Academic literature on the topic 'Esterification Reactivity'

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 'Esterification Reactivity.'

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 "Esterification Reactivity"

1

Zhan, Shaoqi, Xiaochun Tao, Liangzhen Cai, Xiaohui Liu, and Taoping Liu. "The carbon material functionalized with NH2+ and SO3H groups catalyzed esterification with high activity and selectivity." Green Chem. 16, no. 11 (2014): 4649–53. http://dx.doi.org/10.1039/c4gc01395f.

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

Mei, Sheng-Fu, Jin Liu, Zhen Li, Yan Chen, and Jing Zhang. "Reactivity of 2-Phosphonobutane-1,2,4-tricarboxylic Acid Esterification." Asian Journal of Chemistry 26, no. 5 (2014): 1530–32. http://dx.doi.org/10.14233/ajchem.2014.17279.

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

Li, Chong, Yueping Jiang, Baoshan Huang, Menghang Zhang, Yanhong Feng, and Zhitao Yang. "Preparation and Properties of Jute Fiber Long-Chain Fatty Acid Esters in Supercritical Carbon Dioxide." Materials 12, no. 9 (May 8, 2019): 1499. http://dx.doi.org/10.3390/ma12091499.

Full text
Abstract:
A two-step method involving continuous screw-extrusion steam explosion (CSESE) pretreatment and esterification in supercritical carbon dioxide (scCO2) is used to prepare long-chain fatty acid-modified jute fiber. The weight gain percentage (WG %) of CSESE-pretreated jute laurate (JL) was 110.7% when esterification was carried out in scCO2 at 14 MPa and 100 °C for 2 h. The corresponding WG % was 105.5% when esterification was instead carried out in pyridine at 100 °C for 2 h. Scanning electron microscopy and X-ray diffraction indicated that CSESE pretreatment enhanced the reactivity of jute fiber, with esterification in scCO2 simultaneously occurring on the fibers surface and internal walls. The glass transition temperature of esterified jute was approximately 119 °C, indicating that it could be hot processed over a wide temperature range. The esterified jute had an oil absorption ratio of 17.01 g/g, so it can be used as an oil absorption material.
APA, Harvard, Vancouver, ISO, and other styles
4

Roswanda, Robby, Ilham Ardatul Putra, Maria Mardiastuti, and Didin Mujahidin. "Catalyst Free Hydroxyl Protection of Quinine via Esterification." Key Engineering Materials 811 (July 2019): 3–7. http://dx.doi.org/10.4028/www.scientific.net/kem.811.3.

Full text
Abstract:
A method to protect the hydroxyl group of quinine via esterification is developed. The method uses acetyl and benzoyl as the protection group. The method employs no catalyst that generates reasonable yield at 83% for acetyl and 73% for benzoyl. This catalyst free method emphasizes on the importance substrate reactivity to achieve free catalyst procedure. Ester form of quinine synthesized might be further functionalized for various aims in accordance to its rich functional group and building block of quinine.
APA, Harvard, Vancouver, ISO, and other styles
5

Kastratović, Vlatko, Milica Radulović, and Kristina Kastratović. "Esterification of propanoic acid in the presence of a homogeneous catalyst." Kragujevac Journal of Science, no. 44 (2022): 45–55. http://dx.doi.org/10.5937/kgjsci2244045k.

Full text
Abstract:
Esters are organic compounds that are widely used even on an industrial scale, so their production has been extensively investigated. The aim of this work is to carry out the optimization process of esterification of propionic acid with lower monohydric alcohols. The influence of the amount of catalyst (H2SO4), the size and structure of the alcohol, the influence of the a-substituent, the molar acid/alcohol ratio and temperature on the esterification process were investigated. The descending order of reactivity tested alcohol is: 1-butanol> 1-propanol> ethanol> 2-propanol. As the acid/alcohol molar ratio increases, the rate and yield of the esterification reaction increase. The maximum yield of n-propyl propanoate in our experiments was 96.9%, achieved at a molar ratio of propanoic acid/1-propanol/catalyst 1/10/0.20 and a temperature of 65°C for 210 minutes of reaction.
APA, Harvard, Vancouver, ISO, and other styles
6

Vázquez, Ester, and Maurizio Prato. "Functionalization of carbon nanotubes for applications in materials science and nanomedicine." Pure and Applied Chemistry 82, no. 4 (March 13, 2010): 853–61. http://dx.doi.org/10.1351/pac-con-09-10-40.

Full text
Abstract:
Carbon nanotubes (CNTs) can be functionalized using a variety of efficient protocols. Especially, esterification and amidation reactions are exploited along with 1,3-dipolar cycloadditions. The use of microwaves (MWs) to activate the reactivity of CNTs is also reported. Innovative NMR methodologies can be introduced to investigate the covalent attachment of organic moieties to CNTs.
APA, Harvard, Vancouver, ISO, and other styles
7

Wu, Yan, Shiang He, Dongmei Li, Yang Li, and Hao Wang. "Esterification of naphthenic acids with various structures over tungstophosphoric acid-intercalated layer double hydroxide catalysts with various interlayer spacings." Clay Minerals 56, no. 3 (September 2021): 250–59. http://dx.doi.org/10.1180/clm.2022.3.

Full text
Abstract:
AbstractTungstophosphoric acid-intercalated MgAl layer double hydroxides (LDHs) are active catalysts for removing naphthenic acids (NAs) from petroleum via esterification. Due to their active sites being in the interlayer, the interlayer spacing of LDHs might affect their activity, particularly for NAs with various structures. Herein, two tungstophosphoric acid-intercalated MgAl LDHs with various interlayer spacings (d003 = 1.46 and 1.07 nm) synthesized by varying the ion-exchange time were used as catalysts for esterification between NAs and ethylene glycol. Six NAs with various side chains and rings were used as model compounds to investigate the effects of NA structures and d003 values on the activity of LDHs. In general, NAs with large molecule sizes and steric hindrances are less reactive over the same catalyst. The LDH with a larger d003 value favours the esterification of NAs regardless of their structure, particularly NAs with large molecule sizes and steric hindrances. However, a large d003 is less effective for esterification of NAs with conjugated carboxyl groups. An enlarged interlayer space might facilitate NA molecules to access the interlayer of LDHs so as to come into contact with the catalytic sites, making this process responsible for the enhanced reactivity. The esterification kinetics of cyclohexanecarboxylic acid over these LDHs follow a first-order reaction. The activation energies for the LDHs with large and small d003 values are 26.25 and 32.18 kJ mol–1, respectively.
APA, Harvard, Vancouver, ISO, and other styles
8

Kastratovic, Vlatko, and Miljan Bigovic. "Esterification of stearic acid with lower monohydroxylic alcohols." Chemical Industry and Chemical Engineering Quarterly 24, no. 3 (2018): 283–91. http://dx.doi.org/10.2298/ciceq170327040k.

Full text
Abstract:
Esters play a significant role in everyday life but also in the chemical industry. The aim of this study is to investigate the influence of different parameters on the process of esterification of higher monocarboxylic acids with lower monohydroxylic alcohols. We examined the influences of the following variables: the type and amount of the catalyst, the structure of alcohols and fatty acids, the acid/alcohol molar ratio, and the temperature of the esterification process. The descending order of reactivity found alcohols is: 1-butanol > 1-propanol > 2-methyl-1-propanol > ethanol > 2-butanol >2-propanol > 2-methyl-2-propanol. The results of this study show no significant effect of chain lengths of saturated fatty acids on the speed and yield of esterification. The presence of the double bond in unsaturated fatty acids reduces the acid to ester conversion. The highest yield (99%) was obtained in the reaction of stearic acid and 1-butanol with an acid/alcohol/catalyst (H2SO4) mole ratio 1/15/0.75 and at a temperature of 65?C.
APA, Harvard, Vancouver, ISO, and other styles
9

Breitenlechner, Stefan, and Thorsten Bach. "Kinetic Study on the Esterification of Hexanoic Acid with N,N-Dialkylamino Alcohols: Evidence for an Activation by Hydrogen Bonding." Zeitschrift für Naturforschung B 61, no. 5 (May 1, 2006): 583–88. http://dx.doi.org/10.1515/znb-2006-0513.

Full text
Abstract:
The pseudo-first order rate constant for the esterification of hexanoic acid (1) and five different N,N-dialkylamino alcohols (2) was determined in comparison to 1-hexanol (k = 0.67 · 10−5 s−1). The values range from 0.60 · 10−5 s−1 to 9.3 · 10−5 s−1. The data suggest a differing reactivity for structurally related compounds, which is directly correlated to the ability of the corresponding amino alcohol to activate the carboxylic acid by hydrogen bonding. A seven-membered transition state C≠ is postulated for reactions of 2-amino alcohols. The fastest reaction was observed for trans-2-(N,N-dimethylamino) cyclohexanol (2e), in which the amino and the hydroxyl groups are in an almost perfect synperiplanar 1,2-position. Attempts to further enhance the rate of the esterification by the addition of potential catalysts failed. Only Cu(OTf)2 (2.5 mol-%) allowed for a moderate rate increase from 7.5 · 10−5 s−1 (uncatalyzed) to 14.8 · 10−5 s−1 (catalyzed) in the esterification of hexanoic acid (1) with 2-(N,N-dimethylamino)ethanol (2a).
APA, Harvard, Vancouver, ISO, and other styles
10

Rabee, Abdallah, Gamal Mekhemer, Amin Osatiashtiani, Mark Isaacs, Adam Lee, Karen Wilson, and Mohamed Zaki. "Acidity-Reactivity Relationships in Catalytic Esterification over Ammonium Sulfate-Derived Sulfated Zirconia." Catalysts 7, no. 7 (July 5, 2017): 204. http://dx.doi.org/10.3390/catal7070204.

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

Dissertations / Theses on the topic "Esterification Reactivity"

1

Eberhardt, Nathan A. "Synthesis and Reactivity of Nickel POCOP Pincer Complexes for the Reduction of Carbon Dioxide and Related Compounds." University of Cincinnati / OhioLINK, 2017. http://rave.ohiolink.edu/etdc/view?acc_num=ucin1511864001952451.

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

Legoupy, Stéphanie. "Synthèse et réactivité de nouveaux complexes organométalliques chiraux du rhénium." Rennes 1, 1997. http://www.theses.fr/1997REN10148.

Full text
Abstract:
Le travail présenté dans ce mémoire concerne la synthèse et la réactivité de complexes du rhénium. De nouveaux complexes organométalliques chiraux du rhénium des alcools propargyliques et homoallylique ont été synthétisés. Des alcools allyliques secondaires et 1,2-disubstitues ont été coordonnés à l'entité chirale (#5C#5H#5)Re(No)(Pph#3)#+Bf#4#-. Dans le cas du 3-buten-2-ol complexe, les deux diastéréoisomères ont pu été séparés. L'étude de la réactivité de ces complexes du rhénium a montré qu'ils sont compatibles avec des réactions d'oxydation, de Wittig, de réduction, d'estérification, de chloration, de bromation et de fluoration. L'entité organométallique (#5C#5H#5)Re(No)(Pph#3)#+Bf#4#- s'est montrée un bon groupement protecteur d'une seule double liaison au cours de ces réactions. Les substitutions allyliques, catalysées par un acide de Lewis, sur les complexes du rhénium des alcools allyiques ont été étudiées. Quelque soit le nucléophile, ces réactions sont régio- et stéréosélectives et se font avec rétention de configuration. Un mécanisme impliquant un complexe -allyl dicationique du rhénium a été proposé. Le rôle activateur du rhénium a été mis en évidence.
APA, Harvard, Vancouver, ISO, and other styles
3

Yang, Chiehju, and 楊捷如. "Preparation of Sulfated Magnetic Solid Acid Catalyst (SO42−/ Sr7Fe10O22) with Highly Catalytic Reactivity for the Esterification of Oleic Acid and Methanol." Thesis, 2013. http://ndltd.ncl.edu.tw/handle/bvpcdu.

Full text
Abstract:
碩士
靜宜大學
化粧品科學系
101
Transesterification and esterification are the main routes for obtaining esters from plant oils and fatty acids, respectively. Nowadays, these two reactions are catalyzed by homogeneous catalysts (inorganic base or acid) in industry. However, the use of homogeneous catalyst in ester production has serious drawbacks associated with separation, catalyst regeneration and acid/base contamination in the product. Solid catalysts are the promising alternative to homogeneous catalysts. In this work, we focused on the development of solid acid catalyst for the production of esters because the application of solid base catalysts in the production of ester is limited by their low tolerance for the fatty acids containing in natural plant oils. A series of SO42−/M-Fe2O3 (M; doped metal ion in magnetic particle) catalysts were synthesized, and their catalytic abilities were examined in the esterification of oleic acid and methanol at 100°C. The desired metal ions (M) were incorporated in the magnetic support by co-precipitation method. Sulfite groups are then grafted onto the support surface by immersing the support into H2SO4 aqueous solution at a moderate temperature. The physiochemical properties of SO42−/M-Fe2O3 catalysts were characterized by Scanning Electron Microscopy(SEM), X-ray Diffraction(XRD), Brunauer-Emmett-Teller method(BET), Fourier-Transform Infrared Spectrometer(FT-IR) and Hammett Indicators, separately. The results show that the catalysts feature extremely high catalytic reactivity for the esterification in comparison with homogeneous acid catalysts or the solid catalysts reported in the literature. When the esterification of oleic acid with methanol (molar ratio 1:4) was conducted at 100oC in the presence of 10 wt% SO42−/ Sr7Fe10O22, the conversion of oleic acid was over than 95% after 2 hour of the reaction. Heat-treating the used catalyst at 400oC for 1 hours is helpful to recover 85% activity of SO42−/Sr7Fe10O22, and the IR spectrum of used SO42−/Sr7Fe10O22 indicated that the inactivation of SO42−/Sr7Fe10O22, mainly derived from the adsorption of product (oleic acid methyl ester) or reagent (oleic acid) on the active sites.
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Esterification Reactivity"

1

Frey, Perry A., and Adrian D. Hegeman. "Alkyltransferases." In Enzymatic Reaction Mechanisms. Oxford University Press, 2007. http://dx.doi.org/10.1093/oso/9780195122589.003.0019.

Full text
Abstract:
A number of enzymes catalyze alkylation reactions, most of which are reactions of S-adenosyl-L-methionine (SAM) as a methylating agent in the biosynthesis of hormones, modification of DNA, and methyl esterification of proteins involved in signal transduction. Other examples of enzymatic alkylation include prenyl transfer reactions, adenosyltransfer from ATP to methionine in the biosynthesis of SAM, and adenosyltransfer from ATP to cob(I)alamin in the biosynthesis of adenosylcobalamin. Methyl group transfer is also the essential step in the reaction of methionine synthase, which uses 5-methyltetrahydrofolate as an alkylating agent. In an analogous reaction, an analog of 5-methyltetrahydrofolate is the methyl group donor in the methylation of coenzyme M to form methyl coenzyme M, the proximate precursor of methane in methanogenesis (see chap. 4). Glysosyl transfer is an alkylation reaction catalyzed by a large class of enzymes, the glycosyltransferases and glycosidases. The special nature of the glycosyl compounds and their potential for undergoing glycosyltransfer places them in their own class in biochemistry (see chap. 12). The reactivity of glycosyl compounds can be attributed to the contribution of the oxygen atom directly bonded to the glycosyl carbon, the locus of alkylation. In this chapter, we consider other enzymatic alkylations. Alkylation consists of the transfer of a carbon from a leaving group to a nucleophilic acceptor, as in eq.15-1, where R is H or an organic group. The rate is controlled by the reactivity of the nucleophile X:, the stability of the leaving group Y:, and the electrophilic reactivity of the central carbon atom. Alkylation may be regarded as one of the simplest organic chemical reactions because there are few complications in the mechanism. It is the reaction of a nucleophilic molecule with an electrophilic molecule to displace a leaving group. Enzymatic alkylations proceed by polar and not radical mechanisms. In organic chemistry, polar alkylation can occur either by an associative or one-step mechanism, as in fig. 15-1A, or by a dissociative or two-step mechanism through a carbocationic intermediate, as in fig. 15-1B. The chemical nature of the alkylating agent, the propensity of the leaving group to leave, and the polarity of the solvent determine the mechanism.
APA, Harvard, Vancouver, ISO, and other styles

Conference papers on the topic "Esterification Reactivity"

1

John, George, Jose James, Malick Samateh, Siddharth Marwaha, and Vikas Nanda. "Sucralose Hydrogels: Peering into the Reactivity of Sucralose versus Sucrose Using Lipase Catalyzed Trans-Esterification." In 2022 AOCS Annual Meeting & Expo. American Oil Chemists' Society (AOCS), 2022. http://dx.doi.org/10.21748/xkza4963.

Full text
Abstract:
Sucralose differs from sucrose only by virtue of having three Cl groups instead of OH groups. Its intriguing features include being noncaloric, noncariogenic, 600 times sweeter than sucrose, stable at high temperatures/acidic pH's, and void of disagreeable aftertastes. These properties are attractive as food additive, one of which is as hydrogel obtainable via the technique of molecular gelation using a sucralose-derived low-molecular weight gelator (LMWG). The process of molecular gelation entails using specially designed lipid-like amphiphilic molecules capable of self-assembling in a liquid solvent to form a 3D-network. A rational molecular design would involve appending lipophilic alkyl chain to sucralose to afford sucralose-based amphiphiles. Our preliminary study has shown that sucralose, unlike sucrose, is unreactive under biocatalytic conditions using lipase enzyme, which is consistent with its reported lack of reactivity by hydrolytic enzymes in the body. Hence, the aim of this work was (i) to use computation and simulations to further understand sucralose's lack of enzymatic reactivity and (ii) to synthesize the sucralose-based amphiphiles using conventional chemical synthesis and systematically study their tendency towards hydrogelation. Three of the sucralose-based amphiphiles (SL-5, SL-6 and SL-7) proved to be successful hydrogelators. The gelators also showed the ability to gel selected beverages. The LMWGs gelled quantities of water and beverage up to 71 and 55 times their weight, respectively, and remain thermally stable up to 144 °C.
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