Journal articles on the topic 'Malate'

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

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 'Malate.'

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

Liu, Kun, Ying Xu, and Ning-Yi Zhou. "Identification of a Specific Maleate Hydratase in the Direct Hydrolysis Route of the Gentisate Pathway." Applied and Environmental Microbiology 81, no. 17 (June 12, 2015): 5753–60. http://dx.doi.org/10.1128/aem.00975-15.

Full text
Abstract:
ABSTRACTIn contrast to the well-characterized and more common maleylpyruvate isomerization route of the gentisate pathway, the direct hydrolysis route occurs rarely and remains unsolved. InPseudomonas alcaligenesNCIMB 9867, two gene clusters,xlnandhbz, were previously proposed to be involved in gentisate catabolism, and HbzF was characterized as a maleylpyruvate hydrolase converting maleylpyruvate to maleate and pyruvate. However, the complete degradation pathway of gentisate through direct hydrolysis has not been characterized. In this study, we obtained from the NCIMB culture collection aPseudomonas alcaligenesspontaneous mutant strain that lacked thexlncluster and designated the mutant strain SponMu. Thehbzcluster in strain SponMu was resequenced, revealing the correct location of the stop codon forhbzIand identifying a new gene,hbzG. HbzIJ was demonstrated to be a maleate hydratase consisting of large and small subunits, stoichiometrically converting maleate to enantiomerically pured-malate. HbzG is a glutathione-dependent maleylpyruvate isomerase, indicating the possible presence of two alternative pathways of maleylpyruvate catabolism. However, thehbzF-disrupted mutant could still grow on gentisate, while disruption ofhbzGprevented this ability, indicating that the direct hydrolysis route was not a complete pathway in strain SponMu. Subsequently, ad-malate dehydrogenase gene was introduced into thehbzG-disrupted mutant, and the engineered strain was able to grow on gentisate via the direct hydrolysis route. This fills a gap in our understanding of the direct hydrolysis route of the gentisate pathway and provides an explanation for the high yield ofd-malate from maleate by thisd-malate dehydrogenase-deficient natural mutant.
APA, Harvard, Vancouver, ISO, and other styles
2

Asano, Y., M. Ueda, and H. Yamada. "Microbial production of D-malate from maleate." Applied and Environmental Microbiology 59, no. 4 (1993): 1110–13. http://dx.doi.org/10.1128/aem.59.4.1110-1113.1993.

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

Kimura, Takuhei, Yasuro Kawabata, and Eiji Sato. "Enzymatic Production ofl-Malate from Maleate byAlcaligenessp." Agricultural and Biological Chemistry 50, no. 1 (January 1986): 89–94. http://dx.doi.org/10.1080/00021369.1986.10867349.

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

Báthori, Nikoletta B., Ayesha Jacobs, Mawonga Mei, and Luigi R. Nassimbeni. "Resolution of malic acid by (+)-cinchonine and (–)-cinchonidine." Canadian Journal of Chemistry 93, no. 8 (August 2015): 858–63. http://dx.doi.org/10.1139/cjc-2014-0579.

Full text
Abstract:
(+)-Cinchonine and (–)-cinchonidine have been employed to resolve rac-malic acid. The resulting salts contain the D-malate anion in both cases. The cinchoninium and cinchonidinium L-malates were also crystallised, and the structures of all four salts were analysed in terms of their nonbonding interactions. All four structures display extensive hydrogen bonding, and it is shown that the D-malate salts are more efficiently packed.
APA, Harvard, Vancouver, ISO, and other styles
5

van der Werf, M. J., W. J. van den Tweel, and S. Hartmans. "Screening for microorganisms producing D-malate from maleate." Applied and Environmental Microbiology 58, no. 9 (1992): 2854–60. http://dx.doi.org/10.1128/aem.58.9.2854-2860.1992.

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

Cada, Dennis J., Terri Levien, and Danial E. Baker. "Almotriptan Malate." Hospital Pharmacy 36, no. 10 (October 2001): 1066–78. http://dx.doi.org/10.1177/001857870103601010.

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

McIntyre, J. A., and J. Castañer. "Sunitinib Malate." Drugs of the Future 30, no. 8 (2005): 785. http://dx.doi.org/10.1358/dof.2005.030.08.928476.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Izzedine, Hassane, Irina Buhaescu, Olivier Rixe, and Gilbert Deray. "Sunitinib malate." Cancer Chemotherapy and Pharmacology 60, no. 3 (November 30, 2006): 357–64. http://dx.doi.org/10.1007/s00280-006-0376-5.

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

KIMURA, Takuhei, Yasuro KAWABATA, and Eiji SATO. "Enzymatic production of L-malate from maleate by Alcaligenes sp." Agricultural and Biological Chemistry 50, no. 1 (1986): 89–94. http://dx.doi.org/10.1271/bbb1961.50.89.

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

Tunchai, Mattana, Akiko Hida, Shota Oku, Takahisa Tajima, and Junichi Kato. "Chemotactic disruption as a method to control bacterial wilt caused by Ralstonia pseudosolanacearum." Bioscience, Biotechnology, and Biochemistry 85, no. 3 (December 31, 2020): 697–702. http://dx.doi.org/10.1093/bbb/zbaa065.

Full text
Abstract:
ABSTRACT We have demonstrated that chemotaxis to l-malate facilitated motility of Ralstonia pseudosolanacearum MAFF 106611, a causative agent of bacterial wilt, to plant roots. Here, we evaluated the assumption that the disruption of chemotaxis to l-malate leads to inhibition of plant infection by R. pseudosolanacearum MAFF 106611. Chemotactic assays revealed that chemotaxis to l-malate was completely or partially inhibited in the presence of l-, d-, and dl-malate, respectively. Moreover, l-malate served as a carbon and energy source for R. pseudosolanacearum MAFF 106611, while d-malate inhibited the growth of this bacterium. In the sand-soak inoculation virulence assay for tomato plants, the addition of l-, d-, and dl-malate to sand suppressed the plant infection. We concluded that supplementation of l- and dl-malate suppresses tomato plant infection with R. pseudosolanacearum MAFF 106611 by disrupting its chemotaxis to l-malate, while d-malate suppresses it by both the disruption of l-malate chemotaxis and inhibition of growth.
APA, Harvard, Vancouver, ISO, and other styles
11

Landete, José María, Sergi Ferrer, Vicente Monedero, and Manuel Zúñiga. "Malic Enzyme and Malolactic Enzyme Pathways Are Functionally Linked but Independently Regulated in Lactobacillus casei BL23." Applied and Environmental Microbiology 79, no. 18 (July 8, 2013): 5509–18. http://dx.doi.org/10.1128/aem.01177-13.

Full text
Abstract:
ABSTRACTLactobacillus caseiis the only lactic acid bacterium in which two pathways forl-malate degradation have been described: the malolactic enzyme (MLE) and the malic enzyme (ME) pathways. Whereas the ME pathway enablesL. caseito grow onl-malate, MLE does not support growth. Themlegene cluster consists of three genes encoding MLE (mleS), the putativel-malate transporter MleT, and the putative regulator MleR. Themaegene cluster consists of four genes encoding ME (maeE), the putative transporter MaeP, and the two-component system MaeKR. Since both pathways compete for the same substrate, we sought to determine whether they are coordinately regulated and their role inl-malate utilization as a carbon source. Transcriptional analyses revealed that themleandmaegenes are independently regulated and showed that MleR acts as an activator and requires internalization ofl-malate to induce the expression ofmlegenes. Notwithstanding, bothl-malate transporters were required for maximall-malate uptake, although only anmleTmutation caused a growth defect onl-malate, indicating its crucial role inl-malate metabolism. However, inactivation of MLE resulted in higher growth rates and higher final optical densities onl-malate. The limited growth onl-malate of the wild-type strain was correlated to a rapid degradation of the availablel-malate tol-lactate, which cannot be further metabolized. Taken together, our results indicate thatL. caseil-malate metabolism is not optimized for utilization ofl-malate as a carbon source but for deacidification of the medium by conversion ofl-malate intol-lactate via MLE.
APA, Harvard, Vancouver, ISO, and other styles
12

Almuiabed, Ala'ddin M., and Alan Townshend. "Flow-injection determination of malate with immobilized malate dehydrogenase." Analytica Chimica Acta 221 (1989): 337–40. http://dx.doi.org/10.1016/s0003-2670(00)81970-4.

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

Li, Haigang, Fusuo Zhang, Zed Rengel, and Jianbo Shen. "Rhizosphere properties in monocropping and intercropping systems between faba bean (Vicia faba L.) and maize (Zea mays L.) grown in a calcareous soil." Crop and Pasture Science 64, no. 10 (2013): 976. http://dx.doi.org/10.1071/cp13268.

Full text
Abstract:
The processes involving pH modification, carboxylate exudation and phosphorus (P) dynamics in the rhizosphere of crops grown in intercropping are poorly understood. Two groups of maize (Zea mays L.) or faba bean (Vicia faba L.) plants (monocropping) or one group of plant of each species (intercropping) were grown between three 1-mm-thick soil layers; the central soil layer is referred to as inter-rhizosphere, and the two outer soil layers are designated sole-rhizosphere. Faba bean intercropped with maize had an 11% increase in shoot biomass and a 15% increase in P uptake compared with monocropped faba bean. The cropping pattern did not significantly influence maize growth. After 4 weeks of growth, faba bean significantly decreased soil pH in both the sole- and inter-rhizosphere in monocropping, but no effects were apparent for the intercropping rhizosphere. The major carboxylates in the rhizosphere of faba bean were malate (18–45 nmol g–1 soil) and maleate (1.2–2.4 nmol g–1 soil). Only trace amounts of carboxylates were measured in the rhizosphere of monocropped maize. However, intercropped maize had a high concentration of malate (~11 nmol g–1 soil) in both sole- and inter-rhizosphere; the malate was likely exuded by faba bean and was then diffused to the sole-rhizosphere of intercropped maize. The amount of malate exuded by intercropped faba bean was 19% higher than with monocropped plants. The results indicate that diffusion of protons and carboxylates extended the interaction zone between maize and faba bean, and may have contributed to enhancements of P uptake in the intercropping system.
APA, Harvard, Vancouver, ISO, and other styles
14

Mandair, Ravneet, Pinar Karagoz, and Roslyn M. Bill. "A Redox-Neutral, Two-Enzyme Cascade for the Production of Malate and Gluconate from Pyruvate and Glucose." Applied Sciences 11, no. 11 (May 26, 2021): 4877. http://dx.doi.org/10.3390/app11114877.

Full text
Abstract:
A triple mutant of NADP(H)-dependent malate dehydrogenase from thermotolerant Thermococcus kodakarensis has an altered cofactor preference for NAD+, as well as improved malate production compared to wildtype malate dehydrogenase. By combining mutant malate dehydrogenase with glucose dehydrogenase from Sulfolobus solfataricus and NAD+/NADH in a closed reaction environment, gluconate and malate could be produced from pyruvate and glucose. After 3 h, the yield of malate was 15.96 mM. These data demonstrate the feasibility of a closed system capable of cofactor regeneration in the production of platform chemicals.
APA, Harvard, Vancouver, ISO, and other styles
15

Lukas, Hanna, Julia Reimann, Ok Bin Kim, Jan Grimpo, and Gottfried Unden. "Regulation of Aerobic and Anaerobic d-Malate Metabolism of Escherichia coli by the LysR-Type Regulator DmlR (YeaT)." Journal of Bacteriology 192, no. 10 (March 16, 2010): 2503–11. http://dx.doi.org/10.1128/jb.01665-09.

Full text
Abstract:
ABSTRACT Escherichia coli K-12 is able to grow under aerobic conditions on d-malate using DctA for d-malate uptake and the d-malate dehydrogenase DmlA (formerly YeaU) for converting d-malate to pyruvate. Induction of dmlA encoding DmlA required an intact dmlR (formerly yeaT) gene, which encodes DmlR, a LysR-type transcriptional regulator. Induction of dmlA by DmlR required the presence of d-malate or l- or meso-tartrate, but only d-malate supported aerobic growth. The regulator of general C4-dicarboxylate metabolism (DcuS-DcuR two-component system) had some effect on dmlA expression. The anaerobic l-tartrate regulator TtdR or the oxygen sensors ArcB-ArcA and FNR did not have a major effect on dmlA expression. DmlR has a high level of sequence identity (49%) with TtdR, the l- and meso-tartrate-specific regulator of l-tartrate fermentation in E. coli. dmlA was also expressed at high levels under anaerobic conditions, and the bacteria had d-malate dehydrogenase activity. These bacteria, however, were not able to grow on d-malate since the anaerobic pathway for d-malate degradation has a predicted yield of ≤0 ATP/mol d-malate. Slow anaerobic growth on d-malate was observed when glycerol was also provided as an electron donor, and d-malate was used in fumarate respiration. The expression of dmlR is subject to negative autoregulation. The network for regulation and coordination of the central and peripheral pathways for C4-dicarboxylate metabolism by the regulators DcuS-DcuR, DmlR, and TtdR is discussed.
APA, Harvard, Vancouver, ISO, and other styles
16

Perpétuo, Genivaldo Julio, and Jan Janczak. "Anilinium monohydrogenDL-malate." Acta Crystallographica Section C Crystal Structure Communications 59, no. 12 (November 30, 2003): o709—o711. http://dx.doi.org/10.1107/s0108270103023837.

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

Trelease, Richard N., Cheryl A. Hermerath, Rickie B. Turley, and Christine M. Kunce. "Cottonseed Malate Synthase." Plant Physiology 84, no. 4 (August 1, 1987): 1343–49. http://dx.doi.org/10.1104/pp.84.4.1343.

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

Turley, Rickie B., and Richard N. Trelease. "Cottonseed Malate Synthase." Plant Physiology 84, no. 4 (August 1, 1987): 1350–56. http://dx.doi.org/10.1104/pp.84.4.1350.

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

Cole, Erica, Erin Flores, Madeline Basile, Ashini Jayasinghe, Joshua de Groot, Daniel K. Unruh, and Tori Z. Forbes. "Directing dimensionality in uranyl malate and copper uranyl malate compounds." Polyhedron 114 (August 2016): 378–84. http://dx.doi.org/10.1016/j.poly.2016.01.030.

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

Bodson, M. J., W. H. Outlaw, and S. H. Silvers. "Malate content of picoliter samples of Raphanus sativus cytoplasm." Journal of Histochemistry & Cytochemistry 39, no. 4 (April 1991): 435–40. http://dx.doi.org/10.1177/39.4.2005372.

Full text
Abstract:
Malate, which plays many essential roles in plant metabolism, is a potent in vitro inhibitor of the cytosolic enzyme phosphoenolpyruvate carboxylase (PEPC). Because PEPC activity leads to malate biosynthesis, malate is assumed to attenuate its own synthesis in situ. To test this hypothesis, we measured directly the malate content of picoliter samples of Raphanus root-hair cytoplasm using quantitative histochemical techniques. We also obtained an estimate for malate accumulation in these cells. These values were compared with the PEPC activity of individual root hairs (less than 2 ng). The results indicate that high cytoplasmic malate concentration does not severely inhibit PEPC in situ. We suggest that the focus for studies on the regulation of organic anion accumulation be on the interactive effects of malate and other PEPC effectors.
APA, Harvard, Vancouver, ISO, and other styles
21

Zelle, Rintze M., Erik de Hulster, Wouter A. van Winden, Pieter de Waard, Cor Dijkema, Aaron A. Winkler, Jan-Maarten A. Geertman, Johannes P. van Dijken, Jack T. Pronk, and Antonius J. A. van Maris. "Malic Acid Production by Saccharomyces cerevisiae: Engineering of Pyruvate Carboxylation, Oxaloacetate Reduction, and Malate Export." Applied and Environmental Microbiology 74, no. 9 (March 14, 2008): 2766–77. http://dx.doi.org/10.1128/aem.02591-07.

Full text
Abstract:
ABSTRACT Malic acid is a potential biomass-derivable “building block” for chemical synthesis. Since wild-type Saccharomyces cerevisiae strains produce only low levels of malate, metabolic engineering is required to achieve efficient malate production with this yeast. A promising pathway for malate production from glucose proceeds via carboxylation of pyruvate, followed by reduction of oxaloacetate to malate. This redox- and ATP-neutral, CO2-fixing pathway has a theoretical maximum yield of 2 mol malate (mol glucose)−1. A previously engineered glucose-tolerant, C2-independent pyruvate decarboxylase-negative S. cerevisiae strain was used as the platform to evaluate the impact of individual and combined introduction of three genetic modifications: (i) overexpression of the native pyruvate carboxylase encoded by PYC2, (ii) high-level expression of an allele of the MDH3 gene, of which the encoded malate dehydrogenase was retargeted to the cytosol by deletion of the C-terminal peroxisomal targeting sequence, and (iii) functional expression of the Schizosaccharomyces pombe malate transporter gene SpMAE1. While single or double modifications improved malate production, the highest malate yields and titers were obtained with the simultaneous introduction of all three modifications. In glucose-grown batch cultures, the resulting engineered strain produced malate at titers of up to 59 g liter−1 at a malate yield of 0.42 mol (mol glucose)−1. Metabolic flux analysis showed that metabolite labeling patterns observed upon nuclear magnetic resonance analyses of cultures grown on 13C-labeled glucose were consistent with the envisaged nonoxidative, fermentative pathway for malate production. The engineered strains still produced substantial amounts of pyruvate, indicating that the pathway efficiency can be further improved.
APA, Harvard, Vancouver, ISO, and other styles
22

Tejido, M. L., M. J. Ranilla, R. García-Martínez, and M. D. Carro. "In vitromicrobial growth and rumen fermentation of different substrates as affected by the addition of disodium malate." Animal Science 81, no. 1 (August 2005): 31–38. http://dx.doi.org/10.1079/asc42060031.

Full text
Abstract:
AbstractThe effects of two concentrations of disodium malate on thein vitrofermentation of three substrates differing in their forage: concentrate ratio (0·8: 0·2, 0·5: 0·5 and 0·2: 0·8; g/g dry matter; low-, medium- and high-concentrate substrates, respectively) by rumen micro-organisms were studied using batch cultures. Rumen contents were collected from four Merino sheep offered lucerne hay ad libitum and supplemented daily with 400 g concentrate. Disodium malate was added to the incubation bottles to achieve final concentrations of 0, 4 and 8 mmol/l malate and15N was used as a microbial marker. Gas production was measured at regular intervals from 0 to 120 h of incubation to study fermentation kinetics. When gas production values were corrected for gas released from added malate, no effects (P> 0·05) of malate were detected for any of the estimated gas production parameters. In 17-h incubations, the final pH and total volatile fatty acid (VFA) production were increased (P< 0·001) by the addition of malate, but no changes (P> 0·05) were detected in the final amounts of ammonia-N and lactate. When net VFA productions were corrected for the amount of VFA produced from malate fermentation itself, adding malate did not affect (P> 0·05) the production of acetate, propionate and total VFA. Malate reduced methane (CH4) production by proportionately 0·058, 0·013 and 0·054 for the low-, medium- and high-concentrate substrates, respectively. Adding malate to batch cultures increased (P< 0·01) rumen microbial growth (mean values of 16·6, 18·3 and 18·4 mg of microbial N for malate at 0, 4 and 8 mmol/l, respectively), but did not affect (P> 0·05) its efficiency of growth (55·5, 56·7 and 54·3 mg of microbial N per g of organic matter apparently fermented for malate at 0, 4 and 8 mmol/l, respectively). There were no interactions (P> 0·05) malate × substrate for any of the measured variables, and no differences (P> 0·05) in pH, CH4production and microbial growth were found between malate at 4 and 8 mmol/l. The results indicate that malate had a beneficial effect on in vitro rumen fermentation of substrates by increasing VFA production and microbial growth, and that only subtle differences in the effects of malate were observed between substrates. Most of the observed effects, however, seem to be due to fermentation of malate itself.
APA, Harvard, Vancouver, ISO, and other styles
23

Carro, M. D., and M. J. Ranilla. "Effect of the addition of malate onin vitrorumen fermentation of cereal grains." British Journal of Nutrition 89, no. 2 (February 2003): 181–88. http://dx.doi.org/10.1079/bjn2002759.

Full text
Abstract:
Batch cultures of mixed rumen micro-organisms were used to study the effects of different concentrations of malate (Rumalato®; Norel & Nature S.A., Barcelona, Spain; composed of disodium malate–calcium malate (0·16:0·84, w/w)) on the fermentation of four cereal grains (maize, barley, wheat and sorghum). Rumen contents were collected from four Merino sheep fed lucerne hayad libitumand supplemented with 300 g concentrate/d. Rumalato® was added to the incubation bottles to achieve final concentrations of 0, 4, 7 and 10 mM-MALATE. Gas production was measured at regular intervals up to 120 h. Malate increased (P<0·01) the average fermentation rate of all substrates, and the lag time decreased (P<0·05) linearly with increasing concentrations of malate for all substrates, with the exception of sorghum. in 17 h incubations, the final pH and total volatile fatty acid production increased (P<0·001) linearly for all substrates as malate concentration increased from 0 TO 10 mM. Propionate and butyrate production increased (P<0·05), while the value of the acetate: propionate ratio and L-lactate concentrations decreased (P<0·05) linearly with increasing doses of malate. Malate treatment increased (P<0·05) the CO2production and decreased the production of CH4, although this effect was not significant (P>0·05) for maize. Malate at 4 and 7 mm increased (P<0·05) optical density of the cultures measured at 600 nm for maize, with no differences for the other substrates. The results indicate that malate may be used as a feed additive for ruminant animals fed high proportions of cereal grains, because it increased pH and propionate production and decreased CH4production and L-lactate concentrations; however, in general, no beneficial effects of 10 compared with 7 mM-malate were observed.
APA, Harvard, Vancouver, ISO, and other styles
24

Agbanyo, F. R., G. Moses, and N. F. Taylor. "L-Malate transport and proton symport in vesicles prepared from Pseudomonas putida." Biochemistry and Cell Biology 64, no. 11 (November 1, 1986): 1190–94. http://dx.doi.org/10.1139/o86-156.

Full text
Abstract:
In vesicles from glucose-grown Pseudomonas putida, L-malate is transported by nonspecific physical diffusion. L-Malate also acts as an electron donor and generates a proton motive force (Δp) of 129 mV which is composed of a membrane potential (Δψ) of 60 mV and a ΔpH of 69 mV. In contrast, vesicles from succinate-grown cells (a) transport L-malate by a carrier-mediated system with a Km value of 14.3 mM and a Vmax of 313 nmol∙mg protein−1∙min−1, (b) generate no Δψ, ΔpH, or Δp when L-malate is the electron donor, and (c) produce an extravesicular alkaline pH during the transport of L-malate. A kinetic analysis of this L-malate-induced proton transport gives a Km value of 16 mM and a Vmax of 667 nmol H+∙mg protein−1∙min−1. This corresponds to a H+/L-malate ratio of 2.1. The failure to generate a Δp in these vesicles is considered, therefore, to be consistent with the induction in succinate-grown cells of an electrogenic proton symport L-malate transport system.
APA, Harvard, Vancouver, ISO, and other styles
25

Zhang, Lihua, Baiquan Ma, Changzhi Wang, Xingyu Chen, Yong-Ling Ruan, Yangyang Yuan, Fengwang Ma, and Mingjun Li. "MdWRKY126 modulates malate accumulation in apple fruit by regulating cytosolic malate dehydrogenase (MdMDH5)." Plant Physiology 188, no. 4 (January 25, 2022): 2059–72. http://dx.doi.org/10.1093/plphys/kiac023.

Full text
Abstract:
Abstract The content of organic acids greatly influences the taste and storage life of fleshy fruit. Our current understanding of the molecular mechanism of organic acid accumulation in apple (Malus domestica) fruit focuses on the aluminum-activated malate transporter 9/Ma1 gene. In this study, we identified a candidate gene, MdWRKY126, for controlling fruit acidity independent of Ma1 using homozygous recessive mutants of Ma1, namely Belle de Boskoop “BSKP” and Aifeng “AF.” Analyses of transgenic apple calli and flesh and tomato (Solanum lycopersicum) fruit demonstrated that MdWRKY126 was substantially associated with malate content. MdWRKY126 was directly bound to the promoter of the cytoplasmic NAD-dependent malate dehydrogenase MdMDH5 and promoted its expression, thereby enhancing the malate content of apple fruit. In MdWRKY126 overexpressing calli, the mRNA levels of malate-associated transporters and proton pump genes also significantly increased, which contributed to the transport of malate accumulated in the cytoplasm to the vacuole. These findings demonstrated that MdWRKY126 regulates malate anabolism in the cytoplasm and coordinates the transport between cytoplasm and vacuole to regulate malate accumulation. Our study provides useful information to improve our understanding of the complex mechanism regulating apple fruit acidity.
APA, Harvard, Vancouver, ISO, and other styles
26

Thomas, C., P. Sirvent, S. Perrey, E. Raynaud, and J. Mercier. "Relationships between maximal muscle oxidative capacity and blood lactate removal after supramaximal exercise and fatigue indexes in humans." Journal of Applied Physiology 97, no. 6 (December 2004): 2132–38. http://dx.doi.org/10.1152/japplphysiol.00387.2004.

Full text
Abstract:
The present study investigated whether blood lactate removal after supramaximal exercise and fatigue indexes measured during continuous and intermittent supramaximal exercises are related to the maximal muscle oxidative capacity in humans with different training status. Lactate recovery curves were obtained after a 1-min all-out exercise. A biexponential time function was then used to determine the velocity constant of the slow phase (γ2), which denoted the blood lactate removal ability. Fatigue indexes were calculated during all-out (FIAO) and repeated 10-s cycling sprints (FISprint). Biopsies were taken from the vastus lateralis muscle, and maximal ADP-stimulated mitochondrial respiration ( Vmax) was evaluated in an oxygraph cell on saponin-permeabilized muscle fibers with pyruvate + malate and glutamate + malate as substrates. Significant relationships were found between γ2 and pyruvate + malate Vmax ( r = 0.60, P < 0.05), γ2 and glutamate + malate Vmax ( r = 0.66, P < 0.01), and γ2 and citrate synthase activity ( r = 0.76, P < 0.01). In addition, γ2, glutamate + malate Vmax, and pyruvate + malate Vmax were related to FIAO (γ2 − FIAO: r = 0.85; P < 0.01; glutamate + malate Vmax − FIAO: r = 0.70, P < 0.01; and pyruvate + malate Vmax − FIAO: r = 0.63, P < 0.01) and FISprint (γ2 − FISprint: r = 0.74, P < 0.01; glutamate + malate Vmax − FISprint: r = 0.64, P < 0.01; and pyruvate + malate Vmax − FISprint: r = 0.46, P < 0.01). In conclusion, these results suggested that the maximal muscle oxidative capacity was related to blood lactate removal ability after a 1-min all-out test. Moreover, maximal muscle oxidative capacity and blood lactate removal ability were associated with the delay in the fatigue observed during continuous and intermittent supramaximal exercises in well-trained subjects.
APA, Harvard, Vancouver, ISO, and other styles
27

Taillefer, M., T. Rydzak, D. B. Levin, I. J. Oresnik, and R. Sparling. "Reassessment of the Transhydrogenase/Malate Shunt Pathway in Clostridium thermocellum ATCC 27405 through Kinetic Characterization of Malic Enzyme and Malate Dehydrogenase." Applied and Environmental Microbiology 81, no. 7 (January 23, 2015): 2423–32. http://dx.doi.org/10.1128/aem.03360-14.

Full text
Abstract:
ABSTRACTClostridium thermocellumproduces ethanol as one of its major end products from direct fermentation of cellulosic biomass. Therefore, it is viewed as an attractive model for the production of biofuels via consolidated bioprocessing. However, a better understanding of the metabolic pathways, along with their putative regulation, could lead to improved strategies for increasing the production of ethanol. In the absence of an annotated pyruvate kinase in the genome, alternate means of generating pyruvate have been sought. Previous proteomic and transcriptomic work detected high levels of a malate dehydrogenase and malic enzyme, which may be used as part of a malate shunt for the generation of pyruvate from phosphoenolpyruvate. The purification and characterization of the malate dehydrogenase and malic enzyme are described in order to elucidate their putative roles in malate shunt and their potential role inC. thermocellummetabolism. The malate dehydrogenase catalyzed the reduction of oxaloacetate to malate utilizing NADH or NADPH with akcatof 45.8 s−1or 14.9 s−1, respectively, resulting in a 12-fold increase in catalytic efficiency when using NADH over NADPH. The malic enzyme displayed reversible malate decarboxylation activity with akcatof 520.8 s−1. The malic enzyme used NADP+as a cofactor along with NH4+and Mn2+as activators. Pyrophosphate was found to be a potent inhibitor of malic enzyme activity, with aKiof 0.036 mM. We propose a putative regulatory mechanism of the malate shunt by pyrophosphate and NH4+based on the characterization of the malate dehydrogenase and malic enzyme.
APA, Harvard, Vancouver, ISO, and other styles
28

Ryan, PR, E. Delhaize, and PJ Randall. "Malate Efflux From Root Apices and Tolerance to Aluminium Are Highly Correlated in Wheat." Functional Plant Biology 22, no. 4 (1995): 531. http://dx.doi.org/10.1071/pp9950531.

Full text
Abstract:
Aluminium (Al) can stimulate the efflux of malate and other organic acids from root apices of wheat (Triticum aestivum L.) seedlings. This response has been implicated in a mechanism of Al tolerance since the amount of malate released from an Al-tolerant genotype was 5-10-fold greater than the amount released from a near-isogenic, but Al sensitive, genotype. In the present study, 36 wheat cultivars were screened for Al tolerance and for the amount of malate released from their root apices with a standard A1 treatment. Excised root apices (3.0 mm) were used to measure malate efflux, and the relative tolerance to Al was determined from root growth measurements in 3 and 10μM AlCl3 with 200 μM CaCl2, pH 4.3. There was a significant correlation between relative tolerance of the genotypes to Al and the amount of malate released from their root apices. Growth measurements were also used to investigate the amelioration of Al toxicity by exogenous malate. In the presence of 3 μM Al alone, relative root growth of an Al-sensitive genotype was reduced to 13% of the control. Addition of 10 μM malate to the solution increased relative root growth to 50%, and 20 �M malate completely alleviated the Al-induced inhibition of root growth. The results support the hypothesis that the Al-stimulated efflux of malate from root apices is involved in a general mechanism for Al tolerance in wheat.
APA, Harvard, Vancouver, ISO, and other styles
29

Mizutani, Fumio, Soichi Yabuki, and Michihiko Asai. "L-Malate-sensing electrode based on malate dehydrogenase and NADH oxidase." Analytica Chimica Acta 245 (1991): 145–50. http://dx.doi.org/10.1016/s0003-2670(00)80214-7.

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

Bunel, Sergio, Carmen Ibarra, Victor Calvo, Andrei Blaskó, Clifford A. Bunton, and Nancy L. Keder. "Triammine cobalt(III)-L-malate. Malate ion as a tridentate ligand." Polyhedron 10, no. 20-21 (January 1991): 2495–500. http://dx.doi.org/10.1016/s0277-5387(00)86215-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Sonpavde, G., W. Jian, and S. P. Lerner. "Sunitinib malate is active and synergistic with cisplatin against human urothelial carcinoma in a preclinical model." Journal of Clinical Oncology 25, no. 18_suppl (June 20, 2007): 15632. http://dx.doi.org/10.1200/jco.2007.25.18_suppl.15632.

Full text
Abstract:
15632 Background: Sunitinib malate is an oral, multitargeted tyrosine kinase inhibitor of VEGFRs, PDGFRs, KIT, RET, and FLT3, approved multinationally for the treatment of advanced RCC and imatinib-resistant or -intolerant GIST. Angiogenesis and plasma VEGF correlate with poor outcomes in human urothelial carcinoma. We designed a preclinical study to examine the efficacy of sunitinib malate alone and in combination with cisplatin against human urothelial carcinoma in vitro and in a murine xenograft model. Methods: The IC50 for sunitinib malate and cisplatin was determined separately against two human urothelial carcinoma cell lines (TCC-SUP and 5637). Sunitinib malate and cisplatin were also applied concurrently to determine activity of the combination. Immunohistochemical staining was performed to detect expression of VEGFR2 on the cell lines, and to measure modulation of this pathway by sunitinib by measuring phosphorylated (p)VEGFR2. Anti-tumor activity of sunitinib malate alone and in combination with cisplatin was determined in a murine xenograft model bearing 5,637 cells. Results: Both human urothelial carcinoma cell lines were found to express VEGFR2. Sunitinib malate displayed significant activity against both urothelial carcinoma cell lines in vitro at low nanomolar concentrations. Furthermore, sunitinib malate in combination with cisplatin was synergistic in vitro. We observed primarily cytostatic activity for sunitinib malate at both 20 mg/kg and 40 mg/kg orally once daily against a murine xenograft model bearing subcutaneous 5,637 cell tumors during 4 weeks of treatment. Anti-tumor activity of sunitinib malate in combination with cisplatin and correlative studies are being evaluated in the murine xenograft model. Conclusion: Sunitinib malate has anti-tumor activity against human urothelial carcinoma as a single agent and is synergistic in combination with cisplatin in vitro. Sunitinib also has significant efficacy in a murine xenograft model of human urothelial carcinoma. These results warrant further exploration of sunitinib malate as a single agent and in combination with cisplatin chemotherapy in human urothelial carcinoma. No significant financial relationships to disclose.
APA, Harvard, Vancouver, ISO, and other styles
32

Negoro, Hiroaki, Atsushi Kotaka, and Hiroki Ishida. "Mutation in gene coding for glucose-induced degradation-deficient protein contributes to high malate production in yeast strain No. 28 and No. 77 used for industrial brewing of sake." Bioscience, Biotechnology, and Biochemistry 85, no. 5 (March 4, 2021): 1283–89. http://dx.doi.org/10.1093/bbb/zbab031.

Full text
Abstract:
ABSTRACT Saccharomyces cerevisiae produces organic acids including malate during alcohol fermentation. Since malate contributes to the pleasant flavor of sake, high-malate-producing yeast strain No. 28 and No. 77 have been developed by the Brewing Society of Japan. In this study, the genes responsible for the high malate phenotype in these strains were investigated. We had previously found that the deletion of components of the glucose-induced degradation-deficient (GID) complex led to high malate production in yeast. Upon examining GID protein–coding genes in yeast strain No. 28 and No. 77, a nonsense homozygous mutation of GID4 in strain No. 28 and of GID2 in strain No. 77 were identified as the cause of high malate production. Furthermore, complementary tests of these mutations indicated that the heterozygous nonsense mutation in GID2 was recessive. In contrast, the heterozygous nonsense mutation in GID4 was considered semidominant.
APA, Harvard, Vancouver, ISO, and other styles
33

Mortera, Pablo, Martín Espariz, Cristian Suárez, Guillermo Repizo, Josef Deutscher, Sergio Alarcón, Víctor Blancato, and Christian Magni. "Fine-Tuned Transcriptional Regulation of Malate Operons in Enterococcus faecalis." Applied and Environmental Microbiology 78, no. 6 (January 13, 2012): 1936–45. http://dx.doi.org/10.1128/aem.07280-11.

Full text
Abstract:
ABSTRACTInEnterococcus faecalis, themaelocus is constituted by two putative divergent operons,maePEandmaeKR. The first operon encodes a putative H+/malate symporter (MaeP) and a malic enzyme (MaeE) previously shown to be essential for malate utilization in this bacterium. ThemaeKRoperon encodes two putative proteins with significant similarity to two-component systems involved in sensing malate and activating its assimilation in bacteria. Our transcriptional and genetic assays showed thatmaePEandmaeKRare induced in response to malate by the response regulator MaeR. In addition, we observed that both operons were partially repressed in the presence of glucose. Accordingly, the cometabolism of this sugar and malate was detected. The binding of the complex formed by CcpA and its corepressor P-Ser-HPr to acresite located in themaeregion was demonstratedin vitroand explains the carbon catabolite repression (CCR) observed for themaePEoperon. However, our results also provide evidence for a CcpA-independent CCR mechanism regulating the expression of both operons. Finally, a biomass increment of 40 or 75% was observed compared to the biomass of cells grown only on glucose or malate, respectively. Cells cometabolizing both carbon sources exhibit a higher rate of glucose consumption and a lower rate of malate utilization. The growth improvement achieved byE. faecalisduring glucose-malate cometabolism might explain why this microorganism employs different regulatory systems to tightly control the assimilation of both carbon sources.
APA, Harvard, Vancouver, ISO, and other styles
34

Loke, Paxton, and Tiow-Suan Sim. "Molecular cloning, heterologous expression, and functional characterisation of a malate synthase gene fromStreptomyces coelicolorA3(2)." Canadian Journal of Microbiology 46, no. 8 (August 1, 2000): 764–69. http://dx.doi.org/10.1139/w00-044.

Full text
Abstract:
With the rapid generation of genetic information from the Streptomyces coelicolor genome project, deciphering the relevant gene products is critical for understanding the genetics of this model streptomycete. A putative malate synthase gene (aceB) from S. coelicolor A3(2) was identified by homology-based analysis, cloned by polymerase chain reaction, and fully sequenced on both strands. The putative malate synthase from S. coelicolor has an amino acid identity of 77% with the malate synthase of S. clavuligerus, and possesses an open reading frame which codes for a protein of 540 amino acids. In order to establish the identity of this gene, the putative aceB clones were subcloned into the expression vector pET24a, and heterologously expressed in Escherichia coli BL21(DE3). Soluble cell-free extracts containing the recombinant putative malate synthase exhibited a specific activity of 1623 (nmol·mg-1·min-1), which is an increment of 92-fold compared to the non-recombinant control. Thus, the gene product was confirmed to be a malate synthase. Interestingly, the specific activity of S. coelicolor malate synthase was found to be almost 8-fold higher than the specific activity of S. clavuligerus malate synthase under similar expression conditions. Furthermore, the genomic organisation of the three Streptomyces aceB genes cloned thus far is different from that of other bacterial malate synthases, and warrants further investigation.Key words: primary metabolism, polymerase chain reaction, glyoxylate pathway.
APA, Harvard, Vancouver, ISO, and other styles
35

Scholz, Thomas D., Stacia L. Koppenhafer, Cynthia J. Teneyck, and Brian C. Schutte. "Ontogeny of malate-aspartate shuttle capacity and gene expression in cardiac mitochondria." American Journal of Physiology-Cell Physiology 274, no. 3 (March 1, 1998): C780—C788. http://dx.doi.org/10.1152/ajpcell.1998.274.3.c780.

Full text
Abstract:
Developmental downregulation of the malate-aspartate shuttle has been observed in cardiac mitochondria. The goals of this study were to determine the time course of the postnatal decline and to identify potential regulatory sites by measuring steady-state myocardial mRNA and protein levels of the mitochondrial proteins involved in the shuttle. By use of isolated porcine cardiac mitochondria incubated with saturating concentrations of the cytosolic components of the malate-aspartate shuttle, shuttle capacity was found to decline by ∼50% during the first 5 wk of life (from 921 ± 48 to 531 ± 53 nmol ⋅ min−1 ⋅ mg protein−1). Mitochondrial aspartate aminotransferase mRNA levels were greater in adult than in newborn myocardium. mRNA levels of mitochondrial malate dehydrogenase in adult cardiac tissue were 224% of levels in newborn tissue, whereas protein levels were 54% greater in adult myocardium. Aspartate/glutamate carrier protein levels were also greater in adult than in newborn tissue. mRNA and protein levels of the oxoglutarate/malate carrier were increased in newborn myocardium. It was concluded that 1) myocardial malate-aspartate shuttle capacity declines rapidly after birth, 2) divergence of mitochondrial malate dehydrogenase mRNA and protein levels during development suggests posttranscriptional regulation of this protein, and 3) the developmental decline in malate-aspartate shuttle capacity is regulated by decreased oxoglutarate/malate carrier gene expression.
APA, Harvard, Vancouver, ISO, and other styles
36

Tayeh, M. A., and M. T. Madigan. "Malate dehydrogenases in phototrophic purple bacteria. Thermal stability, amino acid composition and immunological properties." Biochemical Journal 252, no. 2 (June 1, 1988): 595–600. http://dx.doi.org/10.1042/bj2520595.

Full text
Abstract:
Purified malate dehydrogenases from four species of non-sulphur purple phototrophic bacteria were examined for their heat-stability, amino acid composition and antigenic relationships. Malate dehydrogenase from Rhodospirillum rubrum, Rhodobacter capsulatus and Rhodomicrobium vannielii (which are all tetrameric proteins) had an unusually high glycine content, but the enzyme from Rhodocyclus purpureus (which is a dimer) did not. R. rubrum malate dehydrogenase was extremely heat-stable relative to the other enzymes, withstanding 65 degrees C for over 1 h with no loss of activity. By contrast, malate dehydrogenase from R. vannielii lost activity above 35 degrees C, and that from R. capsulatus above 40 degrees C. Amino acid compositional relatedness and immunological studies indicated that tetrameric phototrophic-bacterial malate dehydrogenases were highly related to one another, but only distantly related to the tetrameric enzyme from Bacillus. This suggests that, despite differences in their thermal properties, the tetrameric malate dehydrogenases of non-sulphur purple bacteria constitute a distinct biochemical class of this catalyst.
APA, Harvard, Vancouver, ISO, and other styles
37

Miller, A. D., K. Maghlaoui, G. Albanese, D. A. Kleinjan, and C. Smith. "Escherichia coli chaperonins cpn60 (groEL) and cpn10 (groES) do not catalyse the refolding of mitochondrial malate dehydrogenase." Biochemical Journal 291, no. 1 (April 1, 1993): 139–44. http://dx.doi.org/10.1042/bj2910139.

Full text
Abstract:
In vitro refolding of pig mitochondrial malate dehydrogenase is investigated in the presence and absence of Escherichia coli chaperonins cpn60 (groEL) and cpn10 (groES). The refolded yields of active malate dehydrogenase are increased almost 3-fold in the presence of groEL, groES, Mg2+/ATP and K+ ions. Chaperonin-assisted refolding of malate dehydrogenase does not have an absolute requirement for K+ ions but Mg2+/ATP is obligatory. When ATP is replaced by other nucleoside triphosphates, or by non-hydrolysable ATP analogues, assisted refolding is prevented. Optimal chaperonin-assisted refolding requires both groEL and groES homo-oligomers in molar excess over malate dehydrogenase. Kinetic analysis shows that the chaperonins do not catalyse the refolding of malate dehydrogenase but increase the flux of unfolded enzyme through the productive refolding pathway without altering and/or accelerating that pathway. Although not acting as refolding catalysts, the chaperonins are able to assist at least six consecutive cycles of malate dehydrogenase refolding.
APA, Harvard, Vancouver, ISO, and other styles
38

Jia, Dongjie, Peng Wu, Fei Shen, Wei Li, Xiaodong Zheng, Yongzhang Wang, Yongbing Yuan, Xinzhong Zhang, and Zhenhai Han. "Genetic variation in the promoter of an R2R3−MYB transcription factor determines fruit malate content in apple (Malus domestica Borkh.)." Plant Physiology 186, no. 1 (February 24, 2021): 549–68. http://dx.doi.org/10.1093/plphys/kiab098.

Full text
Abstract:
Abstract Deciphering the mechanism of malate accumulation in apple (Malus domestica Borkh.) fruits can help to improve their flavor quality and enhance their benefits for human health. Here, we analyzed malate content as a quantitative trait that is determined mainly by genetic effects. In a previous study, we identified an R2R3−MYB transcription factor named MdMYB44 that was a candidate gene in qtl08.1 (quantitative trait locus mapped to chromosome 8) of fruit malate content. In the present study, we established that MdMYB44 negatively regulates fruit malate accumulation by repressing the promoter activity of the malate-associated genes Ma1 (Al-Activated Malate Transporter 9), Ma10 (P-type ATPase 10), MdVHA-A3 (V-type ATPase A3), and MdVHA-D2 (V-type ATPase D2). Two single-nucleotide polymorphisms (SNPs) in the MdMYB44 promoter, SNP A/G and SNP T/−, were experimentally shown to associate with fruit malate content. The TATA-box in the MdMYB44 promoter in the presence of SNP A enhances the basal activity of the MdMYB44 promoter. The binding of a basic-helix–loop–helix transcription factor MdbHLH49 to the MdMYB44 promoter was enhanced by the presence of SNP T, leading to increased MdMYB44 transcript levels and reduced malate accumulation. Furthermore, MdbHLH49 interacts with MdMYB44 and enhances MdMYB44 activity. The two SNPs could be used in combination to select for sour or non-sour apples, providing a valuable tool for the selection of fruit acidity by the apple breeding industry. This research is important for understanding the complex molecular mechanisms of fruit malate accumulation and accelerating the development of germplasm innovation in apple species and cultivars.
APA, Harvard, Vancouver, ISO, and other styles
39

Hutchinson, J. P., T. S. el-Thaher, and A. D. Miller. "Refolding and recognition of mitochondrial malate dehydrogenase by Escherichia coli chaperonins cpn 60 (groEL) and cpn10 (groES)." Biochemical Journal 302, no. 2 (September 1, 1994): 405–10. http://dx.doi.org/10.1042/bj3020405.

Full text
Abstract:
In vitro refolding of pig mitochondrial malate dehydrogenase is investigated in the presence of Escherichia coli chaperonins cpn60 (groEL) and cpn10 (groES). When the enzyme is initially denatured with 3 M guanidinium chloride, chaperonin-assisted refolding is 100% efficient. C.d. spectroscopy reveals that malate dehydrogenase is almost unfolded in 3 M guanidinium chloride, suggesting that a state with little or no residual secondary structure is the optimal ‘substrate’ for chaperonin-assisted refolding. Malate dehydrogenase denatured to more highly structured states proves to refold less efficiently with chaperonin assistance. The enzyme is shown not to aggregate under the refolding conditions, so that losses in refolding efficiency result from irreversible misfolding. Evidence is advanced to suggest that the chaperonins are unable to rescue irreversibly misfolded malate dehydrogenase. A novel use is made of 100 K Centricon concentrators to study the binding of [14C]acetyl-labelled malate dehydrogenase to groEL by an ultrafiltration binding assay. Analysis of the data by Scatchard plot shows that acetyl-malate dehydrogenase, which has previously been extensively unfolded with guanidinium chloride, binds to groEL at a specific binding site(s). At saturation, one acetyl-malate dehydrogenase homodimer (two polypeptides) is shown to bind to each groEL homooligomer with a binding constant of approx. 10 nM.
APA, Harvard, Vancouver, ISO, and other styles
40

Abdossi, Vahid, Farshid Esmaeili, and Mohsen Kazemi. "Efficiency of essential oils and nano-malate in reduction of ethylene production and extension of vase life of cut Eustoma grandiflorum Mariachii. cv. Blue flowers." Bangladesh Journal of Botany 44, no. 3 (October 13, 2018): 465–68. http://dx.doi.org/10.3329/bjb.v44i3.38556.

Full text
Abstract:
The effect of essential oils and nano-malate in extending the vase-life of Eustoma grandiflorum Mariachii. cv. Blue flower was investigated. The treatment with 3 mM nano-malate increased flower longevity as compared to control. Nano-malate treatment increased chlorophyll, proline and carbohydrate content and membrane stability, while descreasing ACO (ACC-oxidase activity) and MDA (malondialdehyde) content and delay of senescence and peroxidation of lipids. Thyme oil was slightly effective significantly. The application of nano-malate as preservative solutions for E. grandiflorum flowers maintained the vase life of flowers for a longer period.
APA, Harvard, Vancouver, ISO, and other styles
41

Yasin, Yamin, Maszlin Mohamad, Azlin Sanusi, and Ahmad Saat. "Optimization of Lead Removal from Aqueous Solution Using Intercalated Malate Mg-Al Layered Double Hydroxides." Advanced Materials Research 832 (November 2013): 665–69. http://dx.doi.org/10.4028/www.scientific.net/amr.832.665.

Full text
Abstract:
Conventional optimization method was conducted for the removal of lead from aqueous solution by used of intercalated malate layered double hydroxides. The effects of various parameters such as contact time, amount of adsorbent dosage and the pH of the lead solution are studied. The extent of lead ions removal increased with the increased in contact time and amount of malate-Mg-Al used however, the percentage removal was decreased with the increased in concentration and pH. The fundamentals of lead removal from aqueous solution by used of malate-Mg-Al can be explained by the formation of complexes between the malate and Pb2+ ions. The results from this study indicated that layered double hydroxide intercalated with malate could be used as potential adsorbent for the removal of lead ions from aqueous solution.
APA, Harvard, Vancouver, ISO, and other styles
42

Ding, Pingtao, Hailong Guo, and Jonathan D. G. Jones. "Deadlier than the malate." Cell Research 28, no. 6 (May 29, 2018): 609–10. http://dx.doi.org/10.1038/s41422-018-0042-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Bobyleva‐Guarriero, V., D. Battelli, M. Bellei, and H. A. Lardy. "Sources of intramitochondrial malate." FASEB Journal 3, no. 10 (August 1989): 2208–11. http://dx.doi.org/10.1096/fasebj.3.10.2568962.

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

Abdul Rahman, Mohd Basyaruddin, Khairulazhar Jumbri, Kamaliah Sirat, Reza Kia, and Hoong-Kun Fun. "TetraethylammoniumL-malate 1.36-hydrate." Acta Crystallographica Section E Structure Reports Online 65, no. 1 (December 10, 2008): o49—o50. http://dx.doi.org/10.1107/s1600536808040348.

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

Taillandier, P., J. P. Riba, and P. Strehaiano. "Malate utilization bySchizosaccharomyces pombe." Biotechnology Letters 10, no. 7 (July 1988): 469–72. http://dx.doi.org/10.1007/bf01027058.

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

Shi, Miaomiao, Yue Jing, Liuzhi Yang, Xianqing Huang, Hongwei Wang, Yizhe Yan, and Yanqi Liu. "Structure and Physicochemical Properties of Malate Starches from Corn, Potato, and Wrinkled Pea Starches." Polymers 11, no. 9 (September 19, 2019): 1523. http://dx.doi.org/10.3390/polym11091523.

Full text
Abstract:
In this study, corn, potato, and wrinkled pea starches were esterified with malic acid under high temperatures for different lengths of time. The degree of substitution (DS), granule morphology, crystal structure, gelatinization properties, and the digestibility of the malate starch were investigated. Fourier transform infrared spectroscopy (FT–IR) suggested that the malate starch showed a new infrared absorption peak near 1747 cm−1, indicating the occurrence of an esterification reaction. With an increasing treatment time, the degree of substitution (DS) of the malate starch displayed an increasing trend. Scanning electron microscopy (SEM) demonstrated a significant change in the surface structure of the starch granules. X-ray diffractometry (XRD) reflected that the crystal structure of the malate starches was destroyed. The thermogravimetric (TG) curves showed that the maximum heat loss rate of the malate starch was ahead of that of native starch, which caused the decreased degree of crystallinity. These properties of malate starch could allow it to be used for the purpose of starch modification to produce resistant starch and to provide new applications for starch.
APA, Harvard, Vancouver, ISO, and other styles
47

Cerana, R., L. Giromini, and R. Colombo. "Malate-Regulated Channels Permeable to Anions in Vacuoles of Arabidopsis thaliana." Functional Plant Biology 22, no. 1 (1995): 115. http://dx.doi.org/10.1071/pp9950115.

Full text
Abstract:
Anion channels in isolated vacuoles of Arabidopsis thaliana cultured cells were studied by means of the patch clamp technique in the whole-vacuole configuration. In symmetrical 100 mM KCl, a high resistance of the membrane at positive potentials inside the vacuole was observed. In symmetrical 100 mM K2-malate positive potentials inside the vacuole elicited slowly developing inward currents, due to the opening of channels, which, according to measurements of reversal potential, are selective for malate. The activation potential of the channels shifted as a function of the cytoplasmic malate concentration, but it was always such that the channels opened only to mediate malate influx into the vacuole. The channels were also permeable to succinate, fumarate and, to a lesser extent, oxaloacetate. In vacuoles preincubated with cytoplas- mic malate, inward currents were also elicited in the presence of KCl or KNO3 at the cytoplasmic side of the tonoplast. Malate channels were different from the cation slow vacuolar-type channels with regard to their sensitivity to changes in the cytoplasmic concentrations of Ca2+ and ATP, and in temperature between 10 and 20�C.
APA, Harvard, Vancouver, ISO, and other styles
48

Lerondel, Guillaume, Thierry Doan, Nicola Zamboni, Uwe Sauer, and Stéphane Aymerich. "YtsJ Has the Major Physiological Role of the Four Paralogous Malic Enzyme Isoforms in Bacillus subtilis." Journal of Bacteriology 188, no. 13 (July 1, 2006): 4727–36. http://dx.doi.org/10.1128/jb.00167-06.

Full text
Abstract:
ABSTRACT The Bacillus subtilis genome contains several sets of paralogs. An extreme case is the four putative malic enzyme genes maeA, malS, ytsJ, and mleA. maeA was demonstrated to encode malic enzyme activity, to be inducible by malate, but also to be dispensable for growth on malate. We report systematic experiments to test whether these four genes ensure backup or cover different functions. Analysis of single- and multiple-mutant strains demonstrated that ytsJ has a major physiological role in malate utilization for which none of the other three genes could compensate. In contrast, maeA, malS, and mleA had distinct roles in malate utilization for which they could compensate one another. The four proteins exhibited malic enzyme activity; MalS, MleA, and MaeA exhibited 4- to 90-fold higher activities with NAD+ than with NADP+. YtsJ activity, in contrast, was 70-fold higher with NADP+ than with NAD+, with Km values of 0.055 and 2.8 mM, respectively. lacZ fusions revealed strong transcription of ytsJ, twofold higher in malate than in glucose medium, but weak transcription of malS and mleA. In contrast, mleA was strongly transcribed in complex medium. Metabolic flux analysis confirmed the major role of YtsJ in malate-to-pyruvate interconversion. While overexpression of the NADP-dependent Escherichia coli malic enzyme MaeB did not suppress the growth defect of a ytsJ mutant on malate, overexpression of the transhydrogenase UdhA from E. coli partially suppressed it. These results suggest an additional physiological role of YtsJ beyond that of malate-to-pyruvate conversion.
APA, Harvard, Vancouver, ISO, and other styles
49

Heart, Emma, Gary W. Cline, Leon P. Collis, Rebecca L. Pongratz, Joshua P. Gray, and Peter J. S. Smith. "Role for malic enzyme, pyruvate carboxylation, and mitochondrial malate import in glucose-stimulated insulin secretion." American Journal of Physiology-Endocrinology and Metabolism 296, no. 6 (June 2009): E1354—E1362. http://dx.doi.org/10.1152/ajpendo.90836.2008.

Full text
Abstract:
Pyruvate cycling has been implicated in glucose-stimulated insulin secretion (GSIS) from pancreatic β-cells. The operation of some pyruvate cycling pathways is proposed to necessitate malate export from the mitochondria and NADP+-dependent decarboxylation of malate to pyruvate by cytosolic malic enzyme (ME1). Evidence in favor of and against a role of ME1 in GSIS has been presented by others using small interfering RNA-mediated suppression of ME1. ME1 was also proposed to account for methyl succinate-stimulated insulin secretion (MSSIS), which has been hypothesized to occur via succinate entry into the mitochondria in exchange for malate and subsequent malate conversion to pyruvate. In contrast to rat, mouse β-cells lack ME1 activity, which was suggested to explain their lack of MSSIS. However, this hypothesis was not tested. In this report, we demonstrate that although adenoviral-mediated overexpression of ME1 greatly augments GSIS in rat insulinoma INS-1 832/13 cells, it does not restore MSSIS, nor does it significantly affect GSIS in mouse islets. The increase in GSIS following ME1 overexpression in INS-1 832/13 cells did not alter the ATP-to-ADP ratio but was accompanied by increases in malate and citrate levels. Increased malate and citrate levels were also observed after INS-1 832/13 cells were treated with the malate-permeable analog dimethyl malate. These data suggest that although ME1 overexpression augments anaplerosis and GSIS in INS-1 832/13 cells, it is not likely involved in MSSIS and GSIS in pancreatic islets.
APA, Harvard, Vancouver, ISO, and other styles
50

Poolman, B., D. Molenaar, E. J. Smid, T. Ubbink, T. Abee, P. P. Renault, and W. N. Konings. "Malolactic fermentation: electrogenic malate uptake and malate/lactate antiport generate metabolic energy." Journal of Bacteriology 173, no. 19 (1991): 6030–37. http://dx.doi.org/10.1128/jb.173.19.6030-6037.1991.

Full text
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