Academic literature on the topic 'Tartaric acid'

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

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources 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.

Journal articles on the topic "Tartaric acid"

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

Dissertations / Theses on the topic "Tartaric acid"

1

Xie, Wei. "Thermodynamic and structural studies of aqueous chelating agents and their metal complexes at various temperatures and pressures : diethylenetriaminepentaacetic acid (DTPA) and tartaric acid /." Thesis, National Library of Canada = Bibliothèque nationale du Canada, 1999. http://www.collectionscanada.ca/obj/s4/f2/dsk1/tape7/PQDD_0030/MQ47491.pdf.

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

Mhatre, Bharat Suresh. "Super-enantiospecific Autocatalytic Decomposition of Tartaric Acid and Aspartic Acid on Cu Surfaces." Research Showcase @ CMU, 2013. http://repository.cmu.edu/dissertations/232.

Full text
Abstract:
Autocatalytic reaction mechanisms are observed in a range of important chemical processes including catalysis, radical-mediated explosions, and biosynthesis. Because of their complexity, the microscopic details of autocatalytic reaction mechanisms have been difficult to study on surfaces and heterogeneous catalysts. Autocatalytic decomposition reactions of tartaric acid (TA) enantiomers adsorbed on Cu(110) offer molecular-level insight into these processes, which until now, were largely a matter of speculation. The decomposition of TA/Cu(110) is initiated by a slow, irreversible process that forms vacancies in the adsorbed TA layer, followed by a vacancy-mediated, explosive decomposition process that yields CO2 and small hydrocarbon products. Initiation of the explosive decomposition of TA/Cu(110) was studied by measurement of the reaction kinetics, time-resolved low energy electron diffraction (LEED), and time-resolved scanning tunneling microscopy (STM). Initiation results in a decrease in the local coverage of TA and a concomitant increase in the areal vacancy concentration. Once the vacancy concentration reaches a critical value, the explosive, autocatalytic decomposition step dominates the subsequent TA decomposition rate. Aspartic acid is an excellent probe molecule for investigating the surface chemistry of autocatalytic reactions . Because a wide range of isotopically labeled varieties of aspartic acid are commercially available, we have been able to conduct a detailed investigation of its autocatalytic reaction mechanism. Experimental data obtained for variable initial coverage, variable heating rate and isothermal TPRS experiments, while monitoring CO2 desorption is in excellent agreement with a rate law which explicitly accounts for an initiation step and an explosion step which is second order in vacancy concentration Autocatalytic surface explosion mechanisms can be exploited to attain extremely high enantiospecificities in the case of TA decomposition on naturally chiral Cu(hkl)R&S surfaces. Interaction energies of chiral molecules with naturally chiral surfaces are small and typically lead to modest enantioselectivities. However, the highly non-linear kinetics of autocatalytic reaction mechanism amplifies these small differences to result in high enantiospecificities. The observed phenomenon has the characteristics of autocatalytic processes that have been postulated to lead to biomolecular homochirality in life on Earth; processes with relatively small differences in reaction energetics that, nonetheless, lead to extremely high enantioselectivity.
APA, Harvard, Vancouver, ISO, and other styles
3

Basaran, Tolga Yener. "Ion Exchangers In The Recovery Of Tartaric Acid From Aqueous Solutions." Master's thesis, METU, 2006. http://etd.lib.metu.edu.tr/upload/12607367/index.pdf.

Full text
Abstract:
Tartaric acid is a dicarboxylic acid naturally present in grapes, and has many application areas with its salts. It can be produced synthetically, manufactured as a by-product in wine industry, or can be recovered by electrodialysis and solvent extraction methods. Since, ion exchange is one of the oldest processing techniques for the recovery and purification of valuable materials, it can be applied to obtain this valuable organic acid. In this study it is aimed to investigate the effects of resin basicity, initial concentration, and initial pH of the solution on ion exchange equilibrium. The model tartaric acid solutions were prepared for the equilibrium analysis with two different anion exchange resins in a batch type system. A shaker bath at 28 oC with 300-rpm agitation rate was used. The weakly basic resin Lewatit MP62, and strongly basic resin Lewatit M511, which are in polystyrene structure, was obtained from the producer Bayer AG. In the analysis, Shimadzu PDA Detector at 210 nm with Waters Atlantis dC18 column was used. 20 mM NaH2PO4 at pH = 2.7 was introduced to the HPLC as the mobile phase at 0.5 ml/min flow rate. In the investigation of the resin basicity, MP62 presented better performance than M511. The equilibrium experiments were performed at three different initial acid concentrations (0.01, 0.02, and 0.10 M) for both resin, and in the pH ranges pH <
pKa1, pKa1 <
pH <
pKa2, and pKa2 <
pH for weakly basic resin, and in the pH ranges pH <
pKa1, pKa1 <
pH <
pKa2 for strongly basic resin at each concentration. Results show that the pH of the solution is a more important parameter than the initial concentration that affects the ion exchange equilibrium. Also, Langmuir and Freundlich isotherms were plotted, and it was shown that they were in good agreement with the experimental data especially for the systems that are at low total ion concentrations.
APA, Harvard, Vancouver, ISO, and other styles
4

Bakka, Thomas Aleksander. "Synthesis and Mechanistic Studies of Optically Active Tartaric Acid Based Surfactants." Thesis, Norges Teknisk-Naturvitenskaplige Universitet, 2013. http://urn.kb.se/resolve?urn=urn:nbn:no:ntnu:diva-20995.

Full text
Abstract:
Gaining knowledge of the isomerization mechanisms are of pivotal importancefor stereochemically pure synthesis. Optically pure tartaric acid basedsurfactants are seen as important precursors for optically active superstructures.From these superstructures, liposomes are known canditates for drugdelivery nanovessels in the human body.A number of optically pure reference compounds were synthesized and severalanalysis methods were outlined.A mechanistic route for formation of cyclic tartrimides in nonpolar solvents have beenproposed on the background of various mechanistic assessment techniques.
APA, Harvard, Vancouver, ISO, and other styles
5

Lopez, Martinez Marco Antonio. "Studies of tartaric acid modified nickel supported catalysts for enantioselective hydrogenation reactions." Thesis, University of Birmingham, 2011. http://etheses.bham.ac.uk//id/eprint/1491/.

Full text
Abstract:
This thesis investigates the catalytic properties of tartaric acid‐nickel supported catalysts, obtained from hydrotalcite‐like compound precursors, in the enantioselective hydrogenation of methyl acetoacetate to methyl 3‐hydroxybutyrate. Variables of reaction during modification such as pH and tartaric acid concentration, as well as Ni particle size above a minimum threshold of ca. 20 nm, proved not to have a major effect on enantioselectivity. However, the nature of the cations constituting the catalyst support was found to influence the enantioselectivity observed. Specifically, when iron or chromium were constituents of the supporting oxide matrix, enantioselectivities were found to be much higher. For systems containing nickel, magnesium (or zinc) and aluminium as the cations present in the parent hydrotalcite phase, when a series of materials of the same composition obtained from different synthetic methods, the urea hydrolysis method leads to catalysts with enantiodifferentiation ability, whereas materials prepared by the coprecipitation method does not. Also, this thesis researches the use of different types of ordered mesoporous silicas as supports of tartaric‐acid nickel in the aforementioned reaction. Even though the techniques of metal deposition explored did not allow incorporation of Ni in the internal surface of the materials, it was found that the morphology of the support plays an important role in enantioselectivity. In addition, for a given material, the incorporation of Ni via solid state reaction resulted in a catalyst with improved catalytic properties compared to one prepared by wet impregnation techniques.
APA, Harvard, Vancouver, ISO, and other styles
6

Hayes, Monty. "Chiral and racemic calix[6]arenes and their self-assembly /." View online, 2008. http://ecommons.txstate.edu/bchemtad/5.

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

Wade, Charles. "Studies towards the total synthesis of isoavenaciolide and the development of the amino tartrate aldol reaction." Thesis, University of Sheffield, 2001. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.341816.

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

Japu, Cristina. "Cyclic derivatives of D-glucose and tartaric acid as building blocks for renewable polyesters." Doctoral thesis, Universitat Politècnica de Catalunya, 2014. http://hdl.handle.net/10803/284660.

Full text
Abstract:
Three series of aromatic copolyesters derived from poly(ethylene terephthalate) (PET), poly(butylene terephthalate) (PBT) and poly(hexamethylene terephthalate) (PHT) have been synthesized by melt polycondensation in which the terephthalate and oxyalkylene units have been partially or totally replaced by monocyclic and bicyclic diacids and diols obtained by derivatization of renewable monomers such as tartaric acid and D-glucose respectively. Another series of aliphatic copolyesters derived from poly(butylene sebacate) has been prepared by both melt polycondensation and enzymatic polymerization in solution where both the sebacate and butylene units have been partially or totally replaced by the bicyclic diacid or diol derived from D-glucose. All series of polyesters and copolyesters were characterized by FTIR and NMR in solution. The chemical composition of the copolyesters was determined by 1H NMR. On the other hand the microstructure, determined by 13C NMR was observed to be at random for all series. The molecular weights determined by intrinsic viscosimetry and gel permeation chromatography (GPC) were high but decreasing when the content of the sugar monomer increased. The thermal properties were evaluated by differential scanning calorimetry (DSC) and thermogravimetric analysis. It was observed that the thermal stability was not affected when the oxyalkylene unit was the replaced unit, but decreased slightly when the terephthalate unit was replaced one. The new copolyesters showed higher glass transition temperatures than the parent poly(alkylen terephthalate)s, being superiors for the copolyesters obtained from the bicyclic derivative of D-glucose than for the monocyclic derivative of tartaric acid. All copolyesters were less crystalline than the parent homopolyesters, with both melting and enthalpies decreasing with the increased content of the sugar units. The crystalline structure determined by X-ray diffraction was similar to the PET, PBT and PHT polyesters. Finally it was studied the hydrolytic and enzymatic degradability by weight loss, molecular weight determinations and scanning electron microscopy (SEM). All new copolyesters showed enhanced degradability in water and biodegradability.
En esta Tesis se han sintetizado mediante reacción de policondensación en masa tres series de copoliesteres aromáticos derivados del poli(etilen tereftalato) (PET), poli(butilen tereftalato) (PBT) y poli(hexametilen tereftalato) (PHT) en los que se han sustituido parcial o totalmente las unidades tereftalicas o las unidades oxialquilénicas por diácidos y dioles cíclicos y bicíclicos obtenidos por derivatización de monómeros de origen natural como son el ácido tartárico y la D-glucosa. También se han sintetizado mediante policondensación en masa y en disolución por vía enzimática una serie de copoliesteres alifáticos derivados del poli(butilen sebacato) en los que las unidades oxibutilénicas y sebácicas se han reemplazado por dioles y diácidos bicíclicos derivados de la D-glucosa. Todas las series se han caracterizado espectroscópicamente mediante FTIR y RMN en disolución. Mediante RMN de 1H se ha determinado la composición de los copolímeros. Por otro lado mediante RMN de 13C se ha estudiado la microestructura de los mismos. Dicho estudio reveló que todos los copolímeros eran al azar. Los pesos moleculares determinados mediante cromatografía de permeabilidad en gel y viscosimetría capilar resultaron ser aceptables, aunque en todos los casos disminuían a medida que aumentaba el contenido del monómero de origen renovable. Las propiedades térmicas se han evaluado mediante calorimetría diferencial de barrido (DSC) y análisis termogravimétrico. Se ha observado que la estabilidad térmica no se ve apenas afectada cuando la sustitución se realiza sobre la unidad oxialquilénica y que por otro lado desciende cuando es la unidad tereftálica la sustituida. Los nuevos copolímeros presentaron temperaturas de transición vítrea superiores, observándose que este aumento era superior en los copolímeros que incorporaban el diol o el diácido bicíclico derivado de la D-glucosa que en los que incorporaban el diol o el diácido monociclico derivado del ácido tartárico. Los copolimeros eran por lo general menos cristalinos, menos cristalizables y con temperaturas de fusión inferiores. La estructura cristalina de los mismos determinada mediante difracción de rayos-X era semejante a la de los homopolimeros PET, PBT y PHT. Se ha estudiado la degradabilidad hidrolitica y la biodegradabilidad de todos ellos mediante medidas de pérdida de peso, pesos moleculares, microscopía electrónica de barrido (SEM), observándose un aumento de ambas propiedades para los nuevos copolímeros.
APA, Harvard, Vancouver, ISO, and other styles
9

Morello, Alessandro. "Influence of ph and temperature on metatartaric acid efficiency in white wine tartaric stabilization." Master's thesis, ISA/UL, 2012. http://hdl.handle.net/10400.5/8625.

Full text
Abstract:
Mestrado Vinifera EuroMaster - Instituto Superior de Agronomia
The sensitivity of metatartaric acid (MA) to high temperatures is the main limit for its actual use in enology. For this reason MA is generally used only for ready-to-drink wines that are stored for few months in bottle. The objective of this work was to obtain more information about the use of MA in order to prevent tartaric salts precipitation in wine by monitoring its effectiveness along the time. Tartaric stability was followed in a white wine during a ten-week experiment. 10 g/Hl of MA was added to wines with different five pH values, namely 3.0, 3.2, 3.5, 3.7 and 3.9 at room temperature (20°C), and to wines at original pH of 3.2 under three different temperatures commonly found in real storage conditions, precisely 12°C, 20°C and 35°C. To monitor tartaric stability we used a test based on the electrical conductivity developed by Boulton (1983). At the end of the experiment we found both a pH and a temperature effect. It was found a polynomial relation (R2 = 0,85) between tartaric stability and pH. The highest the pH, the highest the instability and the shortest the protection against tartrates precipitation. Furthermore, over 20°C we observed a rapid general decrease of MA effectiveness and that there was a linear relation (R2 = 0,99) between temperature and tartaric stability, so that we can affirm that the highest the temperature, the higher the wine instability; this is probably due to a faster MA hydrolysis. Metatartaric acid is surely able to prevent tartrates precipitation, but since in our experimental conditions it could not protect the wine for longer than one month, we would suggest extreme prudence about its use especially during long storage periods
APA, Harvard, Vancouver, ISO, and other styles
10

Guadagnin, Hellen Costerano. "Corrosion resistance study of AA2524 anodized in sulphuric-tartaric acid and sealed with hybrid coatings." Universidade de São Paulo, 2017. http://www.teses.usp.br/teses/disponiveis/3/3137/tde-20072017-152947/.

Full text
Abstract:
Aluminium alloys are widely used in the aerospace industry due to their lightweight and high specific strength. However, these alloys are particularly sensitive to localized corrosion in chloride environments and need to be protected by a robust system. One of the protection methodologies consists in anodizing. The produced layer increases the corrosion resistance and also serves as anchoring site for organic coatings application. Chromium-based anodizing has been usually employed, nevertheless, as chromate compounds are toxic for health and the environment, chromium-based surface treatments will be prohibited in the aerospace industry in a near future. Tartaric-sulphuric acid (TSA) anodizing is a promising environment compliant alternative, which is already being used at industrial level with appropriate corrosion protection and paint adhesion properties. This study aims at proposing a hybrid sol-gel treatment after TSA anodizing of AA2524 specimens in order to improve the corrosion resistance of the anodized layer while maintaining its compatibility with organic coatings. For this, anodic aluminium oxides (AAO) were produced at different anodizing voltages and protected by dip-coating with a hybrid sol-gel layer obtained from a tetraethoxysilane (TEOS) and 3-glycidoxypropyltrimethoxysilane (GPTMS) hydrolysis solution with high water content. Corrosion resistance evaluation was carried out by means of electrochemical impedance spectroscopy (EIS) in NaCl 0.1 mol L-1 and salt-spray chamber exposure (ASTM B117-11 standard). The morphology of the anodic porous layer was investigated by means of FE-SEM, whereas glow discharge optical emission spectroscopy (GDOES) was employed to evaluate the distribution of the sol-gel layer within the porous AAO. FE-SEM characterization confirmed that the layer properties (pore distribution, porosity and thickness) were strongly dependent on the anodizing conditions, whereas GDOES depth profile showed penetration of the hybrid coating within the pores of the anodized layer. The two characterization techniques showed inefficient surface sol-gel coverage for the samples anodized at higher voltage, likely due to insufficient sol-gel deposition. The results of the EIS characterization tests up to 1008 h (42 days) showed that, irrespectively to the anodizing voltage, the hybrid sol-gel protected AAO was stable with only slight evolution of the diagrams with immersion time. Moreover, the hybrid coating protected samples presented higher low frequency impedance modulus than hydrothermally sealed (HTSed) reference TSA anodized samples, which was confirmed by electrical equivalent circuit (EEC) fitting of the EIS data. EEC fitting also revealed that the resistance of the pores to electrolyte penetration was increased by the application of the sol-gel coating when compared to the resistance of the HTSed reference samples and indicated better anticorrosion performance for the sample anodized at 16 V. These results were confirmed by the salt-spray tests. Investigation on the ageing of the hybrid sol-gel hydrolysis solution showed that its viscosity hardly changed up to two weeks of test and that hybrid coatings applied from these solutions were stable and afforded good corrosion protection to the TSA anodized substrate, an improvement of the anticorrosion properties of the hybrid coating was verified for an ageing time of 168 h. Preliminary tests performed with a solvent-free organic coating (epoxy) indicated good compatibility with the hybrid TEOS-GPTMS coating characterized by very high impedance and good stability upon exposure to the salt-spray chamber.
Ligas de alumínio são muito utilizadas na indústria aeronáutica por serem materiais leves e altamente resistentes. Porém, essas ligas são particularmente sensíveis à corrosão localizada em meios que contêm cloretos, e precisam de sistemas robustos de proteção. Uma das metodologias de proteção consiste em anodização. A camada produzida aumenta a resistência à corrosão e também serve como sítio de ancoragem para aplicação de revestimentos orgânicos. A anodização crômica tem sido usualmente empregada na indústria aeronáutica. No entanto, como compostos contendo íons cromato são tóxicos para a saúde e para o meio-ambiente, tratamentos de superfície à base de cromo serão proibidos na indústria espacial em um futuro próximo. Anodização em banho de ácido sulfúrico-tartárico (TSA) é uma alternativa promissora e ambientalmente compatível, a qual já está sendo usada industrialmente com apropriada proteção à corrosão e adesão para pintura. Este estudo tem como objetivo propor um tratamento utilizando um revestimento híbrido sol-gel para melhorar a resistência à corrosão da liga AA2524 anodizada em TSA e que mantenha sua compatibilidade com revestimentos orgânicos. Para isso, camadas anodizadas de alumínio (CAA) foram produzidas em diferentes voltagens e protegidas por camada de híbrido sol-gel obtida pela hidrólise de tetraetilortosilano (TEOS) e glicidóxipropiltrimetóxisilano (GPTMS) em solução com alto teor de água e aplicada pela técnica de dip-coating. A avaliação da resistência à corrosão foi realizada através de espectroscopia de impedância eletroquímica (EIS) em NaCl 0,1 mol.L-1 e por exposição à câmara de névoa salina (norma ASTM B117-11). A morfologia da camada porosa foi investigada por MEV e a espectroscopia de emissão óptica por descarga luminescente (GDOES) foi empregada para avaliar a distribuição do híbrido sol-gel no interior dos poros da camada porosa. As caracterizações por MEV confirmaram que as propriedades da camada (distribuição dos poros, porosidade e espessura) são fortemente dependentes das condições de anodização, e a composição em profundidade obtida por GDOES mostrou que o revestimento híbrido penetrou nos poros da camada anodizada. As duas técnicas de caracterização mostraram uma cobertura ineficiente da camada sol-gel para as amostras anodizadas nas voltagens mais elevadas, provavelmente devido à deposição insuficiente do híbrido. Os testes de EIS com duração de até 1008 h (42 dias) mostraram que, independentemente da voltagem de anodização empregada, a camada anódica coberta com sol-gel ficou estável ocorrendo apenas pequenas evoluções dos diagramas com o tempo de imersão. Além do mais, as amostras protegidas com o revestimento híbrido apresentaram maiores valores de módulo de impedância em baixa frequência do que as amostras anodizadas em TSA e hidrotermicamente seladas (HTsed) usadas como referências. Essa tendência foi confirmada pelo ajuste com circuitos elétricos equivalentes (EEC) dos resultados de EIS que também mostrou que a aplicação do sol-gel híbrido torna mais difícil a penetração do eletrólito agressivo nos poros da camada anodizada quando comparada com as amostras HTSed, e indicou melhor desempenho anticorrosivo para a amostra anodizada em 16 V. Esses resultados foram confirmados pelos testes de névoa salina. A investigação do envelhecimento da solução de sol-gel mostrou pouca mudança na viscosidade da solução de hidrólise em duas semanas de testes e que os revestimentos híbridos aplicados a partir dessas soluções foram estáveis e promoveram boa proteção à corrosão para as amostras anodizadas em TSA, com melhora das propriedades anticorrosivas após 168 h de envelhecimento. Testes preliminares realizados com revestimento orgânico livre de solvente (epóxi) indicaram boa compatibilidade deste com o revestimento híbrido TEOS-GPTMS. O revestimento epóxi propiciou valores de módulo de impedância elevados e estáveis e também boa estabilidade após exposição à câmara de névoa salina quando aplicado sobre o revestimento híbrido aplicado sobre a liga 2524.
APA, Harvard, Vancouver, ISO, and other styles

Books on the topic "Tartaric acid"

1

1874-1951, Kenrick Frank B., ed. The application of polarimetry to the estimation of tartaric acid in commercial products. Toronto: [s.n.], 1997.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
2

Gawroński, Jacek. Tartaric and malic acids in synthesis: A source book of building blocks, ligands, auxiliaries, and resolving agents. New York: Wiley, 1999.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
3

Melnick, Jason Patrick. Conformational study of bis-allylic and mono-allylic compounds by variable temperature nuclear magnetic resonance. 1994.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
4

Parker, Philip M. The 2007 Import and Export Market for Lactic Acid, Tartaric Acid, Citric Acid, and Their Salts and Esters in China. ICON Group International, Inc., 2006.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
5

Parker, Philip M. The 2007 Import and Export Market for Lactic Acid, Tartaric Acid, Citric Acid, and Their Salts and Esters in India. ICON Group International, Inc., 2006.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
6

The World Market for Lactic Acid, Tartaric Acid, Citric Acid, and Their Salts and Esters: A 2004 Global Trade Perspective. Icon Group International, Inc., 2005.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
7

Parker, Philip M. The World Market for Lactic Acid, Tartaric Acid, Citric Acid, and Their Salts and Esters: A 2007 Global Trade Perspective. ICON Group International, Inc., 2006.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
8

Parker, Philip M. The 2007 Import and Export Market for Lactic Acid, Tartaric Acid, Citric Acid, and Their Salts and Esters in United States. ICON Group International, Inc., 2006.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
9

Gawroński, Jacek, and Krystyna Gawrońska. Tartaric and Malic Acids in Synthesis: A Source Book of Building Blocks, Ligands, Auxiliaries, and Resolving Agents. Wiley-Interscience, 1999.

Find full text
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Tartaric acid"

1

Bährle-Rapp, Marina. "Tartaric Acid." In Springer Lexikon Kosmetik und Körperpflege, 545. Berlin, Heidelberg: Springer Berlin Heidelberg, 2007. http://dx.doi.org/10.1007/978-3-540-71095-0_10337.

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

Gooch, Jan W. "Tartaric Acid." In Encyclopedic Dictionary of Polymers, 730. New York, NY: Springer New York, 2011. http://dx.doi.org/10.1007/978-1-4419-6247-8_11570.

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

Winkelmann, Jochen. "Diffusion coefficient of D-tartaric acid into L-tartaric acid and water." In Diffusion in Gases, Liquids and Electrolytes, 3267. Berlin, Heidelberg: Springer Berlin Heidelberg, 2018. http://dx.doi.org/10.1007/978-3-662-54089-3_2639.

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

Holze, Rudolf. "Ionic conductance of tartaric acid." In Electrochemistry, 500. Berlin, Heidelberg: Springer Berlin Heidelberg, 2016. http://dx.doi.org/10.1007/978-3-662-49251-2_468.

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

Zoecklein, Bruce W., Kenneth C. Fugelsang, Barry H. Gump, and Fred S. Nury. "Tartaric Acid and Its Salts." In Production Wine Analysis, 289–315. Boston, MA: Springer US, 1990. http://dx.doi.org/10.1007/978-1-4615-8146-8_13.

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

Hoffmann, Marcin, and Jacek Rychlewski. "Effects of Solvation for (R,R) Tartaric-Acid Amides." In New Trends in Quantum Systems in Chemistry and Physics, 189–210. Dordrecht: Springer Netherlands, 2000. http://dx.doi.org/10.1007/0-306-46950-2_11.

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

Pardasani, R. T., and P. Pardasani. "Magnetic properties of chromium(III) complex of D(-)tartaric acid." In Magnetic Properties of Paramagnetic Compounds, 213. Berlin, Heidelberg: Springer Berlin Heidelberg, 2015. http://dx.doi.org/10.1007/978-3-662-45972-0_158.

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

Pardasani, R. T., and P. Pardasani. "Magnetic properties of chromium(III) complex of D(-)tartaric acid." In Magnetic Properties of Paramagnetic Compounds, 212. Berlin, Heidelberg: Springer Berlin Heidelberg, 2015. http://dx.doi.org/10.1007/978-3-662-45972-0_157.

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

Pardasani, R. T., and P. Pardasani. "Magnetic properties of dinuclear chromium(III) complex with acetylacetone and tartaric acid." In Magnetic Properties of Paramagnetic Compounds, 407–8. Berlin, Heidelberg: Springer Berlin Heidelberg, 2017. http://dx.doi.org/10.1007/978-3-662-54228-6_231.

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

Pardasani, R. T., and P. Pardasani. "Magnetic properties of dinuclear chromium(III) complex with acctylacetone and tartaric acid." In Magnetic Properties of Paramagnetic Compounds, 409–10. Berlin, Heidelberg: Springer Berlin Heidelberg, 2017. http://dx.doi.org/10.1007/978-3-662-54228-6_232.

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

Conference papers on the topic "Tartaric acid"

1

Krug, W. P., and M. Aronson. "Complex Piezoelectric Response of Tartaric Acid Composites." In Sixth IEEE International Symposium on Applications of Ferroelectrics. IEEE, 1986. http://dx.doi.org/10.1109/isaf.1986.201134.

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

Soltani, A., D. Gebauer, B. Fischer, H. Colfen, and M. Koch. "Monitoring the crystallization of tartaric acid with THz spectroscopy." In 2017 42nd International Conference on Infrared, Millimeter, and Terahertz Waves (IRMMW-THz). IEEE, 2017. http://dx.doi.org/10.1109/irmmw-thz.2017.8066993.

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

"CSA-Based Mortars Manufactured with Tartaric Acid-Based Retarder." In "SP-329: Superplasticizers and Other Chemical Admixtures in Concrete Proceedings Twelfth International Conference, Beijing, China". American Concrete Institute, 2018. http://dx.doi.org/10.14359/51711228.

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

Alonso, José, Santiago Mata, Elena Alonso, Verónica Díez, and Vanessa Cortijo. "FOUR STRUCTURES OF TARTARIC ACID REVEALED IN THE GAS PHASE." In 72nd International Symposium on Molecular Spectroscopy. Urbana, Illinois: University of Illinois at Urbana-Champaign, 2017. http://dx.doi.org/10.15278/isms.2017.wc05.

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

Permana, Ahmadi Jaya, Harsasi Setyawati, Hamami, and Irmina Kris Murwani. "The influence of dicarboxylic acids: Oxalic acid and tartaric acid on the compressive strength of glass ionomer cements." In 5TH INTERNATIONAL CONFERENCE AND WORKSHOP ON BASIC AND APPLIED SCIENCES (ICOWOBAS 2015). AIP Publishing LLC, 2016. http://dx.doi.org/10.1063/1.4943317.

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

Cao, Binghua, and Mengbao Fan. "Quantitative analyses of tartaric acid based on terahertz time domain spectroscopy." In 5th International Symposium on Advanced Optical Manufacturing and Testing Technologies, edited by Yudong Zhang, José Sasián, Libin Xiang, and Sandy To. SPIE, 2010. http://dx.doi.org/10.1117/12.864957.

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

Rajesh, K., A. Mani, K. Anandan, K. Gayathri, and A. Arun. "Role of metal and amino acid on the growth and microhardness properties of tartaric acid crystals." In 7TH NATIONAL CONFERENCE ON HIERARCHICALLY STRUCTURED MATERIALS (NCHSM-2019). AIP Publishing, 2019. http://dx.doi.org/10.1063/1.5114601.

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

Inomata, Katsuhiko, and Yutaka Ukaji. "Asymmetric 1,3-Dipolar Cycloaddition Utilizing Tartaric Acid Ester as a Chiral Auxiliary." In The 4th International Electronic Conference on Synthetic Organic Chemistry. Basel, Switzerland: MDPI, 2000. http://dx.doi.org/10.3390/ecsoc-4-01796.

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

G. Babayan, Bella. "Tartaric Acid New Synthetic Derivatives Antibacterial Activity against the Phytopathogenic Pseudomonas syringae." In 2nd International Conference on Advanced Research in Science, Engineering and Technology. Acavent, 2021. http://dx.doi.org/10.33422/2nd.icarset.2021.03.120.

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

Meng, Fanqing, and Mengkai Lu. "Characterization of linear and nonlinear optical properties of urea-tartaric-acid single crystals." In Photonics China '96, edited by Manfred Eich, Bruce H. T. Chai, and Minhua Jiang. SPIE, 1996. http://dx.doi.org/10.1117/12.252993.

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