• Title/Summary/Keyword: Steric hindrance

Search Result 135, Processing Time 0.024 seconds

A Kinetic Study on Aminolysis of Benzyl 2-Pyridyl Thionocarbonate and t-Butyl 2-Pyridyl Thionocarbonate: Effects of Polarizability and Steric Hindrance on Reactivity and Reaction Mechanism

  • Kim, Min-Young;Bae, Ae Ri;Um, Ik-Hwan
    • Bulletin of the Korean Chemical Society
    • /
    • v.34 no.8
    • /
    • pp.2325-2329
    • /
    • 2013
  • Second-order rate constants $k_N$ have been measured for reactions of benzyl 2-pyridyl thionocarbonate (4b) and t-butyl 2-pyridyl thionocarbonate (5b) with a series of cyclic secondary amines in MeCN at $25.0{\pm}0.1^{\circ}C$. The $k_N$ values for the reactions of 4b and 5b have been compared with those reported previously for the corresponding reactions of benzyl 2-pyridyl carbonate (4a) and t-butyl 2-pyridyl carbonate (5a) to investigate the effect of changing the electrophilic center from C=O to C=S on reactivity and reaction mechanism. The thiono compound 4b is more reactive than its oxygen analogue 4a. The Br${\o}$nsted-type plots for the reactions of 4a and 4b are linear with ${\beta}_{nuc}=0.57$ and 0.37, respectively. The reactions of 4a were previously reported to proceed through a concerted mechanism, while those of 4b in this study have been concluded to proceed through a stepwise mechanism with formation of an intermediate being the rate-determining step on the basis of the ${\beta}_{nuc}$ value of 0.37. Enhanced polarizability upon changing the C=O in 4a by C=S has been suggested to be responsible for the reactivity order and the contrasting reaction mechanisms. In contrast, the reactivity of 5a and 5b is similar, but they are much less reactive than 4a and 4b. Furthermore, the reactions of 5a and 5b have been concluded to proceed through the same mechanism (i.e., a concerted mechanism) on the basis of linear Bronsted-type plots with ${\beta}_{nuc}=0.45$ or 0.47. It has been concluded that the strong steric hindrance exerted by the t-Bu in 5a and 5b causes a decrease in their reactivity and forces the reactions to proceed through a concerted mechanism.

Electrochemical Determinations of Methylanilinium Ion Mixtures by the Stereoselective Complexations of Host-Guest (호스트-게스트의 입체선택적 착물형성에 의한 메틸아닐리늄 이온 혼합물의 전기화학적 분석)

  • Jung, Jong Hwa;Chang, Duk Jin;Lee, Bu-Yong;Seo, Mu Ryong;Kim, Jae Sang;Lee, Shim Sung
    • Analytical Science and Technology
    • /
    • v.5 no.3
    • /
    • pp.239-247
    • /
    • 1992
  • The complexations and selectivities of the 10 species of mono- and dimethylsubstituted anilinium ions with 18-crown-6 in methanol are examined at dropping mercury electrode. The stability constants of these complexes varies drastically due to the steric hindrance by the positions and numbers of methyl groups. And the analyses of the isomeric mixtures of methylanilinium, which are impossible to detect due to the overlapped peaks in normal conditions, were also accomplished by the additions of 18-crown-6 as the supporting complexing agent using the selective complexations by the steric hindrance effects. As results in case of the difference of stability, ${\Delta}log\;K$ were about 0.7~1.3, it was possible to confirm the existence of two species qualitatively. Otherwise when ${\Delta}log\;K$ were large than 1.6, the quantitative determinations of each species could be accomplished sucessfully. From these results it is deduced that the selective recongnition of the positions and numbers of methyl groups as the steric hindrance in anililniums by 18-crown-6 cause the large variation of the magnitudes of negative shift of reduction waves for guest ions in mixtures.

  • PDF

Aminolysis of 2,4-Dinitrophenyl and 3,4-Dinitrophenyl 2-Thiophenecarboxylates: Effect of ortho-Nitro Group on Reactivity and Mechanism

  • Seo, Jin-A;Chun, Sun-Mee;Um, Ik-Hwan
    • Bulletin of the Korean Chemical Society
    • /
    • v.29 no.8
    • /
    • pp.1459-1463
    • /
    • 2008
  • Second-order rate constants (kN) have been measured spectrophotometrically for reactions of 3,4-dinitrophenyl 2-thiophenecarboxylate (2) with a series of alicyclic secondary amines in 80 mol % $H_2O$/20 mol % dimethyl sulfoxide at 25.0 ${^{\circ}C}$. The Brønsted-type plot exhibits a downward curvature for the aminolysis of 2. The curved Brønsted-type plot is similar to that reported for the corresponding reactions of 2,4-dinitrophenyl 2- thiophenecarboxylate (1). The reactions of 1 and 2 have been suggested to proceed through the same mechanism, i.e., through a zwitterionic tetrahedral intermediate ($T^{\pm}$) with a change in the rate-determining step. Substrate 2 is less reactive than 1 toward weakly basic amines (e.g., $pK_a$ < 10.4) but becomes more reactive as the basicity of amines increases further. Dissection of kN into the microscopic rate constants has revealed that the reaction of 2 results in a smaller $k_2/k_{-1}$ ratio but larger $k_1$ than the corresponding reaction of 1. Steric hindrance exerted by the ortho-nitro group has been suggested to be responsible for the smaller $k_1$ value found for the reactions of 1.

Thermal Properties and fracture Toughness of Difunctional Epoxy Resins Cured by Catalytic Initiators (촉매형 개시제로 경화된 이관능성 에폭시 수지의 열적 특성 및 파괴인성)

  • 박수진;허건영;이재락
    • Polymer(Korea)
    • /
    • v.26 no.3
    • /
    • pp.344-352
    • /
    • 2002
  • In this work, two thermal cationic latent catalysts, i.e., triphenyl benzyl phosphonium hexafluoroantimonate (TBPH) and benzyl 2-methylpyrazinium hexafluoroantimonate (BMPH) were newly synthesized. And the thermal and mechanical properties of difunctional epoxy (diglycidylether of bisphenol h, DGEBA) resins initiated by 1 phr of either TBPH or BMPH catalyst were investigated. As experimental results, the epoxy/TBPH system showed higher curing temperature and critical stress intensity factor ($K_{IC}$) than those of epoxy/BMPH. This could be interpreted in terms of slow thermal diffusion rate and bulk structure of four phenyl groups in TBPH. However, the decomposed activation energy determined from Coats-Redfern method was lower in the case of epoxy/TBPH. This result was probably due to the fact that broken short chain structure was developed by steric hindrance of TBPH.

Naphthazarin Derivatives: Synthesis, Cytotoxic Mechanism and Evaluation of Antitumor Activity

  • You, Young-Jae;Zheng, Xiang-Guo;Kim, Yong;Ahn, Byung-Zun
    • Archives of Pharmacal Research
    • /
    • v.21 no.5
    • /
    • pp.595-598
    • /
    • 1998
  • The rate of the GSH conjugate formation, the inhibition of DNA topoisomerase-I and the cytotoxic activity against L1210 cells of the naphthoquinones showed the same order; 5,8-dimethoxy-1,4-naphthoquinone (DMNQ)>6-(1-hydroxyethyl)-DMNQ>2-(1-hydroxyethyl)-DMNQ; the steric hindrance of the substituents, particularly 2-substutuent, in reacting with cellular nucleophiles must be the main cause for lowering the bioactivities. Acetylation of 2-(1-hydroxyethyl)-DMNQ producing 2-(acetyloxyethyl)-DMNQ potentiated the bioactivities; 2-(-hydroxyethyl)-DMNQ did not react with GSH and the enzyme, and showed $ED_{50}$ of 0.146 mg/ml for the cytotoxcity. Furthermore, the acetylation 2-(1-hydroxyethyl)-DMNQ(T/C, 119%) enhanced the T/C values for the mice bearing S-180 tumor {T/C of 2-(1-acetyloxyethyl)-DMNQ, 276%]. It was assumed that the difference in bioactivities ensued by acetylation was based on the mechanism of the so-called bioreductive alkylation.

  • PDF

Aminolyses of 2,4-Dinitrophenyl and 3,4-Dinitrophenyl 2-Furoates: Effect of ortho-Substituent on Reactivity and Mechanism

  • Um, Ik-Hwan;Akhtar, Kalsoom
    • Bulletin of the Korean Chemical Society
    • /
    • v.29 no.4
    • /
    • pp.772-776
    • /
    • 2008
  • Second-order rate constants ($k_N$) have been measured spectrophotometrically for reactions of 3,4-dintrophenyl 2-furoate (2) with a series of secondary alicyclic amines in 80 mol % $H_2O$/20 mol % dimethyl sulfoxide (DMSO) at 25.0 ${^{\circ}C}$. The Bronsted-type plot exhibits a downward curvature for the aminolysis of 2, which is similar to that reported for the corresponding reactions of 2,4-dintrophenyl 2-furoate (1). Substrate 2 is less reactive than 1 toward all the amines studied but the reactivity difference becomes smaller as the amine basicity increases. Dissection of the second-order rate constants into the microscopic rate constants has revealed that the reaction of 2 results in a smaller $k_2/k_{-1}$ ratio but slightly larger $k_1$ value than that of 1. Steric hindrance has been suggested to be responsible for the smaller $k_1$ value found for the reactions of 1, since the ortho-substituent of 1 would inhibit the attack of amines (i.e., the $k_1$ process).

Effect of Substituted Trifluoromethyl Groups on Thermal and Mechanical Properties of Fluorine-containing Epoxy Resin

  • Heo, Gun-Young;Park, Soo-Jin
    • Macromolecular Research
    • /
    • v.17 no.11
    • /
    • pp.870-873
    • /
    • 2009
  • In this study, 2-diglycidylether of benzotrifluoride (2-DGEBTF) and 4-diglycidylether of benzotrifluoride (4-DGEBTF) epoxy resins, which contained fluorine groups in the main chain, were synthesized. The resins were characterized by FTIR, $^1H$ NMR, $^{13}C$ NMR and $^{19}F$ NMR spectroscopy. The 2-DGEBTF and 4-DGEBTF epoxy resins were cured with triethylene tetramine (TETA), and the effect of the fluorine group on the synthesized epoxy resin on the cure behavior, thermal, and mechanical properties was investigated. The 2-DGEBTF/TETA system was more reactive than the 4-DGEBTF/TETA system, whereas the thermal stability factor i.e., the decomposition activation energy ($E_d$), of 4-DGEBTF/TETA was higher than that of 2-DGEBTF/TETA. These results can be explained by the decrease in cross-linking density and decomposition of the short side chains, resulting in the $CF_3$ group at the para position. However, the $K_{IC}$ value of 4-DGEBTF/TETA was higher than that of 2-DGEBTF/TETA. This was attributed to the increase in flexibility in the epoxy backbone, resulting in a difference in steric hindrance and polarlizability.

Analysis of the Heat of Absorption Based on the Chemical Structures of Carbon Dioxide Absorbents (이산화탄소 흡수제의 화학구조별 반응열량 특성 연구)

  • Kwak, No Sang;Lee, Ji Hyun;Eom, Yong Seok;Kim, Jun Han;Lee, In Young;Jang, Kyung Ryoung;Shim, Jae-Goo
    • Korean Chemical Engineering Research
    • /
    • v.50 no.1
    • /
    • pp.135-140
    • /
    • 2012
  • In this study, the heats of absorption of $CO_2$ with aqueous solutions of primary, secondary and tertiary amine aqueous solutions were measured in the commercial reaction calorimeter SIMULAR (HEL, UK). The heats of absorption of 30 wt% amine aqueous solutions of MEA (monoethanolamine, primary amine), EAE(2-(ethylamino)ethanol, secondary amine), and MDEA (methyldiethanolamine, tertiary amine) were measured as function of the $CO_2$ loading ratio at $40^{\circ}C$, in each case. In addition, the heats of absorption of sterically-hindered amine aqueous solutions of AMP(2-amino-2-methyl-1-propanol, primary amine), DEA(diethanolamine, secondary amine) and TEA(triethanolamine, tertiary amine) were measured to observe the steric hindrance effect. The heat of absorption is high in the following order regardless of the steric hindrance: primary amine > secondary amine > tertiary amine. The heats of absorption of amines having sterically-hindered substituents surrounding nitrogen atoms are relatively low compare to that of sterically-free amines, although the difference is very small.

A Mechanistic Study for Aminolysis of p-Nitrophenyl Phenylacetate

  • 엄익환;Yeom, E. Suk;권혜진;권동숙
    • Bulletin of the Korean Chemical Society
    • /
    • v.18 no.8
    • /
    • pp.865-868
    • /
    • 1997
  • Second-order rate constants have been measured spectrophotometrically for the reactions of p-nitrophenyl phenylacetate (1) and benzoate (2) with a series of alicyclic amines in H2O containing 20 mole % DMSO at 25.0 ℃. 1 appears to be more reactive than 2 toward all the amines studied, although phenylacetic acid is a weaker acid than benzoic acid. The higher reactivity of 1 can be attributed to resonance and/or steric effect, since the ground state of 2 can be stabilized by resonance and 1 would experience less steric hindrance due to the presence of CH2 group between phenyl and C=O group. The reactivity of the amines increases with increasing their basicity. The Bronsted-type plots for aminolysis of 1 and 2 show good linearity with βnuc values of 0.81 and 0.85, respectively, indicating that the TS structures of the aminolyses of 1 and 2 are similar. Besides, the linear Bronsted plots obtained in the present system clearly suggest that there is no mechanism change for the given series of the amines and the reactions of 1 and 2 proceed in a same mechanism.

Transglycosylation of Permethylated Methyl D-Glycopyranosides in the Presence of Trimethylsilyl Trifluoromethanesulfonate

  • 이창귀;전정호;서영환
    • Bulletin of the Korean Chemical Society
    • /
    • v.19 no.11
    • /
    • pp.1233-1238
    • /
    • 1998
  • Transglycosylation reactions among methyl 2,3,4,6-tetra-O-methyl-D-glycopyranosides and isomeric butyl alcohols or cyclohexanol took place in the presence of trimethylsilyl trifluoromethanesulfonate (TMSOTf) in dichloromethane. The extent of the reaction after 1 h and 24 h from mixing was determined by gas chromatography (GC). Anomerization of the substrate took place during the course of transglycosylation, which favors α anomer regardless of the anomeric configurations of the starting glycosides. Transglycosylation also favors the a anomer regardless of the steric bulkiness of the alcohol. tert-Butyl alcohol did not give any transglycosylation, suggesting the steric hindrance of approaching the bulky alcohol to the oxonium intermediate. A mechanism for the transglycosylation have been proposed.