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

Journal articles on the topic 'Transfer reaction'

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 'Transfer reaction.'

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

Rasaiah, Jayendran C., and Jianjun Zhu. "Reaction coordinates for electron transfer reactions." Journal of Chemical Physics 129, no. 21 (December 7, 2008): 214503. http://dx.doi.org/10.1063/1.3026365.

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

KIMURA, Yoshifumi, Koji OSAWA, and Issei KOBAYASHI. "Electron Transfer Reaction and Proton Transfer Reaction in Supercritical Water." Review of High Pressure Science and Technology 23, no. 4 (2013): 300–308. http://dx.doi.org/10.4131/jshpreview.23.300.

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

Ashok, Konda, Prashant V. Kamat, and Manapurathu V. George. "Electron transfer reactions. Reaction of dibenzobarrelenes with potassium." Research on Chemical Intermediates 13, no. 3 (September 1990): 203–20. http://dx.doi.org/10.1163/156856790x00094.

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

Ashok, Konda, Pallikkaparambil M. Scaria, Prashant V. Kamat, and Manapurathu V. George. "Electron transfer reactions. Reaction of nitrones with potassium." Canadian Journal of Chemistry 65, no. 9 (September 1, 1987): 2039–49. http://dx.doi.org/10.1139/v87-339.

Full text
Abstract:
Treatment of nitrones (1a–d, 26a,b, 34, 49) with potassium in tetrahydrofuran (THF) gives rise to radical anion (2a–d, 27a,b, 35, 50) and dianion intermediates (3a–d, 28a,b, 36) through electron transfer reactions. These intermediates undergo further transformations to give a variety of products. Thus, the aldehydonitrones (1a–d) give the corresponding aldehydes (10a–c), carboxylic acids (25a–c), and azobenzenes (19a,d), whereas the ketonitrones (26a,b) give deoxygenation products (31a,b). The nitrone 34 gave a mixture of products consisting of benzoic acid (25a), dibenzyl (48), the dimeric product 38, and tetraphenylpyrazine (46), whereas 49 did not give any isolable product. Cyclic voltammetric studies have been employed to measure the reduction potentials for both one-electron and two-electron transfer processes, leading to the radical anions and dianions respectively. These intermediates have been characterized through their electronic spectra and they were quenched by oxygen. Pulse radiolysis of the nitrones 1a–d, 26a,b, 34, and 49 also gave the corresponding radical anions 2a–d, 27a,b, 35, and 50, characterized by their spectra.
APA, Harvard, Vancouver, ISO, and other styles
5

Niki, Katsumi, and Takamasa Sagara. "Electron transfer reaction of electron transfer proteins." Kobunshi 39, no. 11 (1990): 830–33. http://dx.doi.org/10.1295/kobunshi.39.830.

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

Douhal, Abderrazzak, Françoise Lahmani, and Ahmed H. Zewail. "Proton-transfer reaction dynamics." Chemical Physics 207, no. 2-3 (July 1996): 477–98. http://dx.doi.org/10.1016/0301-0104(96)00067-5.

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

Tripathi, R., S. Sodaye, K. Ramachandran, S. K. Sharma, and P. K. Pujari. "Incomplete mass transfer processes in 28Si +93Nb reaction." International Journal of Modern Physics E 27, no. 02 (February 2018): 1850010. http://dx.doi.org/10.1142/s0218301318500106.

Full text
Abstract:
Cross sections of reaction products were measured in [Formula: see text]Si[Formula: see text]Nb reaction using recoil catcher technique involving by off-line gamma-ray spectrometry at beam energies of 105 and 155[Formula: see text]MeV. At [Formula: see text][Formula: see text]MeV, the contribution from different incomplete mass transfer processes is investigated. Results of the present studies show the contribution from deep inelastic collision (DIC), massive transfer or incomplete fusion (ICF) and quasi-elastic transfer (QET). The contribution from massive transfer reactions was confirmed from the fractional yield of the reaction products in the forward catcher foil. The present results are different from those from the reactions with comparatively higher entrance channel mass asymmetry with lighter projectiles, for which dominant transfer processes are ICF and QET which involve mass transfer predominantly from projectile to target. The [Formula: see text] values of the products close to the target mass were observed to be in a wide range, starting from [Formula: see text] of the target ([Formula: see text]Nb) and extending slightly below the [Formula: see text] of the composite system, consistent with the contribution from DIC and QET reactions. At [Formula: see text][Formula: see text]MeV, a small contribution from QET was observed in addition to complete fusion.
APA, Harvard, Vancouver, ISO, and other styles
8

Muneer, Mohammed, Prashant V. Kamat, and Manapurathu V. George. "Electron transfer reactions. Reaction of nitrogen heterocycles with potassium." Canadian Journal of Chemistry 68, no. 6 (June 1, 1990): 969–75. http://dx.doi.org/10.1139/v90-152.

Full text
Abstract:
The results of our studies on potassium-induced transformations of some selected nitrogen heterocycles are presented. The substrates under investigation include 2,3-diphenylindole (1a), 2,3-diphenyl-1-methylindole (1b), 1,2,3-triphenylindole (1c), 2,3,4,5-tetraphenylpyrrole (5a), 1,2,3,5-tetraphenylpyrrole (5b), 1-benzyl-2,3,5-triphenylpyrrole (5c), 2,4,5-triphenyloxazole (15a), 4,5-diphenyl-2-methyloxazole (15b), 2,4-diphenyl-5-methyloxazole (15c), and 2,4,5-triphenylimidazole (19). Treatment of 1a with potassium in THF gave 9H-dibenzo[a,c]carbazole (3a), whereas 1c gave a mixture of 9-phenyl-9H-dibenzo-[a,c]carbazole (3c) and 2,3-diphenylindole (1a). Under identical conditions, 1b gave only the cleavage product 1a. In contrast, when the reactions of 1a,c were carried out with potassium in THF saturated with oxygen, and with potassium superoxide in benzene containing 18-crown-6, a mixture of 2-benzamidobenzophenone (4a), the carbazoles 3a,c, and 1a was formed. Although no product was isolated on treatment of 5a with potassium in THF, the reaction of 5a with potassium in THF saturated with oxygen gave a mixture of tetraphenylpyrazine (7a), the benzoylaminostilbene 8a, the lactam 12a, benzamide (11a), and benzoic acid (14). Similar results were obtained in the reaction of 5a with potassium superoxide. The reaction of N-substituted pyrroles such as 5b,c with potassium gave the NH pyrrole 9b in each case, whereas the reaction of 5b,c with potassium in THF, saturated with oxygen, gave a mixture of 9b, the butanone 10b, the 1,4-dione 13b, the lactam 12b, the amides 11a–c, and benzoic acid (14). Attempted reactions of 5b,c with potassium superoxide did not give any isolable product; most of the starting material could be recovered unchanged in each case. A mixture of N-(1,2-diphenylethyl)benzamide (18a) and benzoic acid (14) was formed in the reaction of the oxazole 15a with potassium, whereas 15b,c, under analogous conditions, gave the N-vinylamides 17b,c and benzoic acid (14). In contrast, treatment of the imidazole 19 with potassium in THF did not give any product; however, when the reaction of 19 was carried out with potassium in THF saturated with oxygen, and with potassium superoxide, dibenzamide (21) was isolated, in each case.Radical ions have been invoked as intermediates in the transformation of the different substrates to the observed products. Cyclic voltammetric studies have been carried out to measure the reduction potentials of these radical anion intermediates. These radical anions have also been generated by pulse radiolysis in methanol, and their absorption spectra recorded. Keywords: nitrogen heterocycles, radical ions, potassium-induced transformations, pulse radiolysis, cyclic voltammetry.
APA, Harvard, Vancouver, ISO, and other styles
9

Yamaguchi, Yasuchika, Noriaki Tatsuta, Kenji Hayakawa, and Ken Kanematsu. "Chirality transfer from the furan ring transfer reaction." Journal of the Chemical Society, Chemical Communications, no. 8 (1989): 470. http://dx.doi.org/10.1039/c39890000470.

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

Arseneau, D. J., D. G. Fleming, M. Senba, I. D. Reid, and D. M. Garner. "The ion–molecule reactivity of the positive muon molecular ions HeMu+ and NeMu+." Canadian Journal of Chemistry 66, no. 8 (August 1, 1988): 2018–24. http://dx.doi.org/10.1139/v88-325.

Full text
Abstract:
Thermal (300 K) ion–molecule reaction rates are measured, using the µSR (muon spin rotation) technique, for the muonated rare gas molecular ions HeMu+ and NeMu+ reacting with NO, O2, N2O, NH3, CF4, C2H4, TMS, and CH3NO2. In almost every case (excepting O2), both charge transfer (ke) and muon transfer (kµ) contribute to the reaction rate. Reaction is believed to occur from ro-vibrational excited states, [HeMu+]* and [NeMu+]*, due to the poor efficiency of He and Ne moderators for collisional deactivation. The total experimental rate constants, kexp = kµ + ke, are generally in excellent agreement with total capture rates predicted by the simple ADO theory, regardless of the degree of internal excitation. Comparisons with literature values for corresponding protonated ion reaction rates with O2 and C2H4 reveal little or no isotope effect, although it is noted that these reactions are dominated by proton transfer, in contrast to the µSR results.
APA, Harvard, Vancouver, ISO, and other styles
11

Szilner, S., L. Corradi, G. Pollarolo, E. Fioretto, A. M. Stefanini, G. [Angelis]de Angelis, J. J. Valiente-Dobón, et al. "Transfer Reaction Studies with Spectrometers." Acta Physica Polonica B 44, no. 3 (2013): 417. http://dx.doi.org/10.5506/aphyspolb.44.417.

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

Blake, Robert S., Paul S. Monks, and Andrew M. Ellis. "Proton-Transfer Reaction Mass Spectrometry." Chemical Reviews 109, no. 3 (March 11, 2009): 861–96. http://dx.doi.org/10.1021/cr800364q.

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

Jiang, Xuefeng. "Sulfur atom transfer (SAT) reaction." Phosphorus, Sulfur, and Silicon and the Related Elements 192, no. 2 (October 24, 2016): 169–71. http://dx.doi.org/10.1080/10426507.2016.1250762.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Kitagawa, Osamu, Akihiko Miura, Yoshiro Kobayashi, and Takeo Taguchi. "Atom-transfer Reaction of Difluoroiodoacetate." Chemistry Letters 19, no. 6 (June 1990): 1011–14. http://dx.doi.org/10.1246/cl.1990.1011.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Sumino, Shuhei, Akira Fusano, and Ilhyong Ryu. "Reductive Bromine Atom-Transfer Reaction." Organic Letters 15, no. 11 (May 22, 2013): 2826–29. http://dx.doi.org/10.1021/ol4011536.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Igarashi, M. "Two-nucleon transfer reaction mechanisms." Physics Reports 199, no. 1 (January 1991): 1–72. http://dx.doi.org/10.1016/0370-1573(91)90078-z.

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

Mijatović, Tea, Suzana Szilner, Lorenzo Corradi, Franco Galtarossa, Samuel Bakes, Daniele Brugnara, Gabriele Carozzi, et al. "Multinucleon transfer reactions and proton transfer channels." EPJ Web of Conferences 223 (2019): 01039. http://dx.doi.org/10.1051/epjconf/201922301039.

Full text
Abstract:
Transfer reactions have always been of great importance for nuclear structure and reaction mechanism studies. So far, in multinucleon transfer studies, proton pickup channels have been completely identified in atomic and mass numbers at energies close to the Coulomb barrier only in few cases. We measured the multinucleon transfer reactions in the 40Ar+208Pb system near the Coulomb barrier, by employing the PRISMA magnetic spectrometer. By using the most neutron-rich stable 40Ar beam we could populate, besidesneutron pickup and proton stripping channels, also neutron stripping and proton pickup channels. Comparison ofcross sections between different systems with the 208Pb target and with projectiles going from neutron-poor to neutron-rich nuclei, as well as between the data and GRAZING calculations, was carried out.Finally, recent results concerning the measurement of the excitation function from the Coulomb barrier to far below for the 92Mo+54Fe system, where both proton stripping and pickup channels were populated with similar strength, will be discussed.
APA, Harvard, Vancouver, ISO, and other styles
18

BOHUN, C. S. "A Stefan model for mass transfer in a rotating disk reaction vessel." European Journal of Applied Mathematics 26, no. 4 (May 4, 2015): 453–75. http://dx.doi.org/10.1017/s0956792515000145.

Full text
Abstract:
In this paper, we focus on the process of mass transfer in the rotating disk apparatus formulated as a Stefan problem with consideration given to both the hydrodynamics of the process and the specific chemical reactions occurring in the bulk. The wide range in the reaction rates of the underlying chemistry allows for a natural decoupling of the problem into a simplified set of weakly coupled convective–reaction–diffusion equations for the slowly reacting chemical species and a set of algebraic relations for the species that react rapidly. An analysis of the chemical equilibrium conditions identifies an expansion parameter and a reduced model that remains valid for arbitrarily large times. Numerical solutions of the model are compared to an asymptotic analysis revealing three distinct time scales and chemical diffusion boundary layer that lies completely inside the hydrodynamic layer. Formulated as a Stefan problem, the model generalizes the work of Levich (Levich and Spalding (1962) Physicochemical hydrodynamics, vol. 689, Prentice-Hall Englewood Cliffs, NJ) and will help better understand the natural limitations of the rotating disk reaction vessel when consideration is made for the reacting chemical species.
APA, Harvard, Vancouver, ISO, and other styles
19

Subbarao, Duvvuri, Reem Hassan Abd Elghafoor Hassan, and Marappagounder Ramasamy. "Heat Transfer with Chemical Reaction in Wall Heated Packed Bed Reactor." Applied Mechanics and Materials 625 (September 2014): 722–25. http://dx.doi.org/10.4028/www.scientific.net/amm.625.722.

Full text
Abstract:
Information on wall heat transfer to packed bed reactors operating with exothermic or endothermic reactions is scarce. Overall wall heat transfer coefficients in a packed bed reactor in presence of an endothermic reaction are measured and observed to be smaller than the expected in the absence of reaction. This observation is in contrast with the reported observations with exothermic reactions in packed beds. A model equation based on energy balance is presented to explain the observations.
APA, Harvard, Vancouver, ISO, and other styles
20

Arnaut, Luís G., and Sebastião J. Formosinho. "Theory of electron transfer reactions in photosynthetic bacteria reaction centers." Journal of Photochemistry and Photobiology A: Chemistry 111, no. 1-3 (December 1997): 111–38. http://dx.doi.org/10.1016/s1010-6030(97)00225-6.

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

Osman, A., and A. A. Farra. "Direct reaction mechanism for heavy ion reactions with particle transfer." Journal of Physics G: Nuclear and Particle Physics 15, no. 6 (June 1, 1989): 871–92. http://dx.doi.org/10.1088/0954-3899/15/6/016.

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

Castejon, Henry, and Kenneth B. Wiberg. "Solvent Effects on Methyl Transfer Reactions. 1. The Menshutkin Reaction." Journal of the American Chemical Society 121, no. 10 (March 1999): 2139–46. http://dx.doi.org/10.1021/ja983736t.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Burshtein, A. I., P. A. Frantsuzov, and A. A. Zharikov. "Interference of reaction channels for highly exothermic electron transfer reactions." Journal of Chemical Physics 96, no. 6 (March 15, 1992): 4261–65. http://dx.doi.org/10.1063/1.462819.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Kaneko, Yu, Shigehiko Hayashi, and Iwao Ohmine. "Proton-Transfer Reactions in Reaction Center of Photosynthetic BacteriaRhodobacter sphaeroides." Journal of Physical Chemistry B 113, no. 26 (July 2, 2009): 8993–9003. http://dx.doi.org/10.1021/jp9008898.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Koper, Marc T. M., and Wolfgang Schmickler. "A Kramers reaction rate theory for electrochemical ion transfer reactions." Chemical Physics 211, no. 1-3 (November 1996): 123–33. http://dx.doi.org/10.1016/0301-0104(96)00248-0.

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

Martini, I. B., E. R. Barthel, and B. J. Schwartz. "Building a molecular-level picture of the ultrafast dynamics of the charge-transfer-to-solvent (CTTS) reaction of sodide (Na¯)." Pure and Applied Chemistry 76, no. 10 (January 1, 2004): 1809–23. http://dx.doi.org/10.1351/pac200476101809.

Full text
Abstract:
Charge-transfer-to-solvent (CTTS) reactions represent the simplest possible electron-transfer reaction. One of the reasons that such reactions have become the subject of recent interest is that transfer of a CTTS electron from an atomic anion to the solvent involves only electronic degrees of freedom, so that all the dynamics involved in the reaction are those of the solvent. Thus, CTTS reactions provide an outstanding spectroscopic window on the dynamics of the solvent during electron transfer. In this paper, we will review our recent work studying the CTTS reaction of the sodium anion, (Na− or sodide) in a series of ether solvents. By comparing the results of ultrafast spectroscopic pump/probe experiments and mixed quantum/classical molecular dynamics simulations, we work to build a molecular-level picture of how solvent motions control the dynamics of CTTS, including the distance to which the electron is ejected and the rates of both the forward and back electron-transfer reactions.
APA, Harvard, Vancouver, ISO, and other styles
27

Wang, Chuan-Hui, Chen-Fu Liu, and Guo-Wu Rao. "Green Application of Phase-Transfer Catalysis in Oxidation: A Comprehensive Review." Mini-Reviews in Organic Chemistry 17, no. 4 (May 31, 2020): 405–11. http://dx.doi.org/10.2174/1570193x16666190617154733.

Full text
Abstract:
Oxidation reactions have emerged as one of the most versatile tools in organic chemistry. Various onium salts such as ammonium, phosphonium, arsonium, bismuthonium, tellurium have been used as phase transfer catalysts in many oxidation reactions. Certainly, considerable catalysts have been widely used in Phase-Transfer Catalysis (PTC). This review focuses on the application of PTC in various oxidation reaction. Furthermore, PTC also conforms to the concept of “Green Chemistry”. <p></p> • Oxidation has become one of the most widely used tools in organic chemistry and phase transfer catalysts has been widely used in oxidation. <p></p> • The application of phase transfer catalysts in oxidation reaction will be summarized. <p></p> • Phase transfer catalysts have important application in various oxidation reaction.
APA, Harvard, Vancouver, ISO, and other styles
28

Do, Cong Cuong, Hoang Phuc Nguyen, and Tri Toan Phuc Nguyen. "Coupled reaction channels study of the ¹⁶O(d,⁶Li) reaction." Nuclear Science and Technology 10, no. 2 (August 9, 2021): 52–59. http://dx.doi.org/10.53747/jnst.v10i2.37.

Full text
Abstract:
The transfer 16O(d,6Li)12C reaction has been studied within the coupled reaction channels (CRC) approach, inluding both the direct and indirect α transfer processes. The obtained results show an important contribution of the indirect α transfer via the 2+ and 4+ states of 12C. The CRC results show that the best-fit α spectroscopic factors of 16O becomes smaller when the indirect transfer processes are taken into account. The α spectroscopic factors deduced from the present CRC analysis of the 16O(d,6Li)12C reaction data measured at Ed=54.25 and 80 MeV are quite close to each other.
APA, Harvard, Vancouver, ISO, and other styles
29

Kumari, Prabla, Alaka Das, Dillip Kumar Baral, A. K. Pattanaik, and P. Mohanty. "Nonenzymatic NADH-Dependent Reduction of Keggin-Type 12-Tungstocobaltate(III) in Aqueous Medium." E-Journal of Chemistry 8, no. 3 (2011): 1152–57. http://dx.doi.org/10.1155/2011/341865.

Full text
Abstract:
The kinetics of the electron transfer reaction of NADH with 12-tungstocobaltate(III) has been studied over the range 5.07 ≤ 104[NADH] ≤ 15.22 mol dm-3, 7.0 ≤ pH ≤ 8.0 and 20 ≤ t ≤ 35oC in aqueous medium. The electron transfer reaction showed first-order dependence each in [NADH]Tand [12-tungstocobaltate(III)]T. The products of the reaction were found to be NAD+and 12-tungstocobaltate(II). The activation parameters ΔH#(kJ mol-1) and ΔS#(JK-1mol-1) of the electron transfer reactions were found to be 64.4±1.8 and -48.86±6.0. Negative value of ΔS#is an indicative of an ordered transition state for the electron transfer reaction.
APA, Harvard, Vancouver, ISO, and other styles
30

Lim, Kieran F., and John I. Brauman. "Reaction dynamics on barrierless reaction surfaces: A model for isoergic gas‐phase proton‐transfer reactions." Journal of Chemical Physics 94, no. 11 (June 1991): 7164–80. http://dx.doi.org/10.1063/1.460724.

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

Ekrami, Saeid, and Hamid Reza Shamlouei. "Ab initio study of C20 nanocluster effects on electrochemical properties of tetraphenylporphyrin." Journal of Porphyrins and Phthalocyanines 22, no. 08 (August 2018): 640–45. http://dx.doi.org/10.1142/s1088424618500773.

Full text
Abstract:
The Density Functional Theory (DFT) method was employed to study the properties of the C[Formula: see text] complex with tetraphenylporphyrin (TPP). Calculations were performed in vacuum and in the presence of different solvents. Strong interaction between the C[Formula: see text] cluster and TPP molecule was observed. To understand the effect of C[Formula: see text] on electrochemical properties of TPP, electron transfers from and toward the porphyrin and C[Formula: see text]-TPP complex were studied. It was shown that the presence of C[Formula: see text] influences the electron transfer reaction toward the porphyrin molecule and causes transfer of one and two electrons to C[Formula: see text]-porphyrin, which is more favorable compared with porphyrin alone. However, C[Formula: see text] has slight effect on electron transfer from porphyrin and on positive ion formation. The effect of solvent type on electron transfer energy was studied for these reactions, and it was shown that solvents with higher permittivity have lower electron transfer reaction energy, which may be predicted from ionic character of the products.
APA, Harvard, Vancouver, ISO, and other styles
32

Fox, A., A. B. Raksit, S. Dheandhanoo, and D. K. Bohme. "Selected-ion flow tube studies of reactions of the radical cation (HC3N)+• in the interstellar chemical synthesis of cyanoacetylene." Canadian Journal of Chemistry 64, no. 2 (February 1, 1986): 399–403. http://dx.doi.org/10.1139/v86-064.

Full text
Abstract:
The radical cation (HC3N)+• was produced in a Selected-Ion Flow Tube (SIFT) apparatus from cyanoacetylene by electron impact and reacted at room temperature in helium buffer gas with a selection of molecules including H2, CO, HCN, CH4, H2O, O2, HC3N, C2H2, OCS, C2H4, and C4H2. The observed reactions exhibited a wide range of reactivity and a variety of pathways including charge transfer, hydrogen atom transfer, proton transfer, and association. Association reactions were observed with CO, O2, HCN, and HC3N. With the latter two molecules association was observed to proceed close to the collision limit, which is suggestive of covalent bond formation perhaps involving azine-like N—N bonds. For HC3N an equally rapid association has been observed by Buckley etal. with ICR (Ion Cyclotron Resonance) measurements at low pressures and this is suggestive of radiative association. The hydrogen atom transfer reaction of ionized cyanoacetylene with H2 is slow while similar reactions with CH4 and H2O are fast. The reaction with CO fails to transfer a proton. These results have implications for synthetic schemes for cyanoacetylene as proposed in recent models of the chemistry of interstellar gas clouds. Proton transfer was also observed to be curiously unfavourable with all other molecules having a proton affinity higher than (C3N)•. Also, hydrogen-atom transfer was inefficient with the polar molecules HCN and HC3N. These results suggest that interactions at close separations may lead to preferential alignment of the reacting ion and molecule which is not suited for proton transfer or hydrogen atom transfer.
APA, Harvard, Vancouver, ISO, and other styles
33

Gong, Yunhao, Ermei Yao, Melissa Stevens, and John E. Tavis. "Evidence that the first strand-transfer reaction of duck hepatitis B virus reverse transcription requires the polymerase and that strand transfer is not needed for the switch of the polymerase to the elongation mode of DNA synthesis." Journal of General Virology 81, no. 8 (August 1, 2000): 2059–65. http://dx.doi.org/10.1099/0022-1317-81-8-2059.

Full text
Abstract:
Deletion of amino acids 79–88 in the duck hepatitis B virus reverse transcriptase had minimal effects on polymerase activities prior to the minus-strand DNA transfer reaction, yet it greatly diminished strand transfer and subsequent DNA synthesis. This mutation also reduced reverse transcription on exogenous RNA templates. The reaction on exogenous RNAs employed the phosphonoformic acid (PFA)-sensitive elongation mode of DNA synthesis rather than the PFA-resistant priming mode, despite the independence of DNA synthesis in this assay from the priming and minus-strand transfer reactions. These data provide experimental evidence that the polymerase is involved directly in the minus-strand transfer reaction and that the switch of the polymerase from the early PFA-resistant mode of DNA synthesis to the later PFA-sensitive elongation mode does not require the strand-transfer reaction.
APA, Harvard, Vancouver, ISO, and other styles
34

Khorasani, Sanaz, and Manuel Fernandes. "Cooperativity in the Reaction of 9-Methylanthracene With a 1,4-Dithiin Molecule." Acta Crystallographica Section A Foundations and Advances 70, a1 (August 5, 2014): C913. http://dx.doi.org/10.1107/s205327331409086x.

Full text
Abstract:
Solid-state chemistry involves the manipulation of molecules and materials through photochemical, thermal, or mechanical reaction methods. Single-crystal-to-single-crystal (SCSC) reactions are rare, but offer the opportunity to study reaction mechanisms and molecular motions in the solid state at the atomic level using single crystal X-ray diffraction. This allows the effect of the surrounding molecules, and hence the reaction cavity, on the reacting molecules to be examined which may ultimately lead to postcrystallization methods for creating new materials or reaction products that cannot easily be obtained via solution. SCSC reactions involving two different molecules are relatively uncommon. A convenient system that allows the study of such reactions is the [4+2] Diels-Alder reaction of 1,4-dithiintetracarboxylic type compounds with anthracene derivatives. In the work reported here, electron donor to acceptor interactions between 9-Methylanthracene and bis(N-cyclobutylimino)-1,4-dithiin lead to the formation of chiral charge transfer (CT) crystals [1]. These undergo a topochemical thermal SCSC [4 + 2] Diels-Alder reaction in the solid state. CT crystals were reacted at 400C, their structures determined by X-ray diffraction at various degrees of conversion, and examined using Hirshfeld surfaces and lattice energy calculations to find evidence of reaction cooperativity and feedback mechanisms. In this case, a maximum reaction conversion of around 96% was obtained indicating that the reaction is non-random within the charge transfer stacks, with close contacts between product molecules in the reacted crystal also providing some evidence for reaction cooperativity along the b axis perpendicular to the CT stacking axis.
APA, Harvard, Vancouver, ISO, and other styles
35

Kiyotani, Tamiko, Masayoshi Kobayashi, Ichiro Tanaka, and Nobuo Niimura. "Observation of electron transfer associated with enzymatic process by muSR." Acta Crystallographica Section A Foundations and Advances 70, a1 (August 5, 2014): C1463. http://dx.doi.org/10.1107/s2053273314085362.

Full text
Abstract:
We propose muSR experiments on trypsin-BPTI complex to visualize the electron and proton transfer processes occurring in the catalytic reaction of the trypsin. The mechanism of an inhibitory effect of the BPTI is interpreted that the reaction products of BPTI remain at a part of the structure and the reverse reaction reforms the stable trypsin–BPTI complex, which has been confirmed by neutron diffraction experiment of the trypsin-BPTI complex [1]. However, it never sees the real image of the proton and electron transfer processes directly. According to the model provided by the results of neutron diffraction experiments, the proton and electron transfer processes are continuously occurring in a crystal of trypsin-BPTI complex and the process induces the local magnetic field. The slow muon is very adequate because the position, where mu+ is captured, is absolutely negatively charged oxyanion hole close to the reaction center of Trypsin. The distance between the oxyanion hole and the active peptide bond is about 10Å. When the turn over time of the catalytic reactions is assumed to be 10msec or so, the induced magnetic field would be estimated as 0.2 micro-T. In order to check the effectiveness of the measurement of the μSR experiments on trypsin-BPTI complex, another measurement of the muSR experiments on the trypsin- MIP complex is adequate [2]. Here, MIP is a kind of the trypsin inhibitor, which completely stops the catalytic reaction of trypsin. In the trypsin-MIP complex, no electron and proton transfers at all in the active site of trypsin and captured mu+ at the oxyanion hole would never be sensitive to the induced magnetic field [3].
APA, Harvard, Vancouver, ISO, and other styles
36

Empel, Claire, Sripati Jana, and Rene M. Koenigs. "C-H Functionalization via Iron-Catalyzed Carbene-Transfer Reactions." Molecules 25, no. 4 (February 17, 2020): 880. http://dx.doi.org/10.3390/molecules25040880.

Full text
Abstract:
The direct C-H functionalization reaction is one of the most efficient strategies by which to introduce new functional groups into small organic molecules. Over time, iron complexes have emerged as versatile catalysts for carbine-transfer reactions with diazoalkanes under mild and sustainable reaction conditions. In this review, we discuss the advances that have been made using iron catalysts to perform C-H functionalization reactions with diazoalkanes. We give an overview of early examples employing stoichiometric iron carbene complexes and continue with recent advances in the C-H functionalization of C(sp2)-H and C(sp3)-H bonds, concluding with the latest developments in enzymatic C-H functionalization reactions using iron-heme-containing enzymes.
APA, Harvard, Vancouver, ISO, and other styles
37

Schöneich, Christian. "Radical rearrangement and transfer reactions in proteins." Essays in Biochemistry 64, no. 1 (January 10, 2020): 87–96. http://dx.doi.org/10.1042/ebc20190046.

Full text
Abstract:
Abstract Radical rearrangement and transfer reactions play an important role in the chemical modifications of proteins in vivo and in vitro. These reactions depend on protein sequence, as well as structure and dynamics. Frequently, these reactions have well-defined precedents in the organic chemistry literature, but their occurrence in proteins provides a stage for a number of novel and, perhaps, unexpected reaction products. This essay will provide an overview over a few representative examples of radical rearrangement and transfer reactions.
APA, Harvard, Vancouver, ISO, and other styles
38

Hewavitharanage, Priyadarshine, Evgeny O. Danilov, and Douglas C. Neckers. "Pentafluorophenyl Transfer: A New Group-Transfer Reaction in Organoborate Salts." Journal of Organic Chemistry 70, no. 26 (December 2005): 10653–59. http://dx.doi.org/10.1021/jo050695s.

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

Khenkin, Alexander M., Irena Efremenko, Jan M. L. Martin, and Ronny Neumann. "The kinetics and mechanism of oxidation of reduced phosphovanadomolybdates by molecular oxygen: theory and experiment in concert." Physical Chemistry Chemical Physics 20, no. 11 (2018): 7579–87. http://dx.doi.org/10.1039/c7cp08610e.

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

Spesyvyi, Anatolii, David Smith, and Patrik Španěl. "Ion chemistry at elevated ion–molecule interaction energies in a selected ion flow-drift tube: reactions of H3O+, NO+ and O2+ with saturated aliphatic ketones." Physical Chemistry Chemical Physics 19, no. 47 (2017): 31714–23. http://dx.doi.org/10.1039/c7cp05795d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Ulmer, P. E. "Polarization transfer in the 4HeH reaction." Nuclear Physics A 689, no. 1-2 (June 2001): 94–103. http://dx.doi.org/10.1016/s0375-9474(01)00823-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Dieterich, S., P. Bartsch, D. Baumann, J. Bermuth, K. Bohinc, R. Böhm, D. Bosnar, et al. "Polarization transfer in the 4HeH reaction." Physics Letters B 500, no. 1-2 (February 2001): 47–52. http://dx.doi.org/10.1016/s0370-2693(01)00052-1.

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

STINSON, STEPHEN. "Novel electron-transfer chain reaction discovered." Chemical & Engineering News 66, no. 17 (April 25, 1988): 22–23. http://dx.doi.org/10.1021/cen-v066n017.p022.

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

Ghosh, Somnath, and Indira Datta. "Diazo Transfer Reaction in Solid State." Synthetic Communications 21, no. 2 (January 1991): 191–200. http://dx.doi.org/10.1080/00397919108020811.

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

Gerlits, Oksana, Jianhui Tian, Amit Das, Paul Langan, William T. Heller, and Andrey Kovalevsky. "Phosphoryl Transfer Reaction Snapshots in Crystals." Journal of Biological Chemistry 290, no. 25 (April 28, 2015): 15538–48. http://dx.doi.org/10.1074/jbc.m115.643213.

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

Arai, Shigeru, Yoshiki Shirai, Toshimasa Ishida, and Takayuki Shioiri. "Phase-transfer-catalyzed asymmetric Darzens reaction." Tetrahedron 55, no. 20 (May 1999): 6375–86. http://dx.doi.org/10.1016/s0040-4020(99)00213-6.

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

Gu, Huan, Yuan Guo, and Zhen Shi. "Covert Mannich Reaction via Carbon Transfer." Synthetic Communications 36, no. 22 (November 1, 2006): 3335–38. http://dx.doi.org/10.1080/00397910600941299.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Bhunnoo, Riaz A., Yulai Hu, Dramane I. Lainé, and Richard C. D. Brown. "An Asymmetric Phase-Transfer Dihydroxylation Reaction." Angewandte Chemie International Edition 41, no. 18 (September 16, 2002): 3479–80. http://dx.doi.org/10.1002/1521-3773(20020916)41:18<3479::aid-anie3479>3.0.co;2-o.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Rao, Santhosh, Puneeth Kumar Someswara Ashwathappa, and Kandikere Ramaiah Prabhu. "Boron-Catalyzed Carbonate Functionality Transfer Reaction." Asian Journal of Organic Chemistry 8, no. 3 (February 7, 2019): 320–23. http://dx.doi.org/10.1002/ajoc.201800751.

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

Loo, Marcia E. Van, Johan Lugtenburg, and Jan Cornelisse. "Single Electron Transfer vs. SN2 Reaction." European Journal of Organic Chemistry 2000, no. 5 (March 2000): 713–21. http://dx.doi.org/10.1002/(sici)1099-0690(200003)2000:5<713::aid-ejoc713>3.0.co;2-u.

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