To see the other types of publications on this topic, follow the link: Aluminum plants.

Journal articles on the topic 'Aluminum plants'

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 'Aluminum plants.'

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

Panda, Sanjib Kumar, František Baluška, and Hideaki Matsumoto. "Aluminum stress signaling in plants." Plant Signaling & Behavior 4, no. 7 (July 2009): 592–97. http://dx.doi.org/10.4161/psb.4.7.8903.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Fageria, N. K., V. C. Ballgar, and R. J. Wright. "Aluminum toxicity in crop plants." Journal of Plant Nutrition 11, no. 3 (March 1988): 303–19. http://dx.doi.org/10.1080/01904168809363804.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Zhang, Liangliang, Ruiqiang Liu, Benjamin W. Gung, Steven Tindall, Javier M. Gonzalez, Jonathan J. Halvorson, and Ann E. Hagerman. "Polyphenol–Aluminum Complex Formation: Implications for Aluminum Tolerance in Plants." Journal of Agricultural and Food Chemistry 64, no. 15 (April 5, 2016): 3025–33. http://dx.doi.org/10.1021/acs.jafc.6b00331.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Barinaga, M. "PLANT SCIENCE: Making Plants Aluminum Tolerant." Science 276, no. 5318 (June 6, 1997): 1497. http://dx.doi.org/10.1126/science.276.5318.1497.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Delhaize, E., and P. R. Ryan. "Aluminum Toxicity and Tolerance in Plants." Plant Physiology 107, no. 2 (February 1, 1995): 315–21. http://dx.doi.org/10.1104/pp.107.2.315.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Zheng, Lu, Ping Lan, Ren Fang Shen, and Wen Feng Li. "Proteomics of aluminum tolerance in plants." PROTEOMICS 14, no. 4-5 (March 2014): 566–78. http://dx.doi.org/10.1002/pmic.201300252.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Freitas, Lucas B. de, Dirceu M. Fernandes, Suelen C. M. Maia, Arianne Moniz, Beatriz G. Mazziero, and Fábio Steiner. "Sources and doses of aluminum in experiments with rice in nutrient solution." Revista Brasileira de Engenharia Agrícola e Ambiental 23, no. 7 (July 2019): 511–17. http://dx.doi.org/10.1590/1807-1929/agriambi.v23n7p511-517.

Full text
Abstract:
ABSTRACT The aluminum source to produce toxicity in upland rice in nutrient solution experiments is not yet well established, althought the aluminum potassium sulfate has been utilized source to produce aluminum toxicity. However, in recent studies have used aluminum chloride. The aim of this study was to evaluate the capacity of aluminum sources and doses to produce toxicity in upland rice plants grown in nutrient solution. The experiment was arranged in a block randomized design, in a 2 x 5 factorial scheme and four repetitions. The treatments were two aluminum sources (aluminum potassium sulfate - AlK(SO4)2.12H2O and aluminum chloride - AlCl3.6H2O) and five aluminum doses in nutrient solution (0, 370, 740, 1100 and 1480 μmol L-1). The experiment was conducted in a greenhouse in Botucatu city, São Paulo state, Brazil, starting in April 2012, and was carried out for 56 days from transplanting of the seedlings. Using aluminum chloride, the rice plants show lower production of root and total dry weight, area and root volume, medium and thick root length, potassium and sulfur contents and accumulations. Using aluminum potassium sulfate, there are lower aluminum activity and availability, besides the formation of large amount of aluminum compounds non-toxic to the plants (aluminum sulfate) in the nutrient solution. The aluminum doses between 1100 to 1480 µmol L-1, corresponding to aluminum activity of 336.8 to 429.0 µmol L-1 of aluminum chloride as source, are more effective to produce aluminum toxicity in upland rice plants grown in nutrient solution.
APA, Harvard, Vancouver, ISO, and other styles
8

He, H., Y. Li, and L. F. He. "Aluminum toxicity and tolerance in Solanaceae plants." South African Journal of Botany 123 (July 2019): 23–29. http://dx.doi.org/10.1016/j.sajb.2019.02.008.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Roy, Ajoy Kumar, Archana Sharma, and Geeta Talukder. "Some aspects of aluminum toxicity in plants." Botanical Review 54, no. 2 (April 1988): 145–78. http://dx.doi.org/10.1007/bf02858527.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Blom, Theo J., and Brian D. Piott. "Florists' Hydrangea Blueing with Aluminum Sulfate Applications during Forcing." HortScience 27, no. 10 (October 1992): 1084. http://dx.doi.org/10.21273/hortsci.27.10.1084.

Full text
Abstract:
The Al content was determined in roots, buds, and stems of dormant florists' hydrangeas [Hydrangea macrophylla subsp. macrophylla var. macrophylla (Thunb.) `Mathilda Gutges' and `Brestenburg'] that were or were not treated in the field with aluminum sulfate. During the greenhouse forcing stage, previously nontreated plants were subjected to four successive weekly subirrigated applications of aluminum sulfate totalling 4, 8, 12, or 16 g/pot. Applications were early (weeks 2, 3, 4, 5) or late (weeks 6, 7, 8, 9), using the start of forcing as week = 0. The Al contents in stems and buds of dormant plants were about five to six times higher in field-treated than in nontreated plants. Roots were the primary location of Al accumulation (≈70%). Aluminum sulfate applications of 12 to 16 g/pot during greenhouse forcing provided commercially acceptable blue plants. Maximum foliar Al concentration was 50% higher in early than in late-treated plants and calculated to occur with 14.5 and 12.2 g aluminum sulfate/pot for early and late-treated plants, respectively. There was a positive correlation (r = 0.74) between blueness ranking and the Al foliar concentration of the two uppermost expanded leaves taken from flowering plants.
APA, Harvard, Vancouver, ISO, and other styles
11

Zhang, Hua, Yakang Song, Zhenglei Fan, Jianyun Ruan, Jianhui Hu, and Qunfeng Zhang. "Aluminum Supplementation Mediates the Changes in Tea Plant Growth and Metabolism in Response to Calcium Stress." International Journal of Molecular Sciences 25, no. 1 (December 30, 2023): 530. http://dx.doi.org/10.3390/ijms25010530.

Full text
Abstract:
Tea plants are more sensitive to variations in calcium concentration compared to other plants, whereas a moderate aluminum concentration facilitates the growth and development of tea plants. Aluminum and calcium show a competitive interaction with respect to the uptake of elements, consequently exerting physiological effects on plants. To further explore these interactions, in this study, we used the solution culture method to treat tea plants with two calcium concentrations (0.8 mM and 5.6 mM) and three aluminum concentrations (0 mM, 0.4 mM, and 1 mM). We then determined the influence of the combined treatments on root growth and quality compound accumulation in the tissues by a combination of phenotype, gene expression, and metabolite analyses. Moderate aluminum supplementation (0.4 mM) alleviated the inhibition of root growth caused by high calcium stress. High calcium stress significantly inhibited the accumulation of most amino acids (e.g., Glutamic acid, Citulline, and Arginine) and organic acids (e.g., a-ketoglutaric acid) in the roots, stems, and leaves, whereas aluminum deficiency significantly increased most amino acids in the roots and leaves (except Serine, Alanine, and Phenylalanine in the roots and Ser in the leaves), with a more than two-fold increase in Arg and Lysine. High calcium stress also induced the accumulation of secondary metabolites such as epigallocatechin gallate and procyanidin in the roots, whereas aluminum supplementation significantly reduced the contents of flavonol glycosides such as quercetin, rutin, myricitrin, and kaempferitrin, as well as caffeine, regardless of calcium concentration. Aluminum supplementation reversed some of the changes in the contents of leaf metabolites induced by calcium stress (e.g., 4-dihydroquercetin, apigenin C-pentoside, phenethylamine, and caffeine). Overall, calcium stress caused severe growth inhibition and metabolic disorders in tea plants, which could be reversed by aluminum supplementation, particularly in maintaining the root tips and the accumulation of secondary metabolites. These results provide a theoretical basis for improving calcium-aluminum nutrient management to promote tea plant growth and quality.
APA, Harvard, Vancouver, ISO, and other styles
12

Chauhan, Devendra Kumar, Vaishali Yadav, Marek Vaculík, Walter Gassmann, Sharon Pike, Namira Arif, Vijay Pratap Singh, Rupesh Deshmukh, Shivendra Sahi, and Durgesh Kumar Tripathi. "Aluminum toxicity and aluminum stress-induced physiological tolerance responses in higher plants." Critical Reviews in Biotechnology 41, no. 5 (April 18, 2021): 715–30. http://dx.doi.org/10.1080/07388551.2021.1874282.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Hosseini, Farnaz, Sara Davari, and Mojtaba Arjomandi. "Review of constructive analytical methods for determining the amount of aluminum in environmental and human biological samples." Analytical Methods in Environmental Chemistry Journal 2, no. 01 (March 12, 2019): 15–32. http://dx.doi.org/10.24200/amecj.v2.i01.51.

Full text
Abstract:
Aluminum is a toxic metal and causes pollution in soil, water, and air. Afterward, a lot of patients suffer renal failure due to the accumulation of aluminum in the tissues of kidneys. Also, a high concentration of aluminum in plants tissues makes agricultural food toxic. Therefore, measuring aluminum in water, soil, air, human organs, tissues of plants and each food (or agricultural product is so necessary for protecting human health. Also, the effect of some parameters such as pH and temperature on decrease or increase in the amount of aluminum in water and other samples are stated. In this review, the analytical methods such as fluorimetric, ICP-MS, colorimetric, graphite furnace/flame atomic absorption spectrometry, etc. which have been applied for measuring the amount of aluminum (especially ) in environmental and human biological samples are assesse
APA, Harvard, Vancouver, ISO, and other styles
14

Benson, D. M. "Aluminum Amendment of Potting Mixes for Control of Phytophthora Damping-off in Bedding Plants." HortScience 30, no. 7 (December 1995): 1413–16. http://dx.doi.org/10.21273/hortsci.30.7.1413.

Full text
Abstract:
Control of preemergence damping-off caused by Phytophthora parasitica Dastur was investigated on three bedding plant species in a 1 peat: 1 vermiculite medium (v/v) limed at 3 kg·m–3 and drenched with aluminum at 10, 25, or 50 meq Al/100 cm3 medium. Aluminum as Al2(SO4)3 was applied as a drench at 0.75, 1.9, or 3.75 g/150 ml water to the surface of infested medium in 650-cm2 plug trays (1300-cm3 tray volume). All concentrations of aluminum were effective in controlling preemergence damping-off of snapdragon (Antirrhinum majus L.) and vinca (Catharanthus roseus G. Don, Madagascar periwinkle), but only 50 meq Al+3/100 cm3 medium was effective for petunia (Petunia ×hybrida Hort. Vilm.-Andr.). At 4 days after seeding and drenching with aluminum sulfate, exchangeable aluminum was 0, 0.5, and 2.03 meq Al+3/100 g medium, respectively, for the three concentrations used. Control of damping-off of snapdragon and vinca with 10 meq Al+3/100 cm3 medium with no detectable exchangeable aluminum 4 days after application suggests that P. parasitica was suppressed by aluminum early in the host–pathogen interaction, whereas petunia was susceptible to damping-off for a longer period before seedling emergence. Aluminum was not phytotoxic to vinca, snapdragon, or petunia grown in a limed medium.
APA, Harvard, Vancouver, ISO, and other styles
15

Storozhev, Yu I., and V. S. Zlobin. "Perspective Solutions of Environmental Problems of Aluminum Plants." Ecology and Industry of Russia 22, no. 12 (December 4, 2018): 10–13. http://dx.doi.org/10.18412/1816-0395-2018-12-10-13.

Full text
Abstract:
The mechanism of formation of harmful components of anode gases of aluminum production is considered. For purification of anode gases the nepheline slime, which is alumina withdrawal from nepheline ore, is offered as adsorbent and the catalyst. The new design of a gas-collecting bell of the Soderbergh,s electrolyzer for collecting and thermal neutralization of anode gases with adjustable air supply is developed.
APA, Harvard, Vancouver, ISO, and other styles
16

Freitas, L. B., D. M. Fernandes, S. C. M. Maia, and B. G. Mazziero. "Aluminum in mineral nutrition of upland rice plants." Revista Brasileira de Ciências Agrárias - Brazilian Journal of Agricultural Sciences 12, no. 1 (March 30, 2017): 26–34. http://dx.doi.org/10.5039/agraria.v12i1a5414.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Brichkova, G. G., A. M. Shishlova, T. V. Maneshina, and N. A. Kartel’. "Tolerance to aluminum in genetically modified tobacco plants." Cytology and Genetics 41, no. 3 (June 2007): 151–55. http://dx.doi.org/10.3103/s0095452707030036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Vlaskin, M. S., A. Z. Zhuk, V. I. Miroshnichenko, and A. E. Sheindlin. "Prospective Schemes of Aluminum–Hydrogen Thermal Power Plants." High Temperature 56, no. 5 (September 2018): 774–82. http://dx.doi.org/10.1134/s0018151x18050267.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

He, Huyi, Jie Zhan, Longfei He, and Minghua Gu. "Nitric oxide signaling in aluminum stress in plants." Protoplasma 249, no. 3 (August 18, 2011): 483–92. http://dx.doi.org/10.1007/s00709-011-0310-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

He, Huyi, Longfei He, and Minghua Gu. "Role of microRNAs in aluminum stress in plants." Plant Cell Reports 33, no. 6 (January 11, 2014): 831–36. http://dx.doi.org/10.1007/s00299-014-1565-z.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Csizinszky, A. A., D. J. Schuster, and J. E. Polston. "Effect of Ultraviolet-reflective Mulches on Tomato Yields and on the Silverleaf Whitefly." HortScience 34, no. 5 (August 1999): 911–14. http://dx.doi.org/10.21273/hortsci.34.5.911.

Full text
Abstract:
Field studies were conducted for three seasons, Fall 1994, Spring 1995, and Fall 1995, on the effect of ultraviolet (UV)-reflective films (mulches) on the silverleaf whitefly (Bemisia argentifolii Bellows and Perring), the incidence of tomato mottle virus (ToMoV), and on fruit yields of staked, fresh-market tomatoes (Lycopersicon esculentum Mill.). The UV-reflective mulches were metallized aluminum (aluminum) and painted aluminum (silver) on either black or white plastic film. The aluminum and silver mulches were evaluated with and without a white (fall) or black (spring) 25-cm-wide painted band in the bed center. Controls were the conventional white (fall) or black (spring) polyethylene mulches. Highest reflected energy (μmol·m–2·s–1) to the plants at 15 cm from the mulch surface was measured on the aluminum mulch with or without a white painted band. Lowest energy was reflected from the white or black controls and from silver on black mulches with or without a black painted band. Whitefly populations in the fall were lower (P ≤ 0.05) on the aluminum than on the silver mulches. In the spring, when whitefly populations were low, whiteflies were more numerous on the black control and silver on white, than on the aluminum mulches. In the fall seasons, the proportion of plants with symptoms of ToMoV transmitted by the silverleaf whitefly were higher on the controls than on the aluminum mulch. In the spring, the proportion of plants with symptoms was not affected by mulch treatments. Yields in the fall were similar with UV-reflective or white control mulches. In the spring, fruit size and marketable yields were greater (P ≤ 0.05) on plants with silver on white mulch than on the control black mulch.
APA, Harvard, Vancouver, ISO, and other styles
22

Wen, Ke, Huanting Pan, Xingang Li, Rong Huang, Qibin Ma, and Hai Nian. "Identification of an ATP-Binding Cassette Transporter Implicated in Aluminum Tolerance in Wild Soybean (Glycine soja)." International Journal of Molecular Sciences 22, no. 24 (December 9, 2021): 13264. http://dx.doi.org/10.3390/ijms222413264.

Full text
Abstract:
The toxicity of aluminum (Al) in acidic soil limits global crop yield. The ATP-binding cassette (ABC) transporter-like gene superfamily has functions and structures related to transportation, so it responds to aluminum stress in plants. In this study, one half-size ABC transporter gene was isolated from wild soybeans (Glycine soja) and designated GsABCI1. By real-time qPCR, GsABCI1 was identified as not specifically expressed in tissues. Phenotype identification of the overexpressed transgenic lines showed increased tolerance to aluminum. Furthermore, GsABCI1 transgenic plants exhibited some resistance to aluminum treatment by ion translocation or changing root components. This work on the GsABCI1 identified the molecular function, which provided useful information for understanding the gene function of the ABC family and the development of new aluminum-tolerant soybean germplasm.
APA, Harvard, Vancouver, ISO, and other styles
23

Simcoe, Charles R. "Aluminum: The Light Metal, Part II." AM&P Technical Articles 172, no. 10 (October 1, 2014): 32–33. http://dx.doi.org/10.31399/asm.amp.2014-10.p032.

Full text
Abstract:
Abstract The cost of electric power is the biggest expense involved in aluminum production. Proximity to economical power sources determined the location of major aluminum reduction plants. This article describes the development of the aluminum industry in the early part of the twentieth century.
APA, Harvard, Vancouver, ISO, and other styles
24

Erzsébet, Fogarasi, Fülöp Ibolya, Berecz Zsuzsa, Márton Kincső, and Croitoru Mircea Dumitru. "Aluminium contamination of several types of tea." Bulletin of Medical Sciences 92, no. 1 (July 1, 2019): 42–46. http://dx.doi.org/10.2478/orvtudert-2019-0003.

Full text
Abstract:
Abstract Introduction: Camellia sinensis, a widely used plant, optimally grows in a low pH soil that in most cases contains high amounts of aluminum. Objectives: The aluminum content of the tea obtained from Camellia sinensis and other plants was compared. The influence of pH on the aluminum content of the tea was also measured. Materials and methods: The aluminum content of 48 samples was measured using a colorimetric method. The method is based on the ability of aluminum to form a stable complex with xylenol orange at low pH; this complex has an absorption maximum of 555 nm. Results: The method was validated for tea obtained with water and for tea obtained with water containing citric acid. The method proved linear over the rage of 0.7 – 7 ug/ml, coefficient of variation ranged between 2.6 – 7.68% (was dependent on the pH of the solution used to obtain the tea), accuracy was suitable for quantitative measurement (92.39-102.92%) and the complex proved to be stable for at least 1 hour. The following concentrations were measured: green tea (1.59 - 7.70 µg/ml), black tea (1.39 - 5.60 µg/ml), fruit tea (1.01 - 5.63 µg/ml) and herbal tea (1.03 - 5.24 µg/ml). Conclusion: The method proved useful and easily applicable for screening aluminum content of plants used for tea brewing. Camellia sinensis both green and black types had significantly higher aluminum contents than other type of teas. Adding citric acid, as would result from use of lemon juice, significantly increased the aluminum extraction from the plants used for tea brewing.
APA, Harvard, Vancouver, ISO, and other styles
25

Haridasan, Mundayatan. "Nutritional adaptations of native plants of the cerrado biome in acid soils." Brazilian Journal of Plant Physiology 20, no. 3 (September 2008): 183–95. http://dx.doi.org/10.1590/s1677-04202008000300003.

Full text
Abstract:
Soils of the cerrado biome, mostly oxisols and deep sandy entisols, are acid, dystrophic and poor in available nutrients. These soils are not very different from soils that occur in the Amazon region. However, the open savanna physiognomies of cerrado with lower biomass of their different components are deficient in nutrients at the ecosystem level, unlike the Amazon forests which retain high nutrient reserves in their live biomass. Field crops are susceptible to aluminum and manganese toxicities, besides nutrient deficiencies, in cerrado soils and do not grow well in the absence of liming and fertilization. However, concepts of nutrient deficiencies and toxicities, well established for cultivated plants, should not be extended to native species in natural ecosystems, indiscriminately. Many native plants in the cerrado biome are resistant or tolerant to soil conditions deemed unfavorable for cultivated plants but their geographic distribution, frequency in native communities, growth and productivity are determined by water and nutrient availability and other edaphic conditions. Species growing on acid soils are aluminum tolerant or resistant, since their capacity to absorb essential nutrients, growth and reproduction is not affected by high aluminum levels in the soil. Many common species of the cerrado, instead of excluding aluminum, absorb and transport it to leaves and accumulate it in different tissues including leaves and seeds whereas others do not survive in the absence of exchangeable aluminum, even though no specific role of Al in plant metabolism is yet established.
APA, Harvard, Vancouver, ISO, and other styles
26

Watanabe, Toshihiro, Seiji Misawa, and Mitsuru Osaki. "Aluminum accumulation in the roots of Melastoma malabathricum, an aluminum-accumulating plant." Canadian Journal of Botany 83, no. 11 (November 2005): 1518–22. http://dx.doi.org/10.1139/b05-111.

Full text
Abstract:
Generally, plants that have Al levels of at least 1000 mg·kg–1 in their leaves are defined as Al accumulators. These plants are often found in very acid soils in the tropics. The mechanisms of Al uptake in strong Al accumulators are still unclear. In this study, we investigated the characteristics of Al uptake and accumulation in the roots of Melastoma malabathricum L., an Al-accumulating plant that grows in acidic soils in the tropics. Melastoma malabathricum roots hardly absorb any La, possibly because of lower affinity of the root apoplast to La than to Al. Exposure to La did not affect the concentration of citrate in the roots; however, application of Al increased the citrate level considerably, corresponding with the amount of Al accumulation in the symplast. 27Al NMR analysis revealed that Al complexes with oxalate, but not with citrate, in the roots of M. malabathricum. This investigation revealed that oxalate, which occurs constitutively at high concentrations, is a ligand for Al accumulation in both root and shoot tissue, and that citrate, the synthesis of which is induced by Al application, is a ligand mainly used for Al translocation from the roots to the shoots.
APA, Harvard, Vancouver, ISO, and other styles
27

Kurina, A. B., I. A. Kosareva, and A. M. Artemyeva. "Genetic diversity of VIR Raphanus sativus L. collections on aluminum tolerance." Vavilov Journal of Genetics and Breeding 24, no. 6 (October 28, 2020): 613–24. http://dx.doi.org/10.18699/vj20.655.

Full text
Abstract:
Radish and small radish (Raphanus sativus L.) are popular and widely cultivated root vegetables in the world, which occupy an important place in human nutrition. Edaphic stressors have a significant impact on their productivity and quality. The main factor determining the phytotoxicity of acidic soils is the increased concentration of mobile aluminum ions in the soil solution. The accumulation of aluminum in root tissues disrupts the processes of cell division, initiation and growth of the lateral roots, the supply of plants with minerals and water. The study of intraspecific variation in aluminum resistance of R. sativus is an important stage for the breeding of these crops. The purpose of this work was to study the genetic diversity of R. sativus crops including 109 accessions of small radish and radish of various ecological and geographical origin, belonging to 23 types, 14 varieties of European, Chinese and Japanese subspecies on aluminum tolerance. In the absence of a rapid assessment methodology specialized for the species studied, a method is used to assess the aluminum resistance of cereals using an eriochrome cyanine R dye, which is based on the recovery or absence of restoration of mitotic activity of the seedlings roots subjected to shock exposure to aluminum. The effect of various concentrations on the vital activity of plants was revealed: a 66-mM concentration of AlCl3 · 6Н2О had a weak toxic effect on R. sativus accessions slowing down root growth; 83 mM contributed to a large differentiation of the small radish accessions and to a lesser extent for radish; 99 mM inhibited further root growth in 13.0 % of small radish accessions and in 7.3 % of radish and had a highly damaging effect. AlCl3 · 6Н2О at a concentration of 99 mM allowed us to identify the most tolerant small radish and radish accessions that originate from countries with a wide distribution of acidic soils. In a result, it was possible to determine the intraspecific variability of small radish and radish plants in the early stages of vegetation and to identify genotypes that are contrasting in their resistance to aluminum. We recommend the AlCl3 · 6Н2О concentration of 83 mM for screening the aluminum resistance of small radish and 99 mM for radish. The modified method that we developed is proposed as a rapid diagnosis of aluminum tolerance for the screening of a wide range of R. sativus genotypes and a subsequent study of contrasting forms during a longer cultivation of plants in hydroponic culture (including elemental analysis of roots and shoots, contrasting in resistance of accessions) as well as reactions of plants in soil conditions.
APA, Harvard, Vancouver, ISO, and other styles
28

Sade, Hemalatha, Balaji Meriga, Varalakshmi Surapu, Jogeswar Gadi, M. S. L. Sunita, Prashanth Suravajhala, and P. B. Kavi Kishor. "Toxicity and tolerance of aluminum in plants: tailoring plants to suit to acid soils." BioMetals 29, no. 2 (January 21, 2016): 187–210. http://dx.doi.org/10.1007/s10534-016-9910-z.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Panda, Sanjib Kumar, and Hideaki Matsumoto. "Molecular Physiology of Aluminum Toxicity and Tolerance in Plants." Botanical Review 73, no. 4 (October 2007): 326–47. http://dx.doi.org/10.1663/0006-8101(2007)73[326:mpoata]2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Kharpukhaeva, T. M. "The lichens in aluminum plants vicinities of Irkutsk region." Проблемы ботаники южной сибири и монголии 20, no. 1 (September 22, 2021): 470–73. http://dx.doi.org/10.14258/pbssm.2021093.

Full text
Abstract:
The aim of our work was to assess the diversity of lichens in the forests polluted by the air emissions fromthe Bratsk (BRAZ), Irkutsk (IRKAZ) and Taishet (TAZ) aluminum smelters, as well as in the background (non-polluted)territories, similar to those by natural and forest-growing conditions. 102 wide-distributed lichen species were found onplots. This amount is due to the fact that the wood tier is sparse and forophytes composition is depleted. The lichen diversity varies in different types of vegetation and according to the influence of anthropogenic factors. Species diversity in vicinities of smelters and in buffer zones are similar, excluding control zones with dark-conifers forests. It’s established that as wemove closer to the aluminum plants – to the IRKAZ and especially BRAZ, the projective cover, the total number of speciesdecreases, and their species composition changes. Clear morphological changes of lichen thalloma were observed in theimpact zone of the BRAZ. Relatively resistant to atmospheric pollution lichens are found in the impact and buffer zonesin vicinities of plants. Lichens are absent in the crowns and trunks, but their abundance increases on bases of trees trunks.
APA, Harvard, Vancouver, ISO, and other styles
31

Romundstad, Pål, S. Andersen, and Tor Haldorsen. "Cancer incidence among workers in six Norwegian aluminum plants." Scandinavian Journal of Work, Environment & Health 26, no. 6 (December 2000): 461–69. http://dx.doi.org/10.5271/sjweh.569.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Romundstad, Pål, Aage Andersen, and Tor Haldorsen. "Nonmalignant mortality among workers in six Norwegian aluminum plants." Scandinavian Journal of Work, Environment & Health 26, no. 6 (December 2000): 470–75. http://dx.doi.org/10.5271/sjweh.570.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Armstrong, Ben, Claude Tremblay, and Gilles Theriault. "Compensating Bladder Cancer Victims Employed in Aluminum Reduction Plants." Journal of Occupational and Environmental Medicine 30, no. 10 (October 1988): 771–75. http://dx.doi.org/10.1097/00043764-198810000-00004.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Kochian, L. V. "Cellular Mechanisms of Aluminum Toxicity and Resistance in Plants." Annual Review of Plant Physiology and Plant Molecular Biology 46, no. 1 (June 1995): 237–60. http://dx.doi.org/10.1146/annurev.pp.46.060195.001321.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

ALVA, A. K., G. L. KERVEN, D. G. EDWARDS, and C. J. ASHER. "REDUCTION IN TOXIC ALUMINUM TO PLANTS BY SULFATE COMPLEXATION." Soil Science 152, no. 5 (November 1991): 351–59. http://dx.doi.org/10.1097/00010694-199111000-00006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Poschenrieder, Charlotte, Benet Gunsé, Isabel Corrales, and Juan Barceló. "A glance into aluminum toxicity and resistance in plants." Science of The Total Environment 400, no. 1-3 (August 2008): 356–68. http://dx.doi.org/10.1016/j.scitotenv.2008.06.003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

O.M., Lebedeva, Lebedev A.H., Fokin M.N., Lomakina S.V., Severorova T.A., M. N. Fokin, S. V. Lomakina, and T. A. Severorova. "Corrosion behaviour of aluminum-base alloys in desalination plants." Desalination 73 (January 1989): 457–58. http://dx.doi.org/10.1016/0011-9164(89)87032-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Mizuno, Kiyotaka, Wataru Takahashi, Toshio Beppu, Takiko Shimada, and Osamu Tanaka. "Aluminum borate whisker-mediated production of transgenic tobacco plants." Plant Cell, Tissue and Organ Culture 80, no. 2 (February 2005): 163–69. http://dx.doi.org/10.1007/s11240-004-9542-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Deflorian, F., S. Rossi, and M. Fedel. "Durability of aluminum cooling system in electric power plants." Surface and Interface Analysis 42, no. 4 (December 8, 2009): 269–74. http://dx.doi.org/10.1002/sia.3142.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Schalk, J. M., and M. LeRon Robbins. "Reflective Mulches Influence Plant Survival, Production, and Insect Control in Fall Tomatoes." HortScience 22, no. 1 (February 1987): 30–32. http://dx.doi.org/10.21273/hortsci.22.1.30.

Full text
Abstract:
Abstract Aluminum mulches lowered soil temperatures and reduced heat stress for young tomato transplants, increasing their survival and height. Yields were improved in all mulch treatments. Greatest yields of large and extra-large fruits were obtained from plants grown on aluminum alone, on aluminum in combination with black plastic, and on black plastic alone. Aphids were repelled by the aluminum mulches, while fruit injury increased due to tomato pinworm and tomato fruit worm feeding in these plots.
APA, Harvard, Vancouver, ISO, and other styles
41

Lux, Heidi B., and Jonathan R. Cumming. "Mycorrhizae confer aluminum resistance to tulip-poplar seedlings." Canadian Journal of Forest Research 31, no. 4 (April 1, 2001): 694–702. http://dx.doi.org/10.1139/x01-004.

Full text
Abstract:
Aluminum (Al) toxicity may limit the growth and nutrient acquisition of sensitive tree species in regions receiving acidic deposition. Symbioses between tree roots and mycorrhizal fungi may offset the negative impacts of Al in the root zone. Liriodendron tulipifera L. (tulip-poplar) is an important tree species in the Appalachian Mountains of the southeastern United States and may be at risk from the high levels of acidic deposition in that area. Mycorrhizal and non-mycorrhizal tulip-poplar seedlings were exposed to Al levels of 0, 50, 100, and 200 µM in sand culture for 6 weeks. Mycorrhizal plants accumulated two to seven times the shoot and root biomass of non-mycorrhizal plants and demonstrated no decreases in biomass with Al exposure. Non-mycorrhizal plants exhibited significant reductions in biomass at and above 100 µM Al. Aluminum toxicity in non-mycorrhizal plants appears to be the result of the disruption of P translocation to leaves and Ca, Mg, P, Cu, and Zn uptake in roots. Mycorrhizal plants accumulated 2 and 1.5 times the concentration of Al in shoots and roots, respectively, indicating that Al resistance was not associated with the exclusion of Al from the plant. Patterns of labile Al in solution, nutrients, and Al accumulation in tissues suggest that arbuscular mycorrhizal fungal ecotypes may alter the form or compartmentation of Al within the rhizosphere and plant, thus protecting seedlings from the effects of exposure to Al in the soil solution.
APA, Harvard, Vancouver, ISO, and other styles
42

Almeida, Aline Mansur, Eduardo Yukio Basílio Wada, and Julio Cesar Wasserman. "Volumetric modeling of two sludge piles from water treatment plants in a Brazilian reservoir." Water Science and Technology 77, no. 2 (October 27, 2017): 355–63. http://dx.doi.org/10.2166/wst.2017.515.

Full text
Abstract:
Abstract Water treatment plants are designed to continuously produce drinkable water, meeting defined criteria of potability. However, besides potable water, these plants produce sludges that are disposed of in the environment. The present work aimed to evaluate the sludges generated in two water treatment plants and disposed of in the margin of the Juturnaíba dam. Since alum has been used as a flocculating agent in these two plants, the concentrations of aluminum were measured in the sludges and in surface sediments. The generated piles are extremely soft to walk on and difficult to measure, so indirect modeling procedures had to be applied. The calculated mass of the sludge piles at each plant are similar and respectively 60,370 and 61,479 tons. The aluminum content of the residues, calculated according to its dosage, was 33.2 and 32.6 g kg−1 in the piles from the two plants. The amount of alum dosed to the water corresponds almost to the excess of aluminum in the sludge, compared to the sediments. It was concluded that regardless of the fact that residues are disposed of in very restricted areas, they are directly in contact with the water and may constitute a threat for the environment and humans’ health.
APA, Harvard, Vancouver, ISO, and other styles
43

SHEVCHENKO, Nadiia, Tetiana MIROSHNICHENKO, Anna MOZGOVSKA, Nataliia BASHTAN, Galyna KOVALENKO, and Tetiana IVCHENKO. "Field performance of cryopreserved seed-derived tomato plants and post-thaw survival of viral-infected meristems." Acta agriculturae Slovenica 118, no. 4 (December 30, 2022): 1. http://dx.doi.org/10.14720/aas.2022.118.4.1823.

Full text
Abstract:
<p class="042abstractstekst">The effectiveness of different cryopreservation techniques of tomato meristems isolated from viral-infected plants ‘Irishka’ cultivar was determined. The pieces of stem were protected with dimethyl sulfoxide and propylene glycol and cooled in vapour phase of liquid nitrogen (–170 °C). For the vitrification and droplet-vitrification protocols, the meristems were treated with loading solution and dehydrated with different plant vitrification solutions (PVS1 modified, PVS2, 88 % PVS3, PVSN). The samples were placed to sterilized aluminum foil pieces, in 1.2 ml cryovials or in 50 µl aluminum pans for differential scanning calorimetry and were directly immersed into liquid nitrogen. Acсording to the dehydration technique, the meristems were dehydrated with sterile airflow for 120 min. The post-thaw survival rates of meristems (from 34.2 to 78.5 %) were observerved only for 50 µl aluminum pans and airflow dehydration. We determined the productivity of plants, obtaned from cryopreserved seeds (‘Seven’, ‘Potiron Ecarlate’ and ‘Druzhba’ cultivars). We observed increasing in total and marketable yields for the plants grown from the cryopreserved seeds for all the cultivars. Total number of diseased plants decreased by 33 % for ‘Seven’, for ‘Potiron Ecarlate’ it did by 6.7 %, for that of ‘Druzhba’ the total percentage of sick and healthy plants did not differ after seeds cryopreservation.</p>
APA, Harvard, Vancouver, ISO, and other styles
44

Alva, A. K., F. P. C. Blamey, D. G. Edwards, and C. J. Asher. "An evaluation of aluminum indices to predict aluminum toxicity to plants grown in nutrient solutions." Communications in Soil Science and Plant Analysis 17, no. 12 (December 1986): 1271–80. http://dx.doi.org/10.1080/00103628609367789.

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

Westberg, Håkan B., Anders I. Seldén, and Tom Bellander. "Exposure to Chemical Agents in Swedish Aluminum Foundries and Aluminum Remelting Plants?A Comprehensive Survey." Applied Occupational and Environmental Hygiene 16, no. 1 (January 2001): 66–77. http://dx.doi.org/10.1080/104732201456140.

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

Cunha, Gabriel Octávio de Mello, Jaime Antonio de Almeida, Clovis Arruda Souza, and Paulo Roberto Ernani. "Nutritional Efficiency and Nutrient Acumulation of Maize Cultivated in Soils With High Al-KCl Content." Journal of Agricultural Science 11, no. 16 (September 30, 2019): 107. http://dx.doi.org/10.5539/jas.v11n16p107.

Full text
Abstract:
Most Brazilian soils are weathered and acidic, generally with exchangeable aluminum levels toxic to plants. Aluminum content can exceed 10 cmolc kg-1 in certain soils, but does not interfere in plant nutrient absorption, transport and accumulation. This study aimed to assess whether the high aluminum levels extracted with 1 mol L-1 KCl solution affect the nutritional efficiency of maize plants grown in acidic soils in areas of Brazilian states (AC: profile 9, PE, SC and RS) never before cultivated. To that end, two greenhouse experiments were performed (soybean: 45 days and maize: 60 days) with increasing levels of lime (0; 25; 50 and 100% H+Al content). After the experiments and collection, chemical analyses were conducted to characterize the soils and determine the chemical composition of the maize plants. The Al levels did not interfere in maize plants, nutrient absorption, transport, use or accumulation in AC9 and Rosario do Sul (RS) soils, but did affect those cultivated in PE, BR and CB soils in treatments without lime. The absence of Al toxicity is related to the increased concentration and activity of basic cations and P in the soil solution.
APA, Harvard, Vancouver, ISO, and other styles
47

Dubinin, P. S., I. S. Yakimov, A. S. Samoilo, S. G. Ruzhnikov, O. E. Bezrukova, A. N. Zaioga, S. D. Kirik, and D. V. Khiystov. "Analytical appro­ aches in the development of industry standard specimens of aluminum production electrolyte." Industrial laboratory. Diagnostics of materials 88, no. 10 (October 24, 2022): 20–29. http://dx.doi.org/10.26896/1028-6861-2022-88-10-20-29.

Full text
Abstract:
A combination of the methods of X-ray phase and X-ray spectral analyzes is used at domestic aluminum plants for operational technological control of the composition of cooled electrolytes. In this case, standard samples of chemical and phase composition are used to calibrate measuring instruments. The synthesis of standard samples from simple components is impossible due to the inadequacy of their microcrystalline structure to real electrolyte samples. Therefore, it is necessary to develop standard samples directly from the material of real electrolytes with a reliably established quantitative chemical and mineralogical phase composition. We managed to develop a set of 30 standard samples of aluminum-produced electrolyte using electrolytes taken from the electrolysis baths of various plants; some of the samples were doped with so­dium, aluminum, calcium, and magnesium fluorides to expand the range of compositions. A metrological certification of the set with the status of "Industry standard samples" was performed based on the data of interlaboratory analysis according to the methods of X-ray control used at the plants and according to the well-known Rietveld X-ray phase method for determining the quantitative phase composition. The set has been successfully implemented at seven RUSAL plants.
APA, Harvard, Vancouver, ISO, and other styles
48

Li, Gang, Jingjing Liu, Yu Tian, Han Chen, and Haiying Ren. "Investigation and Analysis of Rhizosphere Soil of Bayberry-Decline-Disease Plants in China." Plants 11, no. 23 (December 6, 2022): 3394. http://dx.doi.org/10.3390/plants11233394.

Full text
Abstract:
The rampant bayberry decline disease has been regarded as related to soil with the long-term plantation bayberry. These parameters, hydrogen, aluminum, other alkali cations, and plant-related nutrients, were measured from the soil around diseased tree roots 10, 20, and 30 years old. The pH significantly declined in topsoil with increasing tree age and rose with increasing depth of the soil layer with an age of 10, 20, and 30 years. The concentration of exchangeable aluminum has risen significantly with the increase of the tree ages in the top soil layer and also in 0 to 40 cm soils layer with ten-year-old trees. In the top soil layer with a depth of 0 to 10 cm, the cation concentrations of Ca2+, Mg2+, and K+ has fallen significantly with the increase of tree ages. A higher concentration of exchangeable aluminum was observed in the soil with trees more seriously affected by the disease and was accompanied with lower concentrations of Ca2+, Mg2+, and K+. The correlation analysis showed that the soil pH is significantly positively related to the concentration of exchangeable Ca2+, total nitrogen, and total phosphorus and negatively to exchangeable aluminum. These findings provided a new insight to mitigate the disease by regulating the soil parameters.
APA, Harvard, Vancouver, ISO, and other styles
49

Souza, José Roberto Pinto de, Maria Fátima de Guimarães, Marcelo Marques Lopes Muller, and Carlos Henrique dos Santos. "Morphological alterations of corn (Zea mays L.) plants in response to aluminum toxicity in the soil." Brazilian Archives of Biology and Technology 43, no. 4 (2000): 415–20. http://dx.doi.org/10.1590/s1516-89132000000400010.

Full text
Abstract:
The objective of the experiment was to identify the morphological alterations in corn genotypes in response to aluminum toxicity in the soil. A complete randomized block design with five replications was used. The a factorial scheme was composed of two corn genotypes (C525M - tolerant, and HS701B - sensitive) and two neutralization levels (0% and 100% of aluminum saturation). The evaluations were performed at six leaf and 10-11 leaf growth stages, emergence of the "stigma-style", and the physiological maturity. The presence of toxic aluminum didn’t significantly reduce diameter and height of stem, leaf area, dry matter of aerial parts, total dry matter and yield. Stem diameter, leaf area, dry matter of root, dry matter of aerial parts, total dry matter and yield did not allow the separation of genotypes in relation to aluminum toxicity in the soil.
APA, Harvard, Vancouver, ISO, and other styles
50

Rodrigues, Gustavo Araújo, Breno De Jesus Pereira, Anacleto Ranulfo dos Santos, Francielle Medeiros Costa, and Gilvanda Leão dos Anjos. "PIGEON PEA INITIAL GROWTH UNDER DIFFERENT CONCENTRATIONS OF ALUMINUM AND COMMERCIAL SUBSTRATE." JOURNAL OF NEOTROPICAL AGRICULTURE 7, no. 4 (November 30, 2020): 1–6. http://dx.doi.org/10.32404/rean.v7i4.5046.

Full text
Abstract:
The incorporation of organic substrate into the soil can reduce aluminum toxicity in pigeon pea (Cajanus cajan (L.) Millsp.) plants, making it a viable alternative for use by small farmers in acidic soil regions such as the northeast semi-arid regions of Brazil that contain toxic aluminum in the soil. The aim of this study was to evaluate the effect of increasing aluminum doses and different commercial substrate concentrations on the initial growth of the pigeon pea. The experiment was undertaken in a greenhouse using a completely randomized design with a 5 × 3 factorial scheme. The first factor consisted of five doses of aluminum (0, 13.5, 27, 54, and 108 mg L-1) and the second factor consisted of three percentages of commercial substrate Vivato Slim Pro® (0%, 10%, and 20%), with five replicates per treatment, totaling 60 experimental units. The variables evaluated were plant height; total chlorophyll; shoot, root, and total dry matter; and root volume. The use of the commercial substrate attenuated the aluminum toxicity and favored the initial growth of pigeon pea plants, regardless of the concentration used. Thus, the aluminum was detrimental to the initial plant growth at all tested concentrations and showed more pronounced signs of toxicity on the root volume.
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