenetic, geoclimatic, habitat, and morphological features in order to propose new tentative resurrection gesneriads as a way of making them more reachable to the scientific community. Additionally, shared and species-specific physiological responses to desiccation and freezing stress have been gathered as a stress response metabolic basis of the family.
APA, Harvard, Vancouver, ISO, and other styles
50

Finch-Savage, W. E. "Seed development in the recalcitrant species Quercus robur L.: germinability and desiccation tolerance." Seed Science Research 2, no. 1 (March 1992): 17–22. http://dx.doi.org/10.1017/s0960258500001069.

Full text
Abstract:
AbstractChanges in germination and desiccation sensitivity were measured throughout the seed expansion phase of development in fruits of Quercus robur L. The onset of a reduction in sensitivity to desiccation during development on the tree coincided with acquisition of the capacity for seed germination on moist sand substrate. Tolerance of desiccation then increased throughout development to shedding, but viability was still lost at a relatively high moisture content, and seeds did not therefore pass through a fully desiccation-tolerant phase. These data suggest that desiccation sensitivity in Q. robur may have resulted from the premature termination of development.No desiccation was required to initiate germination in prematurely harvested fruits orseeds. Germination rate of seeds on moist sand increased at successive harvests during development, and was also increased by presoaking seeds in water. Variation in germination rate following shedding was not related to seed size or moisture content, but was affected by shedding date. This effect was not observed when the pericarp was removed.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography