Journal articles on the topic 'Tartaric acid'

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

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 'Tartaric acid.'

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

YAMAMOTO, Yoshikazu. "L(+)-Tartaric acid. d-Tartaric acid." Journal of Synthetic Organic Chemistry, Japan 48, no. 1 (1990): 71–72. http://dx.doi.org/10.5059/yukigoseikyokaishi.48.71.

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

Spiller, Gene A., Jon A. Story, Emily J. Furumoto, Jo Carol Chezem, and Monica Spiller. "Effect of tartaric acid and dietary fibre from sun-dried raisins on colonic function and on bile acid and volatile fatty acid excretion in healthy adults." British Journal of Nutrition 90, no. 4 (October 2003): 803–7. http://dx.doi.org/10.1079/bjn2003966.

Full text
Abstract:
Sun-dried raisins are a source of dietary fibre and tartaric acid. The effects of tartaric acid on colon function have not been the focus of extensive research. The purpose of the present study was to evaluate the effects of dietary fibre and tartaric acid from sun-dried raisins on colon function and on faecal bile acid and short-chain fatty acid (SCFA) excretion in healthy adults. Thirteen healthy subjects were fed 120 g sun-dried raisins/d or 5 g cream of tartar (equivalent to the tartaric acid in 120 g sun-dried raisins)/d for 9 weeks, divided into 3-week cycles. The experimental diets were fed in a crossover design after an initial control period. Faeces were collected for the last 4 d of each cycle for analysis of SCFA and bile acids. Intestinal transit time decreased from 42h on the baseline diet to 31h on cream of tartar (P<0·1) and to 28h on sun-dried raisins (P<0·05). Faeces were softer on both sun-dried raisins and cream of tartar, but sun-dried raisins increased faecal wet weight (P<0·05), while cream of tartar did not. Sun-dried raisins caused significant reductions from baseline values in total bile acid concentration (from 1·42 (sd 1·03) to 1·09 (sd 0·76) mg/g, P<0·05), whereas cream of tartar did not (1·40 (sd 1·06) mg/g). Sun-dried raisins also significantly reduced the lithocholic (LC):deoxylithocholic acid (DC) ratio (from 1·63 (sd 0·85) to 1·09 (sd 0·50), P<0·02), whereas cream of tartar reduced the ratio, but to a lesser extent (1·29 (sd 0·79), NS). Both faecal bile acids and the LC:DC ratio are indicators of reduced risk for colon cancer. Sun-dried raisins increased total SCFA excretion (from 5·6 (sd 3·4) to 7·6 (sd 3·0) g/4d, P<0·05), which remained unchanged with cream of tartar (5·6 (sd 3·0) g/4d). Both sun-dried raisins and cream of tartar appear to be good stool softeners and to shorten intestinal transit time, although the fibre in sun-dried raisins has the added benefit of increasing faecal weight. Both sun-dried raisins and cream of tartar modulate the composition of faecal bile acids and SCFA in a way that has potential health benefits.
APA, Harvard, Vancouver, ISO, and other styles
3

Fronczek, Frank R., Richard D. Gandour, Thomas M. Fyles, Philippa J. Hocking, Susan J. McDermid, and P. Daniel Wotton. "Polycarboxylate crown ethers from meso-tartaric acid." Canadian Journal of Chemistry 69, no. 1 (January 1, 1991): 12–19. http://dx.doi.org/10.1139/v91-003.

Full text
Abstract:
The synthesis of crown ethers derived from meso-tartaric acid was investigated. The sodium salt of the bis(dimethylamide) of meso-tartaric acid reacted with diethylene glycol ditosylate to give a mixture of 18-crown-6 tetraamide and 27-crown-9 hexaamide crown ethers. The 2R,3S,11S,12R 18-crown-6 isomer crystallized in triclinic space group [Formula: see text] (a = 7.557(2), b = 8.866(2), c = 10.4133(13) Å, α = 94.13(2), β = 95.86(2), γ = 99.26(2)°, R = 0.040 for 2090 observed of 3129 unique reflections). The structures of the remaining products were then assigned from the NMR spectra. The solution conformations of the amide crown ethers were examined by NMR, and provide a rationale for the product distribution obtained. One of the 18-crown-6 isomers and a mixture of the two 27-crown-9 isomers were hydrolyzed to the respective crown ether carboxylic acids, and the stability constants for complexation of cations were determined by potentiometric titration. The meso tetra- and hexacarboxylates are remarkably nonselective and inefficient cation complexing agents, compared to related crown ethers from R,R-(+)-tartaric acid, due to the unfavorable conformational control exerted by the tartaro units. Key words: crown ether synthesis, complexation, crown ether conformation, meso-tartaric acid, crystal structure.
APA, Harvard, Vancouver, ISO, and other styles
4

Losev, Evgeniy, and Elena Boldyreva. "The effect of amino acid backbone length on molecular packing: crystalline tartrates of glycine, β-alanine, γ-aminobutyric acid (GABA) and DL-α-aminobutyric acid (AABA)." Acta Crystallographica Section C Structural Chemistry 74, no. 2 (January 18, 2018): 177–85. http://dx.doi.org/10.1107/s2053229617017909.

Full text
Abstract:
We report a novel 1:1 cocrystal of β-alanine with DL-tartaric acid, C3H7NO2·C4H6O6, (II), and three new molecular salts of DL-tartaric acid with β-alanine {3-azaniumylpropanoic acid–3-azaniumylpropanoate DL-tartaric acid–DL-tartrate, [H(C3H7NO2)2]+·[H(C4H5O6)2]−, (III)}, γ-aminobutyric acid [3-carboxypropanaminium DL-tartrate, C4H10NO2 +·C4H5O6 −, (IV)] and DL-α-aminobutyric acid {DL-2-azaniumylbutanoic acid–DL-2-azaniumylbutanoate DL-tartaric acid–DL-tartrate, [H(C4H9NO2)2]+·[H(C4H5O6)2]−, (V)}. The crystal structures of binary crystals of DL-tartaric acid with glycine, (I), β-alanine, (II) and (III), GABA, (IV), and DL-AABA, (V), have similar molecular packing and crystallographic motifs. The shortest amino acid (i.e. glycine) forms a cocrystal, (I), with DL-tartaric acid, whereas the larger amino acids form molecular salts, viz. (IV) and (V). β-Alanine is the only amino acid capable of forming both a cocrystal [i.e. (II)] and a molecular salt [i.e. (III)] with DL-tartaric acid. The cocrystals of glycine and β-alanine with DL-tartaric acid, i.e. (I) and (II), respectively, contain chains of amino acid zwitterions, similar to the structure of pure glycine. In the structures of the molecular salts of amino acids, the amino acid cations form isolated dimers [of β-alanine in (III), GABA in (IV) and DL-AABA in (V)], which are linked by strong O—H...O hydrogen bonds. Moreover, the three crystal structures comprise different types of dimeric cations, i.e. (A...A)+ in (III) and (V), and A +...A + in (IV). Molecular salts (IV) and (V) are the first examples of molecular salts of GABA and DL-AABA that contain dimers of amino acid cations. The geometry of each investigated amino acid (except DL-AABA) correlates with the melting point of its mixed crystal.
APA, Harvard, Vancouver, ISO, and other styles
5

Hasbullah, Umar Hafidz Asy'ari, Miftahul Wahidatun Ni’mah, Endang Is Retnowati, and Rini Umiyati. "Physical, Chemical, and Sensory Properties of Robusta Coffee Effervescent Tablets Formulated in Various Organic Acids." Pelita Perkebunan (a Coffee and Cocoa Research Journal) 38, no. 1 (April 20, 2022): 54–69. http://dx.doi.org/10.22302/iccri.jur.pelitaperkebunan.v38i1.489.

Full text
Abstract:
Coffee effervescent products are an innovation in coffee formulation. The compounds that play a role in effervescent are acids and bases. Type of organic acid give an impact on the effervescent characteristics. This study aimed to examine the effect of type of organic acid on physical, chemical, and sensory properties ofRobusta coffee effervescent tablets. This study used a completely randomized design with three acids in the formulation, namely citric acid, tartaric acid, and malic acid. Samples were analyzed in three replications. Making effervescent tablets was done by compression technique in a mixture of all ingredients according tothe formula. The results showed that different acid had a significant effect on physical and chemical parameters. Malic acid caused a faster effervescent time than citric acid and tartaric acid. Malic acid and tartaric acid tended to lower the pH slightly than citric acid. Malic acid and citric acid tended to produce hardertablets than tartaric acid. However, tartaric acid slightly increased tablets’ brightness (L*) compared to malic acid and citric acid. Tartaric acid and malic acid tended to reduce moisture compared to citric acid. The IC50 value of effervescent with malic acid and tartaric acid was lower than that of citric acid. However, therewas a slight decrease in total phenol in both. Meanwhile, the sensory profiles of tablets and effervescent drinks did not differ due to different acids. The recommended formula was that the effervescent using malic acid had an effervescent time of 166 seconds, hardness 321 N, moisture 8%, IC50 5.5 mg mL-1, total phenol4.2 mg gallic acid equivalent (GAE) g-1, and a drink profile that has the best color, aroma, taste, and runs time
APA, Harvard, Vancouver, ISO, and other styles
6

Fukami, Takanori, Shuta Tahara, Chitoshi Yasuda, and Keiko Nakasone. "Structural Refinements and Thermal Properties of L(+)-Tartaric, D(–)-Tartaric, and Monohydrate Racemic Tartaric Acid." International Journal of Chemistry 8, no. 2 (March 10, 2016): 9. http://dx.doi.org/10.5539/ijc.v8n2p9.

Full text
Abstract:
<p>Differential scanning calorimetry, thermogravimetric-differential thermal analysis, and X-ray diffraction measurements were performed on single crystals of L(+)-tartaric, D(–)-tartaric, and monohydrate racemic (MDL-) tartaric acid. The exact crystal structures of the three acids, including the positions of all hydrogen atoms, were determined at room temperature. It was pointed out that one of O–H–O hydrogen bonds in MDL-tartaric acid has an asymmetric double-minimum potential well along the coordinate of proton motion. The weight losses due to thermal decomposition of L- and D-tartaric acid were observed to occur at 443.0 and 443.2 K, respectively, and at 306.1 and 480.6 K for MDL-tartaric acid. The weight losses for L- and D-tartaric acid during decomposition were probably caused by the evolution of 3H<sub>2</sub>O and 3CO gases. By considering proton transfer between two possible sites in the hydrogen bond, we concluded that the weight losses at 306.1 and 480.6 K for MDL-tartaric acid were caused by the evaporation of half the bound water molecules in the sample, and by the evaporation of the remaining water molecules and the evolution of 3H<sub>2</sub>O and 3CO gases, respectively.</p>
APA, Harvard, Vancouver, ISO, and other styles
7

Synoradzki, Ludwik, Pawel Ruśkowski, and Urszula Bernaś. "TARTARIC ACID AND ITSO-ACYL DERIVATIVES. PART 1. SYNTHESIS OF TARTARIC ACID ANDO-ACYL TARTARIC ACIDS AND ANHYDRIDES." Organic Preparations and Procedures International 37, no. 1 (February 2005): 37–63. http://dx.doi.org/10.1080/00304940509355401.

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

Junge, Jonas Yde, Anne Sjoerup Bertelsen, Line Ahm Mielby, Yan Zeng, Yuan-Xia Sun, Derek Victor Byrne, and Ulla Kidmose. "Taste Interactions between Sweetness of Sucrose and Sourness of Citric and Tartaric Acid among Chinese and Danish Consumers." Foods 9, no. 10 (October 9, 2020): 1425. http://dx.doi.org/10.3390/foods9101425.

Full text
Abstract:
Tastes interact in almost every consumed food or beverage, yet many aspects of interactions, such as sweet-sour interactions, are not well understood. This study investigated the interaction between sweetness from sucrose and sourness from citric and tartaric acid, respectively. A cross-cultural consumer study was conducted in China (n = 120) and Denmark (n = 139), respectively. Participants evaluated six aqueous samples with no addition (control), sucrose, citric acid, tartaric acid, or a mixture of sucrose and citric acid or sucrose and tartaric acid. No significant difference was found between citric acid and tartaric acid in the suppression of sweetness intensity ratings of sucrose. Further, sucrose suppressed sourness intensity ratings of citric acid and tartaric acid similarly. Culture did not impact the suppression of sweetness intensity ratings of citric or tartaric acid, whereas it did influence sourness intensity ratings. While the Danish consumers showed similar suppression of sourness by both acids, the Chinese consumers were more susceptible towards the sourness suppression caused by sucrose in the tartaric acid-sucrose mixture compared to the citric acid-sucrose mixture. Agglomerative hierarchical cluster analysis revealed clusters of consumers with significant differences in sweetness intensity ratings and sourness intensity ratings. These results indicate that individual differences in taste perception might affect perception of sweet-sour taste interactions, at least in aqueous solutions.
APA, Harvard, Vancouver, ISO, and other styles
9

Luner, Paul E., Aditya D. Patel, and Dale C. Swenson. "(\pm)-Tartaric acid." Acta Crystallographica Section C Crystal Structure Communications 58, no. 6 (May 21, 2002): o333—o335. http://dx.doi.org/10.1107/s0108270102006650.

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

Jung, Yui Jung. "The Influence of Organic Acid on Color Retention after Dyeing - Focusing on succinic acid and tartaric acid." Journal of Health and Beauty 16, no. 2 (August 31, 2022): 163–72. http://dx.doi.org/10.35131/ishb.2022.16.2.163.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Robert, L., J. Mourgues, Arlette Pamar-Robert, D. Achour, and J. Molinier. "Adsorption of tartaric acid and malic by active carbons." OENO One 29, no. 1 (March 31, 1995): 49. http://dx.doi.org/10.20870/oeno-one.1995.29.1.1719.

Full text
Abstract:
<p style="text-align: justify;">Adsorption of tartaric acid and malic acid by active carbons bas been tested with six samples of carbons. At acid pH ; the adsorbed amounts of tartaric acid and of malic acid are practically the same. For a solution concentration of 20 g/l, adsorbed amounts from 0.008 to 0.29 gramme for one gramme of carbon have been found, variation which may be due to various states of carbon surface oxidation. Increasing the pH of the solutions shows a dramatic decrease of adsorbed amounts, this decrease being more rapid for tartaric acid than for malic acid. At neutral pH, the adsoiption becomes negligible. With acidic solutions containing the two acids altogether, acid malic is more adsorbed than tartaric acid.</p>
APA, Harvard, Vancouver, ISO, and other styles
12

Fonseca, Alvaro. "Utilization of tartaric acid and related compounds by yeasts: taxonomic implications." Canadian Journal of Microbiology 38, no. 12 (December 1, 1992): 1242–51. http://dx.doi.org/10.1139/m92-205.

Full text
Abstract:
A survey of yeasts capable of growing on L(+)-tartaric acid as the sole source of carbon and energy showed that this organic acid is assimilated by a significant number of species of basidiomycetous affinity and is seldom utilized by ascomycetous yeasts. This conclusion was further supported by the fact that among approximately 100 isolates from various natural substrates, using selective media with L(+)-tartaric acid, only one strain of ascomycetous affinity was obtained. In a more comprehensive survey 442 yeast strains belonging to 138 species, mostly of basidiomycetous affinity, were also screened for the assimilation of different aldaric acids: D(−)-tartaric acid, meso-tartaric acid, L(−)-malic acid, D(+)-glucaric acid (saccharic acid), and galactaric acid (mucic acid). L(+)-Tartrate was the most frequently utilized tartaric acid isomer (55% of the total number of strains of basidiomycetous affinity belonging to either the Tremellales/Filobasidiales or the Ustilaginales) when compared with the D(−) and meso forms, which were assimilated by 12 and 18% of the total number of strains, respectively (mainly of tremellaceous species). Saccharic acid was utilized by about 75% of the total number of species of Tremellales affinity and by less than 20% of the ustilaginaceous species. Assimilation of mucic acid occurred in more than 50% of the tremellaceous species and only in 5% of the species related to the Ustilaginales. These tests, not used in standard yeast identification sets, appear to contribute to distinguishing taxa at or above the species level. Key words: assimilation, tartaric acid, aldaric acids, yeasts, taxonomy.
APA, Harvard, Vancouver, ISO, and other styles
13

Cheng, Guo Ling, Xue Bai, and Qun Hui Wang. "Effects of Natural Organic Acids on Growth of Maize and Uptake of Copper and Lead by Maize in Contaminated Soil." Advanced Materials Research 322 (August 2011): 21–24. http://dx.doi.org/10.4028/www.scientific.net/amr.322.21.

Full text
Abstract:
Seven kinds of natural organic acids were added to moderate artificial contaminated soil to investigate the effects of natural organic acids on growth of maize seedlings and phytoextraction of copper and lead by maize. The results show that the effects of organic acids on plant growth are different in Cu and Pb contaminated soil. The natural organic acids can change the dry matter distribution of the shoot and the root, oxalic acid and tartaric acid can increase the root biomass in different degrees. Oxalic acid and tartaric acid can significantly increase the concentrations and uptake of Cu and Pb in the shoots of maize, indicate that oxalic acid and tartaric acid are potential phytoextraction intensifiers in phytoremediation of Cu and Pb contaminated soil.
APA, Harvard, Vancouver, ISO, and other styles
14

Li, Menghan, Jing Su, Huanqi Yang, Lei Feng, Minghui Wang, Gezhe Xu, Jianhui Shao, and Chunhua Ma. "Grape Tartaric Acid: Chemistry, Function, Metabolism, and Regulation." Horticulturae 9, no. 11 (October 26, 2023): 1173. http://dx.doi.org/10.3390/horticulturae9111173.

Full text
Abstract:
Tartaric acid (TA) is the primary organic acid present in grapes and a fundamental constituent of wine, responsible for shaping its taste, aroma, and overall quality. This review presents a comprehensive overview of the advances made in previous investigations on grape tartaric acid. It elucidates the structural properties, distribution characteristics, biosynthesis, catabolism, and transcriptional regulation of grape tartaric acid, and also speculates on the regulatory mechanism of tartaric acid based on the modulation of ascorbic acid-related transcription factors. Furthermore, this review provides insights into the future research directions and objectives, with the goal of providing a reference for the analysis of the complete biosynthetic pathway of grape tartaric acid, thereby enabling precise regulation of tartaric acid.
APA, Harvard, Vancouver, ISO, and other styles
15

Etim, Effiong Ukorebi. "Enhancement of tartaric acid modified washing solutions for lead decontamination of tropical soils." Ovidius University Annals of Chemistry 31, no. 1 (January 1, 2020): 27–32. http://dx.doi.org/10.2478/auoc-2020-0006.

Full text
Abstract:
AbstractTartaric acid is generally not an effective soil washing solution, hence this study focuses on enhancing its usage for soil-Pb decontamination. Three tropical soil types (sandy, clay and loamy) with different lead concentrations were subjected to single batch washing using 0.01, 0.1, 0.5 and 1 M tartaric acid with 5% and 10% KCl modification at 3% soil-pulp-density for 2, 6, 12 and 24 h washing time. The optimum washing conditions were 1 M tartaric acid at 24 h washing time, with Pb removal efficiency: sandy- 94.3%, clay-67.6% and loamy-36.8%. Modification of tartaric acid with 5% and 10% KCl brought about some degree of enhancement of Pb removal efficiency especially for clay and loamy soils. Removal efficiency for 5% KCl modification were: sandy-97.9%, clay-96.2% with 1 M tartaric acid at 24 h washing time, loamy-76.7% for 0.5 M tartaric acid. Similarly, 10% KCl modification were: sandy-96.7%, clay-97.2% for 1 M tartaric acid at 24 h, loamy-82.1% for 0.5 M tartaric acid. Removal efficiency was soil concentration dependent. Generally, removal efficiency increased with increasing tartaric acid concentrations and washing time. Tartaric acid washing is promising and recommended in events of moderate contamination and 10% KCl modification in event of high level contamination. Further study is needed on enhancing very low concentrations of tartaric acid for large scale applications.
APA, Harvard, Vancouver, ISO, and other styles
16

Fuleki, Tibor, Estela Pelayo, and Rodrigo B. Palabay. "Carboxylic Acid Composition of Authentic Varietal and Commercial Grape Juices." Journal of AOAC INTERNATIONAL 76, no. 3 (May 1, 1993): 591–600. http://dx.doi.org/10.1093/jaoac/76.3.591.

Full text
Abstract:
Abstract Quantities were determined of tartaric, malic, citric, lactic, succinic, and shikimic acid present in juices produced by cold and hot pressing from 56 grape cultivars grown in Ontario for the 1988 and 1989 seasons. The individual acids were separated by liquid chromatography (LC) and detected at 210 or 250 nm. Lactic and succinic acids were determined enzymatically. The ranges of acid concentrations found were as follows: tartaric, 4.19-13.51 g/L; malic, 1.68-15.36 g/L; citric, 0.305-1.158 g/L; lactic, 0.015-0.388 g/L; succinic, 0.002-0.075 g/L; and shikimic, 0-0.102 g/L. Tartaric and malic acids were the major acids in every cultivar. The tartaric:malic acid ratio ranged from 0.52 to 4.36, but it was &gt;1.0 in most cases. Genotype had significant effect only on the shikimic acid content. Hot pressing of red cultivars yielded juice with significantly higher pH and tartaric, citric, lactic, and total acid contents than cold pressing. Vintage had no significant effect on acid composition. The total acid content determined by LC was always considerably higher than the titratable acidity, but good correlation existed between the 2 measurements. Most commercial grape juice had similar composition to that of authentic juices. However, tartaric acid content was lower in the majority of commercial juices because of losses during the detartration process. Also, indications of acidulation existed in some of the commercial juices.
APA, Harvard, Vancouver, ISO, and other styles
17

Zhou, Hui, Wei Yang, Miao Liu, Shi Liang Li, and Qian Qian Li. "Effects of EDTA and Organic Acids on Cd Desorption from Zhangshi Irrigation Area Soil." Advanced Materials Research 356-360 (October 2011): 1566–69. http://dx.doi.org/10.4028/www.scientific.net/amr.356-360.1566.

Full text
Abstract:
The effects of the added EDTA and organic acids (oxalic acid, tartaric acid and acetic acid) on Cd desorption of Zhangshi Irrigation Area (ZIR)contaminated soil of Shenyang city was investigated by batch balance experiments, in which the concentrations of acids, pH and temperature were examined. The results showed that EDTA, oxalic acid, tartaric acid and acetic acid modified the desorption behaviors of Cd. And the desorption level was EDTA>tartaric acid >oxalic acid >acetic acid. Also, the desorption amount of Cd increased with the concentration ranges from 5 to 40mmol/L; the desorbed amount obviously reduced with elevating pH when the pH was below 6; the adsorption was facilitated and the desorption was weakened when the pH was above 6. In addition, the Cd desorption amount raised with the temperature increasing.
APA, Harvard, Vancouver, ISO, and other styles
18

Qin, Rongxiu, Haiyan Chen, Rusi Wen, Guiqing Li, and Zhonglei Meng. "Effect of Boric Acid on the Ionization Equilibrium of α-Hydroxy Carboxylic Acids and the Study of Its Applications." Molecules 28, no. 12 (June 12, 2023): 4723. http://dx.doi.org/10.3390/molecules28124723.

Full text
Abstract:
To investigate the synergistic catalytic effects of boric acid and α-hydroxycarboxylic acids (HCAs), we analyzed and measured the effects of the complexation reactions between boric acid and HCAs on the ionization equilibrium of the HCAs. Eight HCAs, glycolic acid, D-(−)-lactic acid, (R)-(−)-mandelic acid, D-gluconic acid, L-(−)-malic acid, L-(+)-tartaric acid, D-(−)-tartaric acid, and citric acid, were selected to measure the pH changes in aqueous HCA solutions after adding boric acid. The results showed that the pH values of the aqueous HCA solutions gradually decreased with an increase in the boric acid molar ratio, and the acidity coefficients when boric acid formed double-ligand complexes with HCAs were smaller than those of the single-ligand complexes. The more hydroxyl groups the HCA contained, the more types of complexes could be formed, and the greater the rate of change in the pH. The total rates of change in the pH of the HCA solutions were in the following order: citric acid > L-(−)-tartaric acid = D-(−)-tartaric acid > D-gluconic acid > (R)-(−)-mandelic acid > L-(−)-malic acid > D-(−)-lactic acid > glycolic acid. The composite catalyst of boric acid and tartaric acid had a high catalytic activity—the yield of methyl palmitate was 98%. After the reaction, the catalyst and methanol could be separated by standing stratification.
APA, Harvard, Vancouver, ISO, and other styles
19

Rajkovic, Milos, Ivana Novakovic, and Aleksandar Petrovic. "Determination of titratable acidity in white wine." Journal of Agricultural Sciences, Belgrade 52, no. 2 (2007): 169–84. http://dx.doi.org/10.2298/jas0702169r.

Full text
Abstract:
The amount of titration acid in must is in the largest number of cases with in the range 5.0-8.0 g/dm3. Wines, as a rule, contain less acids than must, and according to Regulations, titratable acidity is in the range of 4.0-8.0 g/dm3 expressed in tartaric acid, because a part of tartaric acid is deposited in the form of salts (tartar or argol) during alcohol fermentation. For wines that contain less than 4 g/dm3 of titratable acids there arises a suspicion about their origin, that is, that during the preparation some illegal acts were done. Because of that, the aim of this paper is to determine titratable acidity in white wine, using standard methods of determination, which are compared with the results received by potentiometric titration using ion-selective electrode. According to the received results it can be seen that wine titration with indicator gives sufficient reliable values of wine titration acidity. However, as potentiometric titration at pH value 7.00 is more reliable and objective method, the values of titratable acids content in wine, expressed through tartaric acid, are given according to this result. The analysis of differential potentiometric curves shows that these curves can give us an answer to the question of the presence of a larger amount of other nonorganic substances, which have already existed in wine. However, none of the used methods gives absolutely reliable answer what substances are present in analysed samples.
APA, Harvard, Vancouver, ISO, and other styles
20

Vicol, Crina, and Gheorghe Duca. "Synergistic, Additive and Antagonistic Interactions of Some Phenolic Compounds and Organic Acids Found in Grapes." Acta Chimica Slovenica 70, no. 4 (December 20, 2023): 588–600. http://dx.doi.org/10.17344/acsi.2023.8214.

Full text
Abstract:
The antioxidant interactions between several natural phenolic and non-phenolic compounds (catechin, quercetin, rutin, resveratrol, gallic acid and ascorbic acid) and organic acids (tartaric, citric and dihydroxyfumaric acids) were studied using the DPPH method. Main additive and antagonistic interactions have been found for the combinations of catechin, quercetin, resveratrol and gallic acid with tartaric and citric acids; such behavoir can be due to the enhanced stability of the phenolic compounds in acidic media. Rutin and ascorbic acid showed good synergistic effects with tartaric and citric organic acids, which could be due to the polymerization processes in the case of rutin and the change in the mechanism of action in the case of ascorbic acid. In combination with dihydroxyfumaric acid, the mixtures showed dose–dependent synergistic, additive, or antagonistic antioxidant interactions. Good synergistic effects were observed for the binary mixtures of dihydroxyfumaric acid with ascorbic acid, catechin, and rutin.
APA, Harvard, Vancouver, ISO, and other styles
21

Resitca, Vladislav, Anatol Balanuta, Iurie Scutaru, Ecaterina Covaci, Aliona Sclifos, Antoanela Patras, and Ana-Maria Borta. "POSSIBILITY AND NECESSITY OF TARTARIC ACID PRODUCTION IN THE REPUBLIC OF MOLDOVA." Journal of Engineering Science 29, no. 1 (March 2022): 151–63. http://dx.doi.org/10.52326/jes.utm.2022.29(1).14.

Full text
Abstract:
The wine industry has been and remains a source of natural- tartaric acid. The tartaric acid can be obtained from such wastes as grape marcs, yeast, vinasse and wine stone. But the use of these wastes was limited in the Republic of Moldova by the production of tartaric acid lime (calcium tartrate) and wine stone, which were shipped to Ukraine and Armenia where the finished product is obtained. Currently, tartaric acid is used in considerable quantities in the winemaking and food industry, being a quite expensive imported product. The Department of Oenology and Chemistry has developed a complete technological scheme for the use of wine wastes to obtain the finished product – tartaric acid. The realization of the proposed tartaric acid production in the Republic of Moldova is important for the country’s economy and it does not require large investments. Wineries can also help to organize tartaric acid production by providing calcium tartrate, wine stone, pressed or dried yeast, and other ingredients.
APA, Harvard, Vancouver, ISO, and other styles
22

Kalmykova, N. N., E. N. Kalmykova, and T. V. Gaponova. "CHARACTERISTIC OF ORGANIC ACIDS COMPOSITION OF MUSTS AND WINES FROM RED GRAPEVINE VARIETIES OF INTERSPECIFIC ORIGIN." Russian Vine 20 (September 2022): 59–64. http://dx.doi.org/10.32904/2712-8245-2022-20-59-64.

Full text
Abstract:
The aim of the work was to study the composi-tion of organic acids of musts and wines made from red grapevine varieties of interspecific origin growing in the conditions of Rostov re-gion. The objects of the study were must and wines prepared under conditions of micro winemaking from red grapevine varieties of interspecific origin bred by ARRIV&W, a branch of FSBSI FRASC ‒ Denisovsky, Cab-ernet Severniy, Augusta, Cabernet Sauvignon (control), according to the classical technology adopted for red dry wines. As a result of the study the following results were obtained. In must of all studied varieties the predominance of tartaric acid over malic acid was observed in several times, while in the interspecific grape-vine varieties the concentration of tartaric acid was 5‒6 times higher than malic acid, while in the control only 2.5 times higher. The greatest amount of tartaric acid was observed in the must from Cabernet Severniy grapes ‒ 5.6 g/dm3. Malic acid accumulation was signifi-cantly lower in all variants (0.48‒1.1 g/dm3) compared with the control Cabernet Sauvignon (2 g/dm3). There was a slight accumulation of succinic acid in the must from Denisovsky and Cabernet Severniy grapes. In all variants, with the exception of Augusta wine, the concentra-tion of tartaric and malic acids decreased. The total proportion of tartaric and malic acids in the experimental wines was 38‒74.7 % of all acids in wine. The highest concentration of succinic acid was observed in wines from Den-isovsky and Cabernet Sauvignon varieties.
APA, Harvard, Vancouver, ISO, and other styles
23

Dong, Hao, Qing Liu, Yuanyu Tian, and Yingyun Qiao. "Tartaric Acid–Zinc Nitrate as an Efficient Brønsted Acid-Assisted Lewis Acid Catalyst for the Mannich Reaction." Journal of Chemical Research 42, no. 9 (September 2018): 463–66. http://dx.doi.org/10.3184/174751918x15355426661373.

Full text
Abstract:
Tartaric acid–zinc nitrate has been found to be an efficient Brønsted acid-assisted Lewis acid catalytic system for the facile synthesis of β-amino carbonyl compounds through the one-pot Mannich reaction of aldehydes, aromatic amines and ketones in ethanol at room temperature. Remarkable enhancement of reactivity by tartaric acid (Brønsted acid) was observed in these reactions in the presence of anhydrous zinc nitrate (Lewis acid), due to coordination of the tartaric acid ligand to zinc ions increasing the acidity of the system. This procedure shows some advantages such as mild reaction conditions, short reaction times and high yields.
APA, Harvard, Vancouver, ISO, and other styles
24

Bhanot, Lalitha, Anuj Kumar, Diwakar Shende, and Kailas Wasewar. "Extraction of the Food Additive Tartaric Acid Using Octanol, Methyl Isobutyl Ketone, Kerosene, Mustard Oil, And Groundnut Oil." Hungarian Journal of Industry and Chemistry 51, no. 2 (November 8, 2023): 15–20. http://dx.doi.org/10.33927/hjic-2023-13.

Full text
Abstract:
Tartaric acid (TA) is a dicarboxylic acid found in bananas, grapes, apples, papaya, cherries, pineapple, pears, mangoes, and tamarind. Due to its widespread use in the food, cosmetic and pharmaceutical industries, it is an essential carboxylic acid. Tartaric acid is produced commercially from wine-industry byproducts and is also present in the industry's effluent. Separating tartaric acid from wastewater is challenging. In this research, tartaric acid was separated from the aqueous phase using chemical and organic solvents such as groundnut oil, mustard oil, kerosene, octanol, and methyl isobutyl ketone (MIBK). Experiments were conducted at 298 K to determine the extraction efficiency (E%) and distribution coefficient (KD). The maximum extraction efficiencies of tartaric acid were found to be 49.01, 25.62, 16.73, 15.89 and 14.29% when using MIBK, octanol, kerosene, mustard oil and groundnut oil, respectively. The results demonstrate the significance of solvent choice in the extraction of tartaric acid with solvents such as MIBK and octanol being more effective at extracting TA from aqueous solutions. On the other hand, the sustainability of the method for separating tartaric acid was highlighted when organic solvents were applied.
APA, Harvard, Vancouver, ISO, and other styles
25

Emmett, Michael. "Recurrent tartaric acid acidosis?" Kidney International 79, no. 2 (January 2011): 258–59. http://dx.doi.org/10.1038/ki.2010.443.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Mohammed, Abdul-Halim A.-K., and Safa'a Yaseen. "Preparation of Tartaric Acid." Iraqi Journal of Chemical and Petroleum Engineering 1, no. 1 (December 31, 2000): 50–55. http://dx.doi.org/10.31699/ijcpe.2000.1.9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Sato, Kei, Fumikazu Ikemori, Sathiyamurthi Ramasamy, Akihiro Fushimi, Kimiyo Kumagai, Akihiro Iijima, and Yu Morino. "Four- and Five-Carbon Dicarboxylic Acids Present in Secondary Organic Aerosol Produced from Anthropogenic and Biogenic Volatile Organic Compounds." Atmosphere 12, no. 12 (December 20, 2021): 1703. http://dx.doi.org/10.3390/atmos12121703.

Full text
Abstract:
To better understand precursors of dicarboxylic acids in ambient secondary organic aerosol (SOA), we studied C4–C9 dicarboxylic acids present in SOA formed from the oxidation of toluene, naphthalene, α-pinene, and isoprene. C4–C9 dicarboxylic acids present in SOA were analyzed by offline derivatization gas chromatography–mass spectrometry. We revealed that C4 dicarboxylic acids including succinic acid, maleic acid, fumaric acid, malic acid, DL-tartaric acid, and meso-tartaric acid are produced by the photooxidation of toluene. Since meso-tartaric acid barely occurs in nature, it is a potential aerosol tracer of photochemical reaction products. In SOA particles from toluene, we also detected a compound and its isomer with similar mass spectra to methyltartaric acid standard; the compound and the isomer are tentatively identified as 2,3-dihydroxypentanedioic acid isomers. The ratio of detected C4–C5 dicarboxylic acids to total toluene SOA mass had no significant dependence on the initial VOC/NOx condition. Trace levels of maleic acid and fumaric acid were detected during the photooxidation of naphthalene. Malic acid was produced from the oxidation of α-pinene and isoprene. A trace amount of succinic acid was detected in the SOA produced from the oxidation of isoprene.
APA, Harvard, Vancouver, ISO, and other styles
28

Jančářová, Irena, Luděk Jančář, Alice Náplavová, and Vlastimil Kubáň. "Changes of organic acids and phenolic compounds contents in grapevine berries during their ripening." Open Chemistry 11, no. 10 (October 1, 2013): 1575–82. http://dx.doi.org/10.2478/s11532-013-0288-2.

Full text
Abstract:
AbstractChanges of total content of phenolic substances, alteration in total titratable acidity and differences in tartaric acid content in grapes of four white (Müller-Thurgau — MT, Pinot Blanc — Rulandské bílé in Czech, RB, Sauvignon (Sg), and Muscat Ottonel — Muškát Ottonel in Czech, MO) and two blue (Dornfelder — Df and Blue Frankish — Frankovka in Czech, Fr) grapevine varieties throughout their growth, ripening and maturing (July–November). Potentiometric titration was applied for the determination of total titratable acids in grapes (expressed as tartaric acid equivalents in g L−1). A spectrophotometric method according Rebelein based on the formation of a colored complex of ammonium metavanadate and tartaric acid was used for determination of tartaric acid in green juice made by pressing unripe grapes. A spectrophotometric method based on reduction of phosphomolybdato-tungsten complex in alkaline solution using Folin-Ciocalteau reagent was applied for determination of total content of phenolic substances (TCP).
APA, Harvard, Vancouver, ISO, and other styles
29

Putranto, W. A., Y. Umardhani, Sulistyo, Yurianto, and A. P. Bayuseno. "Analysis of calcium carbonate polymorphs deposited in water piping system and the effect of tartaric acid additive." MATEC Web of Conferences 159 (2018): 01054. http://dx.doi.org/10.1051/matecconf/201815901054.

Full text
Abstract:
Calcium carbonate (CaCO3) formed in a water piping system was investigated in the presence of chemical additives tartaric acid (0.00 and 10.00 ppm) and various temperatures ((27 and 50ºC). The flow rate inside pipe (35 ml/min) were selected. Solutions of CaCl2 and Na2CO3 were prepared in water with equimolar to Ca2+ concentration of 3000 ppm. The induction time of scale nucleation varied from 24 min to 44 min. An increasing temperature of the solution resulted in more CaCO3 scale, mass, while the higher tartaric acid made the reduced mass of scales by 90%. SEM/EDS analysis verified CaCO3 with a plate like morphology. Also the XRPD Rietveld method provided the confirmation of a major phase of calcite and vaterite minerals followed by the minor of aragonite in the absence of tartaric acid and at room temperature. However, the presence of 10 ppm tartaric acid yielded the increasing result of calcite, while aragonite precipitation was hampered under the influence of tartaric acid. The presence of tartaric acid at increasing temperature of 50ºC could delay formation of calcite, whereas aragonite could be formed significantly. Thus the study showed the capacity of the tartaric acid in influencing CaCO3 crystallization.
APA, Harvard, Vancouver, ISO, and other styles
30

Godinho, Oswaldo E. S., Nilson E. Desouza, Luiz M. Aleixo, and Ari U. Ivaska. "Determination of Tartaric Acid and the Sum of Malic and Citric Acids in Grape Juices by Potentiometric Titration." Journal of AOAC INTERNATIONAL 71, no. 5 (September 1, 1988): 1028–32. http://dx.doi.org/10.1093/jaoac/71.5.1028.

Full text
Abstract:
Abstract Application of a linear algebraic method to the potentiometric titration of a mixture of tartaric and malic acids makes it possible to determine the individual concentrations of both acids in the same sample. These 2 acids have also been determined in grape juice free of citric acid after their separation from the juice matrix by precipitation as barium salts, followed by selective solubilization. It is also possible to determine tartaric acid and the sum of malic acid and citric acid in grape juice when the latter is present.
APA, Harvard, Vancouver, ISO, and other styles
31

Taleb Dhiab, Ammar, and Younus Abbas Khalaf Al-Saadi. "Effect of Adding Tartaric and Salicylic Acids and Their Mixture to Water and Diet in Egg Quality Characteristics of Aged Laying Hens." Diyala Agricultural Sciences Journal 13, no. 2 (December 30, 2021): 38–51. http://dx.doi.org/10.52951/dasj.21130206.

Full text
Abstract:
The study aimed to know the effect of adding tartaric and salicylic acids alone and a mixture of water and diet on the egg quality characteristics of aged laying hens for brown Lohmann. 210 laying hens 60 weeks old were used The duration of the experiment, which lasted 112 days(16 weeks), was divided into four equal periods at a rate of 28 days for each period,, distributed in equal numbers to 21 ground hens, 10 laying hens for each hen, and distributed to seven treatments with three replications of the treatment, T1 was fed a standard diet without adding (control), T2 was fed a standard diet supplemented with 0.2% tartaric acid with water, T3 was fed a standard diet supplemented with 0.2% salicylic acid with water, T4 was fed a standard diet supplemented with 0.4% tartaric acid and salicylic acid was added with water, T5 was fed a standard diet supplemented with % 0.2 tartaric acid in the diet, T6 was fed a standard diet supplemented with 0.2% salicylic acid in the diet, T7 was fed a standard diet supplemented with 0.4% a mixture of tartaric and salicylic acid in the diet. The results of adding the two acids in the water and feed showed that there was a significant improvement (P≤0.01) in the quality characteristics of the eggs produced, as it significantly improved the shell weight, shell thickness, albumin height, albumin weight, Haugh unit, rate of yolk height, yolk weight, and the yolk diameter was significantly reduced for the coefficients of The addition compared with the control treatment.
APA, Harvard, Vancouver, ISO, and other styles
32

Marchitan, Natalia. "Reactive Extraction of Tartaric Acid." Chemistry Journal of Moldova 4, no. 2 (December 2009): 28–33. http://dx.doi.org/10.19261/cjm.2009.04(2).15.

Full text
Abstract:
The present paper describes the results of reactive extraction of tartaric acid in model systems, which can be used for its separation from secondary wine products. As extractant have been used a normal/isododecyl mixed secondary amine Amberlite LA-2. The following parameters of the separation process have been varied: nature of diluent and modifier; modifier concentration; concentration, temperature and pH of the tartaric acid solution and the stirring time, and the work intervals have been established. It was concluded that in determinated conditions the extent of tartaric acid extraction attains value 85-95%.
APA, Harvard, Vancouver, ISO, and other styles
33

LEIROSE, G. D., M.-F. GRENIER-LOUS TALOT, and A. H. OLIVEIRA. "L-TARTARIC ACID: PRODUCTION TECHNOLOGY, ECONOMIC GROWTH AND QUALITY CONTROL." Periódico Tchê Química 15, no. 30 (August 20, 2018): 12–18. http://dx.doi.org/10.52571/ptq.v15.n30.2018.15_periodico30_pgs_12_18.pdf.

Full text
Abstract:
Natural substances are the basis of many types of industries and represent a growing market. The study of these products and the development of analytical methods should accompany this growth to ensure quality and provenance to consumers. An example to be discussed is the L(+)-Tartaric acid, an organic compound of molecular formula C4H6O6. This organic acid is widely applied in wine, food and pharmaceutical industry. It is obtained naturally through the fermentation of fruits, especially grape and tamarind. Synthetically, there are two ways of obtaining L(+)-tartaric acid on industrial scale. It can be synthesized by the reaction of maleic anhydride with hydrogen peroxide, which is derived from petroleum by-products. And by biotechnological synthesis, in which cis-epoxy succinic acid, also derived from petroleum, is converted into L(+)-tartaric acid by hydrolase enzyme. The market for tartaric acid is growing and is considered promising. Currently, there is a lack of legislation and specific rules that allow classification of tartaric acid according to its origin. This legal vacuum precludes quality assurance for the consumer. This lack of safety is a matter of great concern as applications of tartaric acid come directly to final consumer.
APA, Harvard, Vancouver, ISO, and other styles
34

Synoradzki, Ludwik, Urszula Bernaś, and Pawel Ruśkowski. "TARTARIC ACID AND ITSO-ACYL DERIVATIVES. PART 2. APPLICATION OF TARTARIC ACID AND OFO-ACYL TARTARIC ACIDS AND ANHYDRIDES. RESOLUTION OF RACEMATES." Organic Preparations and Procedures International 40, no. 2 (April 2008): 163–200. http://dx.doi.org/10.1080/00304940809458084.

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

Herlina, Nita Kuswardhani, Maria Belgis, and Adinda Tiara. "Characterization of Physical and Chemical Properties of Effervescent Tablets Temulawak (Curcuma zanthorrhiza) in the Various Proportion of Sodium Bicarbonate and Tartaric Acid." E3S Web of Conferences 142 (2020): 03006. http://dx.doi.org/10.1051/e3sconf/202014203006.

Full text
Abstract:
The purpose of this study was to determine the effect of comparative treatment of the proportion of Sodium bicarbonate and tartaric acid on the physical and chemical properties of effervescent tablets temulawak, as well as knowing the proportions of sodium bicarbonate and tartaric acid right to produce a good effervescent tablets temulawak. The research method uses a single completely randomized design namely the ratio of sodium bicarbonate: tartaric acid, (F1 = 2.0: 2.5; F2 = 2.5: 2.0; F3 = 3.0; 1.3; F4 = 3 , 5: 1,0, and F5 = 4.0: 0.5y). each treatment was repeated 3 (three) times. The resulted data were analyzed by ANOVA test. The results showed that the proportion of sodium bicarbonate and tartaric acid significantly affected color lightness, hardness, hygroscopicity, solubility times, water content, ash content, and not significantly affect viscosity and pH. The right proportion of sodium bicarbonate acid tartaric acid for making effervescent tablets temulawak is A1 treatment (proportion of sodium bicarbonate and tartaric acid 2.0: 2.5) with the attributes of water content of 66.72%, hardness of 2.20 kg, hygroscopicity of 25.43 g, solubility times of 35 seconds, viscosity 1.75 MPa.S, water content 1.19%, ash content 3.10%, and pH = 5.1.
APA, Harvard, Vancouver, ISO, and other styles
36

Xuan, Jinsong, and Yingang Feng. "Enantiomeric Tartaric Acid Production Using cis-Epoxysuccinate Hydrolase: History and Perspectives." Molecules 24, no. 5 (March 5, 2019): 903. http://dx.doi.org/10.3390/molecules24050903.

Full text
Abstract:
Tartaric acid is an important chiral chemical building block with broad industrial and scientific applications. The enantioselective synthesis of l(+)- and d(−)-tartaric acids has been successfully achieved using bacteria presenting cis-epoxysuccinate hydrolase (CESH) activity, while the catalytic mechanisms of CESHs were not elucidated clearly until very recently. As biocatalysts, CESHs are unique epoxide hydrolases because their substrate is a small, mirror-symmetric, highly hydrophilic molecule, and their products show very high enantiomeric purity with nearly 100% enantiomeric excess. In this paper, we review over forty years of the history, process and mechanism studies of CESHs as well as our perspective on the future research and applications of CESH in enantiomeric tartaric acid production.
APA, Harvard, Vancouver, ISO, and other styles
37

EL-SHENAWY, MOUSTAFA A., and ELMER H. MARTH. "Organic Acids Enhance the Antilisterial Activity of Potassium Sorbate." Journal of Food Protection 54, no. 8 (August 1, 1991): 593–97. http://dx.doi.org/10.4315/0362-028x-54.8.593.

Full text
Abstract:
Tryptose broth containing 0.0, 0.05, 0.15, or 0.3% potassium sorbate was acidified to pH 5.0 or 5.6 with acetic, tartaric, lactic or citric acid; inoculated to contain ca. 103 CFU Listeria monocytogenes/ml; and incubated at 13 or 35°C. The pathogen was inactivated in tryptose broth containing (a) 0.3% sorbate and acidified to pH 5.0 with acetic, tartaric, lactic, or citric acid although the time required for inactivation varied from ca. 30 h to &gt; 10 d and (b) 0.15% sorbate and acidified to pH 5.0 with tartaric acid. Growth of the pathogen was inhibited to various degrees by other combinations of sorbate and organic acids. L. monocytogenes grew at pH 5.6 regardless of organic acid or incubation temperature used and at pH 5.0 in all instances except when acetic acid and incubation at 13°C were used.
APA, Harvard, Vancouver, ISO, and other styles
38

Gonzalez, Susana V., and Per Carlsen. "Investigation of tartaric acid amide formation by thermolysis of tartaric acids with alkylamines." Arkivoc 2011, no. 9 (June 25, 2011): 325–36. http://dx.doi.org/10.3998/ark.5550190.0012.924.

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

NAGADEEP, J., P. KAMARAJ, M. ARTHANAREESWARI, and P. A. VIVEKANAND. "Identification of Tartaric Acid Adduct Impurities in Dipyridamole Capsule Formulation Related Substances Method." Asian Journal of Chemistry 33, no. 2 (2021): 307–13. http://dx.doi.org/10.14233/ajchem.2021.22984.

Full text
Abstract:
Two tartaric acid adduct degradation products namely dipyridamole tartaric acid monoester and dipyridamole ditartaric acid ester are observed in a dipyridamole capsule formulation. The adduct impurities are inevitable in the formulation due to the interaction of multilayers of dipyridamole on tartaric acid pellets. Present study reported a simple procedure for generating these two major adducts degradation products from a mixture of dipyridamole drug substance and tartaric acid by stress study. The obtained stress mixture was characterized by liquid chromatography-tandem mass spectrometry (LC-MS) to assure the identity of adduct degradant impurities. The obtained solid stress mixture was stable for more than one year and the prepared solution can be used as reference solution to identify both the degradants during related substance analysis. Practically, the identification of tartaric acid degradants applied to the British pharmacopeia monograph related substances method, where no mechanism for identification of these adduct impurities was described. This study establishes relative retention times for the British pharmacopeia method, which enables the chemist to monitor these two major degradants during quality control release testing and shelf life stability. The same kind of experimental approach for identifying tartaric acid adduct impurities in the British pharmacopeia method can be extended to any of the in-house laboratory-developed related substance methods.
APA, Harvard, Vancouver, ISO, and other styles
40

Kousar, Mubeen, Umme Salma, Taous Khan, and Abdul Jabbar Shah. "Antihypertensive Potential of Tartaric Acid and Exploration of Underlying Mechanistic Pathways." Dose-Response 20, no. 4 (October 2022): 155932582211357. http://dx.doi.org/10.1177/15593258221135728.

Full text
Abstract:
Tartaric acid is capable of balancing blood pressure. It is the main constituent of antihypertensive agents (grapes and wine) and has not been scientifically explored as an antihypertensive remedy. This study aimed to investigate the antihypertensive effect of a low-dose tartaric acid in vivo and explore underlying mechanisms in vitro. Intravenous administration of tartaric acid at the dose of 50 µg/kg caused a % fall in mean arterial pressure (MAP) in normotensive and hypertensive rats [51.5 ± 1.7 and 63.5 ± 2.9% mmHg]. This hypotensive effect was partially inhibited by atropine (1 mg/kg) and L-NAME (100 µg/kg) pretreatment. In hypertensive rats, oral administration of tartaric acid (.1, .5, 1, 5, and 10 mg/kg) for 2 weeks resulted in 65 ± 7.3 mmHg MAP at 10 mg/kg. This antihypertensive effect was comparable to the orally administered verapamil (10 mg/kg) for 2 weeks which caused a decrease in MAP 60.4 ± 3.8 mmHg. Tartaric acid relaxed phenylephrine (PE) and High K+-induced contractions with EC50 values of .157 (.043-.2) and 1.93 (.07-2) µg/mL in vitro. This endothelium-dependent relaxation was inhibited with atropine (1 µM) and L-NAME (10 µM) pretreatment. Tartaric acid also suppressed phenylephrine contractions in Ca+2 free/EGTA medium and on voltage-dependent calcium channels, causing the concentration–response curves toward right. Tartaric acid induced negative inotropic and chronotropic effects with EC50 values of .26 (.14-.4) and .60 (.2-.8) in rat atria. It showed its effect by complete blockade against atropine and partially in propranolol pretreatment. These findings provide scientific basis to low-dose tartaric acid as an antihypertensive and vasodilatory remedy through muscarinic receptor-linked nitric oxide (NO) pathway and Ca+2 channel antagonist.
APA, Harvard, Vancouver, ISO, and other styles
41

Tudorascu, Marius, Spiridon Oprea, Afrodita Doina Marculescu, and Stefania Tudorascu. "Enzymatic Iodination of Maleic and Fumaric Acids Diethyl Esters." Revista de Chimie 59, no. 12 (January 9, 2009): 1400–1404. http://dx.doi.org/10.37358/rc.08.12.2071.

Full text
Abstract:
The mechanism of the enzymatic iodination process of diethylmaleate and diethylfumarate (which present no miscibility with water) in the presence of lactoperoxidase, both in diluted hydrogen peroxide solution and in a generating system of hydrogen peroxide using ammonium and calcium iodides as halide sources in disperse system (after an ultrasonic pretreatment) was studied. The obtained sole product (diethyl-2, 3-diiodosuccinate) after the enzymatic iodination process was directly hydrolyzed to a tartaric acid present in an optically inactive form. The mechanism of obtaining the intermediate and final products and respectively, the existence of both D, L-tartaric acid and meso-tartaric acids (as lithium bitartrates) were also investigated.
APA, Harvard, Vancouver, ISO, and other styles
42

Reis, L. G., A. V. Chaves, S. R. O. Williams, and P. J. Moate. "Comparison of enantiomers of organic acids for their effects on methane production in vitro." Animal Production Science 54, no. 9 (2014): 1345. http://dx.doi.org/10.1071/an14199.

Full text
Abstract:
This study aimed to evaluate the effect of organic acids on in vitro fermentation characteristics. Four organic acids (tartaric, malic, fumaric and citric) and their enantiomers (L-tartaric, D-tartaric, DL-tartaric, L-malic and DL-malic) were analysed using in vitro batch culture incubations, at four concentrations (0, 5, 10 and 15 mM). Cumulative total gas and methane (CH4) production (mL/g DM) were measured at 6, 12 and 24 h; ammonia, pH, volatile fatty acids (VFA) and in vitro dry matter digestibility (IVDMD) were determined after 24 h of fermentation. Overall, addition of acids at 5 to 15 mM increased (P < 0.0001) cumulative gas and CH4 production. No effect (P > 0.10) of enantiomers, individual acid or interaction acid × concentration was detected at 12 and 24 h for cumulative gas or CH4 production. Addition of DL-malic, L-malic and fumaric acids increased (P < 0.0001) the percentage of propionic acid in the ruminal fluid total VFA compared with all concentrations of the other organic acids or their enantiomers. Ammonia concentration was not affected (P ≥ 0.28) by the addition of organic acids, concentrations or interactions. These findings are evidence that ruminal microorganisms can metabolise both D- and L-enantiomers of organic acids. None of the organic acids and their enantiomers at four different concentrations demonstrated potential as CH4 mitigation agents.
APA, Harvard, Vancouver, ISO, and other styles
43

Mukhamedyarova, Lilia I., Sergey G. Bezryadin, Elena Yu Klukvina, Vladimir V. Chevela, and Valentina Yu Ivanova. "Composition, stability and stereo effects of zirconium(IV) dl-tartrate formation." Butlerov Communications 57, no. 2 (February 28, 2019): 28–34. http://dx.doi.org/10.37952/roi-jbc-01/19-57-2-28.

Full text
Abstract:
The system of zirconium (IV) – dl-tartaric acid for metal: ligand 1: 1, 1: 2 and 1: 3 ratios in aqueous solution has been studied by means of using potentiometric titration method in combination with mathematical modeling. The comparison of Bjerrum functions from pH for zirconium(IV) systems: d-tartaric acid and zirconium (IV): dl-tartaric acid, has revealed the following features in the behavior of the curves: the degree of titration for the complexes at a fixed pH value for systems with dl-tartaric acid is more than for d-acid. The CPESSP software complex has calculated the composition, stability constants and molar fractions of zirconium(IV) tartrate accumulation. It has been also found that at a ratio of 1: 1 for Zr (IV) and ligand (H4Tart) ions in the system under study ZrHTart+ is formed, which is tetramerized into Zr4Tart40 and, further, tetranuclear particles of varying degrees of deprotonization are formed, as well as mononuclear forms. In a strongly alkaline pH environment > 10, Bjerrum curves for d- and dl-tartaric acids overlap each other and correspond to hydroxocomplexes of varying degrees of titration. For the 1: 2 ratio, the composition of the complexes for the zirconium(IV) – dl-H4T system is slightly different; compared to the zirconium(IV) – dH4T system, differences are clearly observed for both low and high concentrations. Based on these data, a complex formation scheme in the Zr(IV) – dl-tartaric acid system has been proposed for all the ratios studied. The characteristics of stereoselective diastereomer formation have been calculated. It has been revealed that in the medium of racemic tartrate, ddd- and lll-Zr(H2Tart)2(HTart)3-forms, as well as Zr(H2Tart)(НTart)24-Zr(HTart)35- are formed on a stereoselective basis.
APA, Harvard, Vancouver, ISO, and other styles
44

Wagalgave, Sopan M., Mahmood D. Aljabri, Keerti Bhamidipati, Deepak A. Shejule, Dinesh N. Nadimetla, Mohammad Al Kobaisi, Nagaprasad Puvvada, Sidhanath V. Bhosale, and Sheshanath V. Bhosale. "Characteristics of the pH-regulated aggregation-induced enhanced emission (AIEE) and nanostructure orchestrate via self-assembly of naphthalenediimide–tartaric acid bola-amphiphile: role in cellular uptake." New Journal of Chemistry 45, no. 19 (2021): 8775–85. http://dx.doi.org/10.1039/d0nj05845a.

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

Suhesti, Tuti Sri, Warsinah Warsinah, and Pitra Wulandari. "Optimizing effervescent granules of butterfly pea (Clitoria ternatea L) flower ethanol extract as antioxidant." Acta Pharmaciae Indonesia : Acta Pharm Indo 10, no. 1 (July 27, 2022): 5803. http://dx.doi.org/10.20884/1.api.2022.10.1.5803.

Full text
Abstract:
Background: Butterfly pea (Clitoria ternatea) contains secondary metabolites, including flavonoids, saponins, terpenoids, tannins, and anthocyanins which have antioxidant activity. Objective: This research aims to produce the effervescent granule preparations of the butterfly pea flower ethanol extract with the optimal concentrations of citric acid and tartaric acid. Methods: Butterfly pea flower was extracted using 70% ethanol. Effervescent granules were made using the wet granulation method in eight formulas containing citric acid and tartaric acid. The physical properties of granules were evaluated, including extract quality, flow rate, dissolution time, and pH. Results: The concentration of the mixed components of citric acid 48.65 mg and tartaric acid 576.30 mg was the most optimal combination of acid sources for effervescent granules of butterfly pea flower extract with a desirability value of 1,000. Conclusion: The variation in the concentration of citric acid and tartaric acid affected flow rate, dissolution time, and pH of the effervescent granule preparation.
APA, Harvard, Vancouver, ISO, and other styles
46

Gonta, Maria, Gheorghe Duca, and Diana Porubin. "Establishment of the Antioxidant/Antiradical Activity of the Inhibitors Using the DPPH–Radical." Chemistry Journal of Moldova 3, no. 1 (June 2008): 118–26. http://dx.doi.org/10.19261/cjm.2008.03(1).01.

Full text
Abstract:
This research paper presents the results of the investigation of antioxidant activities of various inhibitors, which are constituents of winery products: quercitin, rezveratrol, dihydroxyfumaric acid. Also, the antioxidant activity of tartaric and dihydroxyfumaric (DFH4) acids derivatives has been determined: sodium dihydroxyfumarate, dimethylic ester of DFH4 and dimethylic ester of tartaric acid. The enotannin extracts obtained from grape seeds have been evaluated: the non-oxidized enotannin extract Eneox and the oxidized one -Enoxil.
APA, Harvard, Vancouver, ISO, and other styles
47

Ceriani, Federica, Luca Casanova, and Marco Ormellese. "Use of Organic Acids as Additives for Plasma Electrolytic Oxidation (PEO) of Titanium." Coatings 14, no. 6 (June 3, 2024): 703. http://dx.doi.org/10.3390/coatings14060703.

Full text
Abstract:
The present study investigates the influence of organic acids, added to the electrolytic solution, on the structure, morphology, and corrosion behaviour of plasma electrolytic oxidation (PEO) coatings produced on titanium grade 2. Particular attention is paid to the role of functional groups in the modification of the oxide’s properties. For this reason, all three selected acids, namely glutaric, glutamic, and tartaric acid, display two carboxylic groups, thus they interact with the substrate material mainly through –COO− adsorption. However, glutamic acid also has an amine group, while tartaric acid has two hydroxyl groups. The presence of such additional functional groups is found to impact the formation of the PEO coatings. According to scanning electron microscopy (SEM) analyses, the number of defects and their dimension increase with an increasing number of active groups present in the organic molecules. Then, when glutaric acid with only two carboxyl groups, is employed as an additive, smaller pores are produced. The dimension of defects increases when glutamic and tartaric acid are used. X-ray diffraction (XRD) testing demonstrates that rutile and anatase are present in all the coatings and that when using tartaric acid, a relatively high level of amorphism is reached. The electrochemical and corrosion behaviours are evaluated by potentiodynamic polarization (PDP) and electrochemical impedance spectroscopy (EIS) performed in a heated sulphuric acid solution. It is found that all types of coatings provide protection against corrosion, with oxides produced using glutamic acid showing the lowest corrosion current density (0.58 mA·m−2) and low corrosion rate (1.02 μm·y−1).
APA, Harvard, Vancouver, ISO, and other styles
48

Journal, Baghdad Science. "Separation and Determination of Some Organic Acids in Dry Calyces of Iraqi Hibiscus Sabdariffa Linn." Baghdad Science Journal 12, no. 2 (June 7, 2015): 340–49. http://dx.doi.org/10.21123/bsj.12.2.340-349.

Full text
Abstract:
A new reversed phase- high performance liquid chromatographic (RP-HPLC) method with Ultraviolet-Visible spectrophotometry has been optimized and validated for the simultaneous extraction and determination of organic acids present in Iraqi calyces of Hibiscus Sabdraffia Linn. The method is based on using ultrasonic bath for extracting organic acids. Limit of detection in µg/ml of Formic acid, Acetic acid, Oxalic acid, Citric acid, Succinic acid, Tartaric acid, and Malic acid 126.8498×10-6, 113.6005×10-6, 97.0513×10-6, 49.7925×10-6, 84.0753×10-6, 92.6551×10-6, and 106.1633×10-6 ,respectively. The concentration of organic acids found in dry spacemen of calyces of Iraqi Hibiscus Sabdraffia Linn. under study: Formic acid, Acetic acid, Oxalic acid, Citric acid, Succinic acid, Tartaric acid, and Malic acid are 114.896 µg/g, 64.722 µg/g, 342.508 µg/g, 126.902 µg/g, 449.91 µg/g, 268.52 µg/g, and 254.07 µg/g respectively.
APA, Harvard, Vancouver, ISO, and other styles
49

Ibrahim, Sarah K., Berween A. Hasan, and Kareem D. Khalaf. "Separation and Determination of Some Organic Acids in Dry Calyces of Iraqi Hibiscus Sabdariffa Linn." Baghdad Science Journal 12, no. 2 (June 7, 2015): 340–49. http://dx.doi.org/10.21123/bsj.2015.12.2.340-349.

Full text
Abstract:
A new reversed phase- high performance liquid chromatographic (RP-HPLC) method with Ultraviolet-Visible spectrophotometry has been optimized and validated for the simultaneous extraction and determination of organic acids present in Iraqi calyces of Hibiscus Sabdraffia Linn. The method is based on using ultrasonic bath for extracting organic acids. Limit of detection in µg/ml of Formic acid, Acetic acid, Oxalic acid, Citric acid, Succinic acid, Tartaric acid, and Malic acid 126.8498×10-6, 113.6005×10-6, 97.0513×10-6, 49.7925×10-6, 84.0753×10-6, 92.6551×10-6, and 106.1633×10-6 ,respectively. The concentration of organic acids found in dry spacemen of calyces of Iraqi Hibiscus Sabdraffia Linn. under study: Formic acid, Acetic acid, Oxalic acid, Citric acid, Succinic acid, Tartaric acid, and Malic acid are 114.896 µg/g, 64.722 µg/g, 342.508 µg/g, 126.902 µg/g, 449.91 µg/g, 268.52 µg/g, and 254.07 µg/g respectively.
APA, Harvard, Vancouver, ISO, and other styles
50

Lord, Richard S., Cheryl K. Burdette, and J. Alexander Bralley. "Significance of Urinary Tartaric Acid." Clinical Chemistry 51, no. 3 (March 1, 2005): 672–73. http://dx.doi.org/10.1373/clinchem.2004.036368.

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