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

Journal articles on the topic 'Naphthoic 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 'Naphthoic 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

Meckenstock, Rainer U., Eva Annweiler, Walter Michaelis, Hans H. Richnow, and Bernhard Schink. "Anaerobic Naphthalene Degradation by a Sulfate-Reducing Enrichment Culture." Applied and Environmental Microbiology 66, no. 7 (July 1, 2000): 2743–47. http://dx.doi.org/10.1128/aem.66.7.2743-2747.2000.

Full text
Abstract:
ABSTRACT Anaerobic naphthalene degradation by a sulfate-reducing enrichment culture was studied by substrate utilization tests and identification of metabolites by gas chromatography-mass spectrometry. In substrate utilization tests, the culture was able to oxidize naphthalene, 2-methylnaphthalene, 1- and 2-naphthoic acids, phenylacetic acid, benzoic acid, cyclohexanecarboxylic acid, and cyclohex-1-ene-carboxylic acid with sulfate as the electron acceptor. Neither hydroxylated 1- or 2-naphthoic acid derivatives and 1- or 2-naphthol nor the monoaromatic compounds ortho-phthalic acid, 2-carboxy-1-phenylacetic acid, and salicylic acid were utilized by the culture within 100 days. 2-Naphthoic acid accumulated in all naphthalene-grown cultures. Reduced 2-naphthoic acid derivatives could be identified by comparison of mass spectra and coelution with commercial reference compounds such as 1,2,3,4-tetrahydro-2-naphthoic acid and chemically synthesized decahydro-2-naphthoic acid. 5,6,7,8-Tetrahydro-2-naphthoic acid and octahydro-2-naphthoic acid were tentatively identified by their mass spectra. The metabolites identified suggest a stepwise reduction of the aromatic ring system before ring cleavage. In degradation experiments with [1-13C]naphthalene or deuterated D8-naphthalene, all metabolites mentioned derived from the introduced labeled naphthalene. When a [13C]bicarbonate-buffered growth medium was used in conjunction with unlabeled naphthalene, 13C incorporation into the carboxylic group of 2-naphthoic acid was shown, indicating that activation of naphthalene by carboxylation was the initial degradation step. No ring fission products were identified.
APA, Harvard, Vancouver, ISO, and other styles
2

Song, Yuan Jun, Chun Zhang, and Jing Jing Li. "The Purification and Analysis of 6 - Hydroxy -2 - Naphthoic Acid." Advanced Materials Research 602-604 (December 2012): 1391–95. http://dx.doi.org/10.4028/www.scientific.net/amr.602-604.1391.

Full text
Abstract:
When synthetizing 2 - hydroxy - 6 - naphthoic acid with 2 - naphthol as raw material , we use a variety of methods to characterize and analyze the intermediates and target product . Through melting point method, the melting point of HNA we synthetized is showed to be slightly lower than standard samples. High performance liquid chromatographic shows that the purity of 2 - methoxy - 6 - naphthoic acid 2 - hydroxy - 6 - naphthoic acid reach up to 81.98% and 99.0% respectively. The purity of 2-hydroxyl-6-naphthoic acid is showed to be 97.1% with ultraviolet spectrophotometry.
APA, Harvard, Vancouver, ISO, and other styles
3

Ptáček, Aleš, and Jiří Kulič. "Effect of OH- Concentration on Alkaline Hydrolysis of Diphenyl (4-Nitrophenyl) Phosphate Catalyzed by 2-Iodosobenzoic and 3-Iodoso-2-naphthoic Acids." Collection of Czechoslovak Chemical Communications 59, no. 5 (1994): 1137–44. http://dx.doi.org/10.1135/cccc19941137.

Full text
Abstract:
The hydrolysis of diphenyl (4-nitrophenyl) phosphate catalyzed by 2-iodosobenzoic and 3-iodoso-2-naphthoic acids has been studied at different pH values in the presence of hexadecyltrimethylammonium bromide as a micellar agent. It was found that 3-iodoso-2-naphthoic acid is better catalyst than 2-iodosobenzoic acid. At amounts of the acids higher than stoichiometric, the reaction is independent of pH in the 8.00 to 10.00 region while on using substoichiometric amounts, the reaction rate depends on OH- ion concentration only when the acid to diphenyl (4-nitrophenyl) phosphate molar ratio amounts to 12.5 : 1 for 2-iodosobenzoic acid and 6.25 : 1 for 3-iodoso-2-naphthoic acid.
APA, Harvard, Vancouver, ISO, and other styles
4

Fitzgerald, L. J., and R. E. Gerkin. "Redetermination of the structures of 1-naphthoic acid and 2-naphthoic acid." Acta Crystallographica Section C Crystal Structure Communications 49, no. 11 (November 15, 1993): 1952–58. http://dx.doi.org/10.1107/s0108270193002641.

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

Pařík, Patrik, and Miroslav Ludwig. "Acid-Base Properties of Substituted Naphthoic Acids in Nonaqueous Media." Collection of Czechoslovak Chemical Communications 62, no. 11 (1997): 1737–46. http://dx.doi.org/10.1135/cccc19971737.

Full text
Abstract:
Thirteen substituted 1-naphthoic acids have been prepared and their dissociation constants, along with those of twenty-five substituted 2-naphthoic acids, have been measured potentiometrically in methanol, N,N-dimethylformamide, pyridine, and acetonitrile. The pKHA values obtained have been treated by linear regression using four sets of substituent constants. The experimental data have also been interpreted by statistical methods using latent variables. The first latent variable calculated by these methods can be used as a new set of substituent constants for describing substituent effects in naphthalene skeleton.
APA, Harvard, Vancouver, ISO, and other styles
6

Blackburn, A. C., L. J. Fitzgerald, and R. E. Gerkin. "2-Naphthoic Acid at 153K." Acta Crystallographica Section C Crystal Structure Communications 52, no. 11 (November 15, 1996): 2862–64. http://dx.doi.org/10.1107/s0108270196008876.

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

Souza, Bruno S., Ramon Vitto, Faruk Nome, Anthony J. Kirby, and Adailton J. Bortoluzzi. "3-Acetoxy-2-naphthoic acid." Acta Crystallographica Section E Structure Reports Online 66, no. 11 (October 20, 2010): o2848. http://dx.doi.org/10.1107/s1600536810040365.

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

Blackburn, A. C., and R. E. Gerkin. "1,4-Dimethoxy-2-naphthoic Acid." Acta Crystallographica Section C Crystal Structure Communications 53, no. 10 (October 15, 1997): 1425–27. http://dx.doi.org/10.1107/s0108270197008299.

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

Song, Yun-Sung, and Soon W. Lee. "6-Nicotinamido-2-naphthoic acid." Acta Crystallographica Section E Structure Reports Online 68, no. 7 (June 2, 2012): o1978. http://dx.doi.org/10.1107/s1600536812024051.

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

Song, Yuan Jun, Yan Wei Li, Bao Qing Pan, and Hao Nan Song. "Synthesis of 2-Acetoxy-6-Naphthoic Acid with P-Methylbenzenesulfonic Acid as Catalyst." Applied Mechanics and Materials 152-154 (January 2012): 306–11. http://dx.doi.org/10.4028/www.scientific.net/amm.152-154.306.

Full text
Abstract:
2-acetoxy-6-naphthoic acid (ANA) was synthesized in the presence of 2-hydroxy -6-naphthoic acid (HNA) and acetic anhydride (CH3CO)2O with toluenesulfonic acid (PTSA) as catalyst. The effects of reactant ratio, temperature, time were investigated in the acetylation process. 1H-NMR、13C-NMR、FT-IR、HPLC measurements indicate the optimal acetylation reaction condition when the ratio of HNA: (CH3CO)2O : PTSA are 1: 2.3:0.025 under the temperature at 90~95°C for 40min.
APA, Harvard, Vancouver, ISO, and other styles
11

FITZGERALD, L. J., and R. E. GERKIN. "ChemInform Abstract: Redetermination of the Structures of 1-Naphthoic Acid and 2-Naphthoic Acid." ChemInform 25, no. 7 (August 19, 2010): no. http://dx.doi.org/10.1002/chin.199407036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Annweiler, Eva, Arne Materna, Michael Safinowski, Andreas Kappler, Hans H. Richnow, Walter Michaelis, and Rainer U. Meckenstock. "Anaerobic Degradation of 2-Methylnaphthalene by a Sulfate-Reducing Enrichment Culture." Applied and Environmental Microbiology 66, no. 12 (December 1, 2000): 5329–33. http://dx.doi.org/10.1128/aem.66.12.5329-5333.2000.

Full text
Abstract:
ABSTRACT Anaerobic degradation of 2-methylnaphthalene was investigated with a sulfate-reducing enrichment culture. Metabolite analyses revealed two groups of degradation products. The first group comprised two succinic acid adducts which were identified as naphthyl-2-methyl-succinic acid and naphthyl-2-methylene-succinic acid by comparison with chemically synthesized reference compounds. Naphthyl-2-methyl-succinic acid accumulated to 0.5 μM in culture supernatants. Production of naphthyl-2-methyl-succinic acid was analyzed in enzyme assays with dense cell suspensions. The conversion of 2-methylnaphthalene to naphthyl-2-methyl-succinic acid was detected at a specific activity of 0.020 � 0.003 nmol min−1 mg of protein−1 only in the presence of cells and fumarate. We conclude that under anaerobic conditions 2-methylnaphthalene is activated by fumarate addition to the methyl group, as is the case in anaerobic toluene degradation. The second group of metabolites comprised 2-naphthoic acid and reduced 2-naphthoic acid derivatives, including 5,6,7,8-tetrahydro-2-naphthoic acid, octahydro-2-naphthoic acid, and decahydro-2-naphthoic acid. These compounds were also identified in an earlier study as products of anaerobic naphthalene degradation with the same enrichment culture. A pathway for anaerobic degradation of 2-methylnaphthalene analogous to that for anaerobic toluene degradation is proposed.
APA, Harvard, Vancouver, ISO, and other styles
13

Kricheldorf, Hans R., and Thorsten Adebahr. "Whiskers 9. Synthesis of whisker-like crystals of poly(6-hydroxy-2-naphthoic acid)." High Performance Polymers 6, no. 2 (April 1994): 109–21. http://dx.doi.org/10.1177/095400839400600202.

Full text
Abstract:
Poly(6-hydroxy-2-naphthoic acid), poly(2,6-HNA), was synthesized by three different methods: polycondensation of 6-acetoxy-2-naphthoic acid, or its trimethylsilyl ester, and polycondensation of free 6-hydroxy-2-naphthoic acid with acetic arnhydride in an 'one-pot procedure'. The last method gave whiskers only when a monomer of high purity was used. Different temperatures (350 or 400C) and reaction media tMarlotherm-S, Marlotherm-SCB or Santotherm) were applied. Furthermore, the reaction time and monomer concentration were varied. 6-Acetoxy-2-naphthoic acid yielded whiskers only at concentrations < 0.1 mol I1', whereas its trimethylsilyl ester gave whiskers even at a concentration of 0.4 mol 11. The crystal growth of whisker-like crystals proved to be sensitive to the presence of radical scavengers or acidic transesterification catalysts.
APA, Harvard, Vancouver, ISO, and other styles
14

Annweiler, Eva, Walter Michaelis, and Rainer U. Meckenstock. "Identical Ring Cleavage Products during Anaerobic Degradation of Naphthalene, 2-Methylnaphthalene, and Tetralin Indicate a New Metabolic Pathway." Applied and Environmental Microbiology 68, no. 2 (February 2002): 852–58. http://dx.doi.org/10.1128/aem.68.2.852-858.2002.

Full text
Abstract:
ABSTRACT Anaerobic degradation of naphthalene, 2-methylnaphthalene, and tetralin (1,2,3,4-tetrahydronaphthalene) was investigated with a sulfate-reducing enrichment culture obtained from a contaminated aquifer. Degradation studies with tetralin revealed 5,6,7,8-tetrahydro-2-naphthoic acid as a major metabolite indicating activation by addition of a C1 unit to tetralin, comparable to the formation of 2-naphthoic acid in anaerobic naphthalene degradation. The activation reaction was specific for the aromatic ring of tetralin; 1,2,3,4-tetrahydro-2-naphthoic acid was not detected. The reduced 2-naphthoic acid derivatives tetrahydro-, octahydro-, and decahydro-2-naphthoic acid were identified consistently in supernatants of cultures grown with either naphthalene, 2-methylnaphthalene, or tetralin. In addition, two common ring cleavage products were identified. Gas chromatography-mass spectrometry (GC-MS) and high-resolution GC-MS analyses revealed a compound with a cyclohexane ring and two carboxylic acid side chains as one of the first ring cleavage products. The elemental composition was C11H16O4 (C11H16O4-diacid), indicating that all carbon atoms of the precursor 2-naphthoic acid structure were preserved in this ring cleavage product. According to the mass spectrum, the side chains could be either an acetic acid and a propenic acid, or a carboxy group and a butenic acid side chain. A further ring cleavage product was identified as 2-carboxycyclohexylacetic acid and was assumed to be formed by β-oxidation of one of the side chains of the C11H16O4-diacid. Stable isotope-labeling growth experiments with either 13C-labeled naphthalene, per-deuterated naphthalene-d 8, or a 13C-bicarbonate-buffered medium showed that the ring cleavage products derived from the introduced carbon source naphthalene. The series of identified metabolites suggests that anaerobic degradation of naphthalenes proceeds via reduction of the aromatic ring system of 2-naphthoic acid to initiate ring cleavage in analogy to the benzoyl-coenzyme A pathway for monoaromatic hydrocarbons. Our findings provide strong indications that further degradation goes through saturated compounds with a cyclohexane ring structure and not through monoaromatic compounds. A metabolic pathway for anaerobic degradation of bicyclic aromatic hydrocarbons with 2-naphthoic acid as the central intermediate is proposed.
APA, Harvard, Vancouver, ISO, and other styles
15

Deck, Lorraine, Jacob Greenberg, Lisa Whalen, David Vander Jagt, and Robert Royer. "Synthesis of Naphthoic Acids as Potential Anticancer Agents." Synlett 30, no. 01 (December 3, 2018): 104–8. http://dx.doi.org/10.1055/s-0037-1611342.

Full text
Abstract:
As part of ongoing research to investigate structural requirements for lactate dehydrogenase inhibition by highly substituted naphthoic acids, nine new aryl-substituted dihydroxynaphthoic acids were synthesized from three known precursors. Described here are efficient preparations of the 1-naphthoic acid target compounds by using Suzuki coupling reactions, formylations, oxidations, and demethylations. Lactate dehydrogenase inhibition studies conducted with five of the compounds revealed values of the inhibitory constant K i in the low micromolar range.
APA, Harvard, Vancouver, ISO, and other styles
16

Stringfellow, William T., and Michael D. Aitken. "Comparative physiology of phenanthrene degradation by two dissimilar pseudomonads isolated from a creosote-contaminated soil." Canadian Journal of Microbiology 40, no. 6 (June 1, 1994): 432–38. http://dx.doi.org/10.1139/m94-071.

Full text
Abstract:
Two species of bacteria, identified as Pseudomonas stutzeri (P-16) and Pseudomonas saccharophila (P-15) by fatty acid methyl ester analysis, were found in a phenanthrene enrichment culture of a creosote-contaminated soil. The organisms are shown to be physiologically dissimilar, and their genetic relatedness is discussed. Phenanthrene degradation by both organisms followed Michaelis–Menten kinetics, allowing for the determination of half-saturation (Ks) and maximum activity coefficients, using nonlinear regression. Both organisms utilized kinetically similar enzymes for phenanthrene uptake and oxidation, as evidenced by similar Ks coefficients of approximately 0.2 mg/L and temperature optima of 40 °C, but levels of expression differed with different media. Each organism degraded phenanthrene via salicylic acid, but patterns of intermediate metabolism were shown to differ. P-15 excreted 1-hydroxy-2-naphthoic acid during growth on phenanthrene and demonstrated Michaelis–Menten kinetics for the oxidation of 1-hydroxy-2-naphthoic acid by resting cells. P-16 excreted only trace amounts of 1-hydroxy-2-naphthoic acid and demonstrated linear kinetics in response to 1-hydroxy-2-naphthoic acid concentration. P-15 was found to form thick biofilms on phenanthrene crystals and was characterized by a hydrophobic cell surface, whereas P-16 grew mostly in suspension and was hydrophilic. Neither organism produced significant amounts of biosurfactants when grown on phenanthrene. The implications of these findings for the design of systems to remediate contaminated soil are discussed.Key words: phenanthrene, 1-hydroxy-2-naphthoic acid, biodegradation, kinetics, polycyclic aromatic hydrocarbons.
APA, Harvard, Vancouver, ISO, and other styles
17

Harp, Bhakti Petigara, and Julie N. Barrows. "Reversed-Phase Liquid Chromatographic Determination of Two Manufacturing Intermediates in D&C Red No. 34 and Its Lakes." Journal of AOAC INTERNATIONAL 92, no. 6 (November 1, 2009): 1639–43. http://dx.doi.org/10.1093/jaoac/92.6.1639.

Full text
Abstract:
Abstract A reversed-phase LC method was developed to determine two manufacturing intermediates in the monosulfo monoazo color additive D&C Red No. 34 and its lakes. The analytes are 2-amino-1-naphthalenesulfonic acid (Tobias acid) and 3-hydroxy-2-naphthalenecarboxylic acid (3-hydroxy-2-naphthoic acid). This method can be used for batch certification of the color additives by the U.S. Food and Drug Administration to ensure that each lot meets published specifications for coloring drugs and cosmetics. The new method uses lithium oxalate in methanolwater to dissolve the color additives for analysis. The analytes were identified by comparison of their LC retention times and UV absorption spectra with those of standards. Peak area calibrations were generally linear (R &gt; 0.999) and recoveries were 105 for Tobias acid and 103 for 3-hydroxy-2-naphthoic acid. The limits of determination (LOD) were 0.01 for Tobias acid and 0.03 for 3-hydroxy-2-naphthoic acid. The RSDs at the specification levels were 0.9 for Tobias acid and 3.2 for 3-hydroxy-2-naphthoic acid. Survey analyses of 14 samples of certified D&C Red No. 34 straight colors and lakes from six domestic and foreign manufacturers yielded results for Tobias acid that generally agreed with results previously obtained by using a gravity elution column chromatographic method. Nine of the results for 3-hydroxy-2-naphthoic acid were 2 to 5 times higher than the results obtained using the column chromatographic method. We attribute the lower accuracy of the column chromatographic method to incomplete solubility of the samples using the method conditions and difficulty with interpreting the UV spectrophotometric results.
APA, Harvard, Vancouver, ISO, and other styles
18

Pande, Archana, and Yogendra N. Shukla. "Naphthoic acid derivative from Valeriana wallichii." Phytochemistry 32, no. 5 (March 1993): 1350–51. http://dx.doi.org/10.1016/s0031-9422(00)95120-3.

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

Madeley, Lee G., Demetrius C. Levendis, and Andreas Lemmerer. "Isonicotinamide–2-naphthoic acid (1/1)." Acta Crystallographica Section E Structure Reports Online 67, no. 12 (November 25, 2011): o3440. http://dx.doi.org/10.1107/s1600536811050057.

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

Thenmozhi, S., J. ArulClement, A. K. MohanaKrishnan, and A. SubbiahPandi. "3-(2,4,6-Trimethylbenzoyl)-2-naphthoic acid." Acta Crystallographica Section E Structure Reports Online 66, no. 11 (October 13, 2010): o2794. http://dx.doi.org/10.1107/s1600536810040183.

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

Liu, Xiao-Gang, and Tao Chen. "6-Amino-2-naphthoic acid monohydrate." Acta Crystallographica Section E Structure Reports Online 63, no. 2 (January 12, 2007): o597—o598. http://dx.doi.org/10.1107/s1600536807000116.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Casassas, Enric, Miquel Esteban, and Santiago Alier. "Pulse polarographic study of the behaviour of some o,o'-dihydroxyazo-compounds." Collection of Czechoslovak Chemical Communications 54, no. 5 (1989): 1219–26. http://dx.doi.org/10.1135/cccc19891219.

Full text
Abstract:
The reduction of several o,o'-dihydroxyazo-compounds is studied by means of pulse polarographic techniques (DPP, NPP and RPP). The compounds studied are the following: 2-(2'-hydroxyphenylazo)-phenol (o,o'-dihydroxyazobenzene), 1-(2'-hydroxy-1'-naphthylazo)-2-naphthol-4-sulphonic acid (calcon or Eriochrome Blue Black R), 1-(2'-hydroxy-4'-sulpho-1'-naphthylazo)-2-hydroxy-3-naphthoic acid (calcon carboxylic acid), and 1-(1'-hydroxy-2'-naphthylazo)-6-nitro-2-naphthol-4-sulphonic acid (Eriochrome Black T). Correlations between Ip and Epand experimental variables (pH, T, conc.) and instrumental parameters (dropping time, t, and pulse magnitude, ΔE) are established. Reaction mechanisms formerly proposed are discussed on the basis of the new obtained results, and the ranges are defined where adsorption and/or acid-base catalysis are operative.
APA, Harvard, Vancouver, ISO, and other styles
23

Deveryshetty, Jaigeeth, and Prashant S. Phale. "Biodegradation of phenanthrene by Pseudomonas sp. strain PPD: purification and characterization of 1-hydroxy-2-naphthoic acid dioxygenase." Microbiology 155, no. 9 (September 1, 2009): 3083–91. http://dx.doi.org/10.1099/mic.0.030460-0.

Full text
Abstract:
Pseudomonas sp. strain PPD can metabolize phenanthrene as the sole source of carbon and energy via the ‘phthalic acid’ route. The key enzyme, 1-hydroxy-2-naphthoic acid dioxygenase (1-HNDO, EC 1.13.11.38), was purified to homogeneity using a 3-hydroxy-2-naphthoic acid (3-H2NA)-affinity matrix. The enzyme was a homotetramer with a native molecular mass of 160 kDa and subunit molecular mass of ∼39 kDa. It required Fe(II) as the cofactor and was specific for 1-hydroxy-2-naphthoic acid (1-H2NA), with K m 13.5 μM and V max 114 μmol min−1 mg−1. 1-HNDO failed to show activity with gentisic acid, salicylic acid and other hydroxynaphthoic acids tested. Interestingly, the enzyme showed substrate inhibition with a K i of 116 μM. 1-HNDO was found to be competitively inhibited by 3-H2NA with a K i of 24 μM. Based on the pH-dependent spectral changes, the enzyme reaction product was identified as 2-carboxybenzalpyruvic acid. Under anaerobic conditions, the enzyme failed to convert 1-H2NA to 2-carboxybenzalpyruvic acid. Stoichiometric studies showed the incorporation of 1 mol O2 into the substrate to yield 1 mol product. These results suggest that 1-HNDO from Pseudomonas sp. strain PPD is an extradiol-type ring-cleaving dioxygenase.
APA, Harvard, Vancouver, ISO, and other styles
24

Wang, Yang, Qiang Li, Xian Feng Gong, Xue Ling Zhao, and Yu Zhang. "Study on Synthesis of 2-Hydroxy-6-Naphthoic Acid from 2-Naphthol and Improve the Synthetic Process." Advanced Materials Research 634-638 (January 2013): 2044–48. http://dx.doi.org/10.4028/www.scientific.net/amr.634-638.2044.

Full text
Abstract:
2-Hydroxy-6-naphthoic acid is an main raw material for the synthesis of polyaromatic ester, it was synthesized from 2-Naphthol by methoxylation, bromization, grignard reaction and demethylation before recrystallization (C2H5OH:H2O=2.8:1 m/m) with an overall yield of 78.5% and content of 99.1%. Furthermore, the results showed that the 6-bromo-2-methoxynaphthalene was prepared using Sn as reducing agent with 96.2% of the yield which is higher than 6-bromo-2-methoxynaphthalene was prepared using solid carbon dioxide and grignard reagent with 23.8% of the yield.
APA, Harvard, Vancouver, ISO, and other styles
25

Al-Farhan, Khalid A. "Triphenylphosphine oxide–1-naphthoic acid (1/1)." Acta Crystallographica Section C Crystal Structure Communications 60, no. 7 (June 30, 2004): o531—o532. http://dx.doi.org/10.1107/s010827010401279x.

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

Schmidt, Martin U., Guido Wagner, and Michael Bolte. "Redetermination of 3-hydroxy-2-naphthoic acid." Acta Crystallographica Section E Structure Reports Online 58, no. 8 (July 25, 2002): o918—o919. http://dx.doi.org/10.1107/s1600536802012667.

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

Chowdhury, Piyali Pal, Jayita Sarkar, Soumik Basu, and Tapan K. Dutta. "Metabolism of 2-hydroxy-1-naphthoic acid and naphthalene via gentisic acid by distinctly different sets of enzymes in Burkholderia sp. strain BC1." Microbiology 160, no. 5 (May 1, 2014): 892–902. http://dx.doi.org/10.1099/mic.0.077495-0.

Full text
Abstract:
Burkholderia sp. strain BC1, a soil bacterium, isolated from a naphthalene balls manufacturing waste disposal site, is capable of utilizing 2-hydroxy-1-naphthoic acid (2H1NA) and naphthalene individually as the sole source of carbon and energy. To deduce the pathway for degradation of 2H1NA, metabolites isolated from resting cell culture were identified by a combination of chromatographic and spectrometric analyses. Characterization of metabolic intermediates, oxygen uptake studies and enzyme activities revealed that strain BC1 degrades 2H1NA via 2-naphthol, 1,2,6-trihydroxy-1,2-dihydronaphthalene and gentisic acid. In addition, naphthalene was found to be degraded via 1,2-dihydroxy-1,2-dihydronaphthalene, salicylic acid and gentisic acid, with the putative involvement of the classical nag pathway. Unlike most other Gram-negative bacteria, metabolism of salicylic acid in strain BC1 involves a dual pathway, via gentisic acid and catechol, with the latter being metabolized by catechol 1,2-dioxygenase. Involvement of a non-oxidative decarboxylase in the enzymic transformation of 2H1NA to 2-naphthol indicates an alternative catabolic pathway for the bacterial degradation of hydroxynaphthoic acid. Furthermore, the biochemical observations on the metabolism of structurally similar compounds, naphthalene and 2-naphthol, by similar but different sets of enzymes in strain BC1 were validated by real-time PCR analyses.
APA, Harvard, Vancouver, ISO, and other styles
28

Pal Chowdhury, Piyali, Soumik Basu, Arindam Dutta, and Tapan K. Dutta. "Functional Characterization of a Novel Member of the Amidohydrolase 2 Protein Family, 2-Hydroxy-1-Naphthoic Acid Nonoxidative Decarboxylase from Burkholderia sp. Strain BC1." Journal of Bacteriology 198, no. 12 (April 11, 2016): 1755–63. http://dx.doi.org/10.1128/jb.00250-16.

Full text
Abstract:
ABSTRACTThe gene encoding a nonoxidative decarboxylase capable of catalyzing the transformation of 2-hydroxy-1-naphthoic acid (2H1NA) to 2-naphthol was identified, recombinantly expressed, and purified to homogeneity. The putative gene sequence of the decarboxylase (hndA) encodes a 316-amino-acid protein (HndA) with a predicted molecular mass of 34 kDa. HndA exhibited high identity with uncharacterized amidohydrolase 2 proteins of variousBurkholderiaspecies, whereas it showed a modest 27% identity with γ-resorcylate decarboxylase, a well-characterized nonoxidative decarboxylase belonging to the amidohydrolase superfamily. Biochemically characterized HndA demonstrated strict substrate specificity toward 2H1NA, whereas inhibition studies with HndA indicated the presence of zinc as the transition metal center, as confirmed by atomic absorption spectroscopy. A three-dimensional structural model of HndA, followed by docking analysis, identified the conserved metal-coordinating and substrate-binding residues, while their importance in catalysis was validated by site-directed mutagenesis.IMPORTANCEMicrobial nonoxidative decarboxylases play a crucial role in the metabolism of a large array of carboxy aromatic chemicals released into the environment from a variety of natural and anthropogenic sources. Among these, hydroxynaphthoic acids are usually encountered as pathway intermediates in the bacterial degradation of polycyclic aromatic hydrocarbons. The present study reveals biochemical and molecular characterization of a 2-hydroxy-1-naphthoic acid nonoxidative decarboxylase involved in an alternative metabolic pathway which can be classified as a member of the small repertoire of nonoxidative decarboxylases belonging to the amidohydrolase 2 family of proteins. The strict substrate specificity and sequence uniqueness make it a novel member of the metallo-dependent hydrolase superfamily.
APA, Harvard, Vancouver, ISO, and other styles
29

Kricheldorf, Hans R., and Andreas Gerken. "New Polymer Syntheses 87. Thermosetting Nematic or Cholesteric Diesters Having Propargyl Endgroups." High Performance Polymers 9, no. 2 (June 1997): 75–90. http://dx.doi.org/10.1088/0954-0083/9/2/001.

Full text
Abstract:
4-Propargyloxy benzoic acid, 6-propargyloxy-2-naphthoic acid and 4′-propargyloxybiphenyl-4-carboxylic acid were synthesized from the corresponding hydroxy acids and converted into the acid chlorides. Diesters were then prepared by the reaction of these acid chlorides with diphenols such as hydroquinone, methylhydroquinone n-octylthiohydroquinone and phenylhydroquinone. Most of these diesters form an enantiotropic nematic melt but some diesters form a monotropic liquid crystal phase. Cholesteric phases were observed for diesters of ( S)-2-methylbutylthiohydroquinone. Diesters derived from 6-propargyloxy-2-naphthoic acid begin to crosslink above 200 °C with an optimum cure temperature of 240–250 °C. The crosslinking of the 4′-propargyloxybiphenyl-4-carboxylic acid derivatives occurs at a temperature 60 °C higher. Several diesters crosslink in the nematic state, but most of them undergo isotropization before completion of the crosslinking process. Numerous cholesteric blends of nematic diesters and chiral diesters of isosorbide were studied and Grandjean textures with bright colours were found. Thermal crosslinking in the cholesteric phase was feasible in several cases.
APA, Harvard, Vancouver, ISO, and other styles
30

Flakus, Henryk T., and Michał Chełmecki. "Polarization IR spectra of hydrogen bonded 1-naphthoic acid and 2-naphthoic acid crystals: electronic effects in the spectra." Journal of Molecular Structure 659, no. 1-3 (October 2003): 103–17. http://dx.doi.org/10.1016/j.molstruc.2003.08.007.

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

Khokhar, Vaishali, and Siddharth Pandey. "Prototropic forms of hydroxy derivatives of naphthoic acid within deep eutectic solvents." Physical Chemistry Chemical Physics 23, no. 15 (2021): 9096–108. http://dx.doi.org/10.1039/d1cp00845e.

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

Arlin, Jean-Baptiste, and Alan R. Kennedy. "Alkaline earth metal salts of 1-naphthoic acid." Acta Crystallographica Section C Crystal Structure Communications 68, no. 8 (July 19, 2012): m213—m218. http://dx.doi.org/10.1107/s0108270112030399.

Full text
Abstract:
The structures of the Mg, Ca, Sr and Ba salts of 1-naphthoic acid are examined and compared with analogous structures of salts of benzoate derivatives. It is shown thatcatena-poly[[[diaquabis(1-naphthoato-κO)magnesium(II)]-μ-aqua] dihydrate], {[Mg(C11H7O2)2(H2O)3]·2H2O}n, exists as a one-dimensional coordination polymer that propagates only through Mg—OH2—Mg interactions along the crystallographicbdirection. In contrast with related benzoate salts, the naphthalene systems are large enough to prevent inorganic chain-to-chain interactions, and thus species with inorganic channels rather than layers are formed. The Ca, Sr and Ba salts all have metal centres that lie on a twofold axis (Z′ = 1 \over 2) and all have the common namecatena-poly[[diaquametal(II)]-bis(μ-1-naphthoato)-κ3O,O′:O;κ3O:O,O′], [M(C11H7O2)2(H2O)2]n, whereM= Ca, Sr or Ba. The Ca and Sr salts are essentially isostructural, and all three species form one-dimensional coordination polymers through a carboxylate group that forms threeM—O bonds. The polymeric chains propagatevia c-glide planes and throughMOMO four-membered rings. Again, inorganic channel structures are formed rather than layered structures, and the three structures are similar to those found for Ca and Sr salicylates and other substituted benzoates.
APA, Harvard, Vancouver, ISO, and other styles
33

Ahn, So-Jin, and Yong-Kul Lee. "Factors influencing the formation of 2-hydroxy-6-naphthoic acid from carboxylation of naphthol." Journal of Industrial and Engineering Chemistry 19, no. 6 (November 2013): 2060–63. http://dx.doi.org/10.1016/j.jiec.2013.03.016.

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

Krishnakumar, V., R. Mathammal, and S. Muthunatesan. "Structures and vibrational frequencies of 2-naphthoic acid and 6-bromo-2-naphthoic acid based on density functional theory calculations." Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 70, no. 1 (June 2008): 201–9. http://dx.doi.org/10.1016/j.saa.2007.06.039.

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

Xiang, Tao, and Keith P. Johnston. "Acid-base behavior in supercritical water: β-naphthoic acid-ammonia equilibrium." Journal of Solution Chemistry 26, no. 1 (January 1997): 13–30. http://dx.doi.org/10.1007/bf02439441.

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

Abrahamsson, Carl-Olof, Ulf Ellervik, Johan Eriksson-Bajtner, Mårten Jacobsson, and Katrin Mani. "Xylosylated naphthoic acid–amino acid conjugates for investigation of glycosaminoglycan priming." Carbohydrate Research 343, no. 9 (July 2008): 1473–77. http://dx.doi.org/10.1016/j.carres.2008.04.006.

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

Joshi, Deepti, and T. K. Joshi. "Synthesis and Characterization of Some Trivalent and Tetravalent Derivatives of 2- Hydroxy-1-naphthoic Acid." E-Journal of Chemistry 1, no. 2 (2004): 110–14. http://dx.doi.org/10.1155/2004/405086.

Full text
Abstract:
The trivalent derivatives (B, Al, As, Sb & Fe) and tetravalent derivatives (Si, Ti & Se) of 2-hydroxy-1-naphthoic acid have been prepared by the interaction of their corresponding isopropoxide with letter in different molar ratioviz. 1:3 &1:4 in benzene medium. The prepared compounds generally obtained as coloured solids and amongst them those containing isopropoxy groups were found to be hygroscopic. All these compound were characterized by azeotrope and elemental analysis as well as by IR, PMR and mass spectral measurements. These spectral data have facilitated in elucidating the mode of bonding of the said metals and non-metals in these compounds with 2-hydroxy-1-naphthoic acid.
APA, Harvard, Vancouver, ISO, and other styles
38

Liu, F., F. Duan, K. He, Y. Ma, K. A. Rahn, and Q. Zhang. "An enhanced procedure for measuring organic acids and methyl esters in PM<sub>2.5</sub>." Atmospheric Measurement Techniques Discussions 8, no. 3 (March 4, 2015): 2379–407. http://dx.doi.org/10.5194/amtd-8-2379-2015.

Full text
Abstract:
Abstract. We have developed an enhanced analytical procedure to measure organic acids and methyl esters in fine aerosol with much greater specificity and sensitivity than previously available. This capability is important because of these species and their low concentrations, even in highly polluted atmospheres like Beijing, China. The procedure first separates the acids and esters from the other organic compounds with anion-exchange solid- phase extraction (SPE), then, quantifies them by gas chromatography coupled with mass spectrometry. This allows us to accurately quantify the C4-C11 dicarboxylic and the C8-C30 monocarboxylic acids. Then the acids are separated from the esters on an aminopropyl SPE cartridge, whose weak retention isolates and enriches the acids from esters prevents the fatty acids and dimethyl phthalate from being overestimated. The resulting correlations between the aliphatic acids and fatty acid methyl esters (FAMEs) suggest that FAMEs had sources similar to those of the carboxylic acids, or were formed by esterifying carboxylic acids, or that aliphatic acids were formed by hydrolyzing FAMEs. In all, 17 aromatic acids were identified and quantified using this procedure coupled with gas chromatography-tandem mass spectrometry, including the five polycyclic aromatic hydrocarbon (PAH) acids 2-naphthoic, biphenyl-4-carboxylic, 9-oxo-9H-fluorene-1-carboxylic, biphenyl-4,4´-dicarboxylic, and phenanthrene-1-carboxylic acid, plus 1,8-naphthalic anhydride. Correlations between the PAH-acids and the dicarboxylic and aromatic acids indicated that the first three acids and 1,8-naphthalic anhydride were mainly secondary, the last two mainly primary.
APA, Harvard, Vancouver, ISO, and other styles
39

Gupta, Akul Sen, Kamaldeep Paul, and Vijay Luxami. "A new ‘Turn-on’ PET-CHEF based fluorescent sensor for Al3+ and CN− ions: applications in real samples." Analytical Methods 10, no. 9 (2018): 983–90. http://dx.doi.org/10.1039/c7ay02779f.

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

Furuya, Toshiki, Yuka Arai, and Kuniki Kino. "Biotechnological Production of Caffeic Acid by Bacterial Cytochrome P450 CYP199A2." Applied and Environmental Microbiology 78, no. 17 (June 22, 2012): 6087–94. http://dx.doi.org/10.1128/aem.01103-12.

Full text
Abstract:
ABSTRACTCaffeic acid is a biologically active molecule that has various beneficial properties, including antioxidant, anticancer, and anti-inflammatory activities. In this study, we explored the catalytic potential of a bacterial cytochrome P450, CYP199A2, for the biotechnological production of caffeic acid. When the CYP199A2 enzyme was reacted withp-coumaric acid, it stoichiometrically produced caffeic acid. The crystal structure of CYP199A2 shows that Phe at position 185 is situated directly above, and only 6.35 Å from, the heme iron. This F185 residue was replaced with hydrophobic or hydroxylated amino acids using site-directed mutagenesis to create mutants with novel and improved catalytic properties. In whole-cell assays with the known substrate of CYP199A2, 2-naphthoic acid, only the wild-type enzyme hydroxylated 2-naphthoic acid at the C-7 and C-8 positions, whereas all of the active F185 mutants exhibited a preference for C-5 hydroxylation. Interestingly, several F185 mutants (F185V, F185L, F185I, F185G, and F185A mutants) also acquired the ability to hydroxylate cinnamic acid, which was not hydroxylated by the wild-type enzyme. These results demonstrate that F185 is an important residue that controls the regioselectivity and the substrate specificity of CYP199A2. Furthermore,Escherichia colicells expressing the F185L mutant exhibited 5.5 times higher hydroxylation activity forp-coumaric acid than those expressing the wild-type enzyme. By using the F185L whole-cell catalyst, the production of caffeic acid reached 15 mM (2.8 g/liter), which is the highest level so far attained in biotechnological production of this compound.
APA, Harvard, Vancouver, ISO, and other styles
41

Mane, Ramchandra Bhimrao, and Abhijit Jaysingrao Kadam. "A New Synthesis of Occidol." Collection of Czechoslovak Chemical Communications 64, no. 3 (1999): 533–38. http://dx.doi.org/10.1135/cccc19990533.

Full text
Abstract:
Sodium borohydride reduction of 5,8-dimethyl-3,4-dihydronaphthalen-1-(2H)-one (4) yielded 5,8-dimethyl-1,2,3,4-tetrahydro-1-naphthol (5). The tetralol 5 on Vilsmeier-Haack reaction with N,N-dimethylacetamide yielded 1-(5,8-dimethyl-3,4-dihydro-2-naphthyl)ethan-1-one (7) which on hydrogenation over Pd/C afforded 1-(5,8-dimethyl-1,2,3,4-tetrahydro-2-naphthyl)ethan-1-one (8). The tetralol 5 on Vilsmeier-Haack formylation gave 5,8-dimethyl-3,4-dihydro-2-naphthaldehyde (9) which on reduction with lithium aluminium hydride followed by oxidation with the Jones reagent furnished 5,8-dimethyl-1,2,3,4-tetrahydro-2-naphthoic acid (11). The acid 11 on treatment with excess of methyllithium yielded (±)-occidol (1); with two moles of methyllithium it yielded ketone 8, which on reaction with methyllithium furnished (±)-occidol (1).
APA, Harvard, Vancouver, ISO, and other styles
42

Liu, Xiaochun, Haifeng Zhao, Linfeng Wei, Xinjian Ren, Xinyang Zhang, Faming Li, Peng Zeng, and Mingzhen Liu. "Ligand-modulated electron transfer rates from CsPbBr3 nanocrystals to titanium dioxide." Nanophotonics 10, no. 8 (June 1, 2020): 1967–75. http://dx.doi.org/10.1515/nanoph-2020-0631.

Full text
Abstract:
Abstract In most perovskite nanocrystal (PeNC)-based optoelectronic and photonic applications, surface ligands inevitably lead to a donor–bridge–acceptor charge transfer configuration. In this article, we demonstrate successful modulation of electron transfer (ET) rates from all-inorganic CsPbBr3 PeNCs to mesoporous titanium dioxide films, by using different surface ligands including single alkyl chain oleic acid and oleylamine, cross-linked insulating (3-aminopropyl)triethoxysilane and aromatic naphthoic acid molecules as the ligand-bridge. We systematically investigated the ET process through time-resolved photoluminescence spectroscopy. Calculations verified the ligand-bridge barrier effect of the three species upon the ET process. Transient absorption measurements excluded carrier-delocalization effect of the naphthoic acid ligands and confirmed the bridge-barrier effect. Our work provides a perspective for composable and appropriate ligands design for diverse practical purposes.
APA, Harvard, Vancouver, ISO, and other styles
43

Liu, F., F. K. Duan, K. B. He, Y. L. Ma, K. A. Rahn, and Q. Zhang. "An enhanced procedure for measuring organic acids and methyl esters in PM<sub>2.5</sub>." Atmospheric Measurement Techniques 8, no. 11 (November 19, 2015): 4851–62. http://dx.doi.org/10.5194/amt-8-4851-2015.

Full text
Abstract:
Abstract. A solid-phase extraction (SPE) pretreatment procedure allowing organic acids to be separated from methyl esters in fine aerosol has been developed. The procedure first separates the organic acids from fatty acid methyl esters (FAMEs) and other nonacid organic compounds by aminopropyl-based SPE cartridge and then quantifies them by gas chromatography/mass spectrometry. The procedure prevents the fatty acids and dimethyl phthalate from being overestimated, and so allows us to accurately quantify the C4–C11 dicarboxylic acids (DCAs) and the C8–C30 monocarboxylic acids (MCAs). Results for the extraction of DCAs, MCAs, and AMAs in eluate and FAMEs in effluate by SAX and NH2 SPE cartridges exhibited that the NH2 SPE cartridge gave higher extraction efficiency than the SAX cartridge. The recoveries of analytes ranged from 67.5 to 111.3 %, and the RSD ranged from 0.7 to 10.9 %. The resulting correlations between the aliphatic acids and FAMEs suggest that the FAMEs had sources similar to those of the carboxylic acids, or were formed by esterifying carboxylic acids, or that aliphatic acids were formed by hydrolyzing FAMEs. Through extraction and cleanup using this procedure, 17 aromatic acids in eluate were identified and quantified by gas chromatography/tandem mass spectrometry, including five polycyclic aromatic hydrocarbon (PAH): acids 2-naphthoic, biphenyl-4-carboxylic, 9-oxo-9H-fluorene-1-carboxylic, biphenyl-4,4´-dicarboxylic, and phenanthrene-1-carboxylic acid, plus 1,8-naphthalic anhydride. Correlations between the PAH acids and the dicarboxylic and aromatic acids suggested that the first three acids and 1,8-naphthalic anhydride were secondary atmospheric photochemistry products and the last two mainly primary.
APA, Harvard, Vancouver, ISO, and other styles
44

Mishra, Hirdyesh, Hem Chandra Joshi, Hira Ballabh Tripathi, Shruti Maheshwary, Narayanasami Sathyamurthy, Manoranjan Panda, and Jayaraman Chandrasekhar. "Photoinduced proton transfer in 3-hydroxy-2-naphthoic acid." Journal of Photochemistry and Photobiology A: Chemistry 139, no. 1 (February 2001): 23–36. http://dx.doi.org/10.1016/s1010-6030(00)00415-9.

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

Kuhlich, Paul, Robert Göstl, Ramona Metzinger, Christian Piechotta, and Irene Nehls. "3,5,5,6,8,8-Hexamethyl-5,6,7,8-tetrahydro-2-naphthoic acid (AHTN–COOH)." Acta Crystallographica Section E Structure Reports Online 66, no. 10 (September 30, 2010): o2687. http://dx.doi.org/10.1107/s1600536810038572.

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

Kumar, Vinod. "2-Naphthoic acid prototropism within ionic liquid based media." Journal of Molecular Liquids 339 (October 2021): 116831. http://dx.doi.org/10.1016/j.molliq.2021.116831.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Kolkmann, Rainer, and Eckhard Leistner. "4-(2′-Carboxyphenyl)-4-oxobutyryl Coenzyme A Ester, an Intermediate in Vitamin K2 (Menaquinone) Biosynthesis." Zeitschrift für Naturforschung C 42, no. 11-12 (December 1, 1987): 1207–14. http://dx.doi.org/10.1515/znc-1987-11-1212.

Full text
Abstract:
Enzyme preparations from Mycobacterium phlei, Escherichia coli and Galium mollugo cell suspension cultures were incubated in the presence of 4-(2′-carboxyphenyl)-4-oxobutyrate (i.e. o- succinylbenzoic acid. OSB. 1). ATP. coenzyme A and Mg2+. The main product isolated from the incubation mixture was 4-(2′-carboxyphenyl)-4-oxobutyryl coenzyme A ester (2) as determined by comparison with synthetic coenzyme A esters. Synthetic and enzymically formed 4-(2′-car-boxyphenyl)-4-oxobutyryl coenzyme A ester (2) was shown to be enzymically converted to an intermediate in vitamin K2 biosynthesis viz. 1.4-dihydroxy-2′-naphthoic acid (5). The enzymic formation of 2-(3′-Carboxypropionyl)benzoyl coenzyme A ester (3) and 4-(2′-carboxyphenyl)-4-oxobutyryl-di-coenzyme A ester (4) was also observed. They appeared in minor amounts, how­ever. These esters were not convertible to 1.4-dihydroxy-2-naphthoic acid (5).
APA, Harvard, Vancouver, ISO, and other styles
48

Zhang, Min, Zhangyi Fu, Anping Luo, Xingwen Pu, Menglei Wang, Ying Huang, Yudong Yang, and Jingsong You. "Palladium-catalyzed C8–H arylation and annulation of 1-naphthalene carboxylic acid derivatives with aryl iodides." Organic Chemistry Frontiers 8, no. 13 (2021): 3274–79. http://dx.doi.org/10.1039/d1qo00428j.

Full text
Abstract:
Disclosed herein is palladium-catalyzed C8–H arylation and annulation of 1-naphthoic acid derivatives with aryl iodides in a low reactant molar ratio via an electrophilic aromatic substitution (SEAr) process.
APA, Harvard, Vancouver, ISO, and other styles
49

Pang, Yanyan, Peiqi Xing, Xiujuan Geng, Yujing Zhu, Faqian Liu, and Lei Wang. "Supramolecular assemblies of 2-hydroxy-3-naphthoic acid and N-heterocycles via various strong hydrogen bonds and weak X⋯π (X = C–H, π) interactions." RSC Advances 5, no. 51 (2015): 40912–23. http://dx.doi.org/10.1039/c5ra03837e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Abdel-Kader, Nora S., Samir A. Abdel-Latif, Aida L. El-Ansary, and Amira G. Sayed. "Combined experimental, DFT theoretical calculations and biological activity of sulfaclozine azo dye with 1-hydroxy-2-naphthoic acid and its complexes with some metal ions." New Journal of Chemistry 43, no. 44 (2019): 17466–85. http://dx.doi.org/10.1039/c9nj04594e.

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