Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 42
Filtrar
1.
Acc Chem Res ; 53(11): 2680-2691, 2020 11 17.
Artículo en Inglés | MEDLINE | ID: mdl-33063516

RESUMEN

SmI2 was introduced to organic chemistry as a single electron transfer agent in 1977. After ca. 15 years of latency, the scientific community has realized the high potential of this reagent, and its chemistry has started blooming. This versatile reagent has mediated a myriad of new bond formations, cyclizations, and other reactions. Its popularity stems largely from the fact that three different intermediates, radical anions, radicals, and anions, depending on the ligand or additive used, could be obtained. Each of these intermediates could in principle lead to a different product. While these options vastly enrich the repertoire of SmI2, they necessitate a thorough mechanistic understanding, especially concerning how appropriate ligands direct the SmI2 to the desired intermediate. Our first paper on this subject dealt with the reduction of an activated double bond. The results were puzzling, especially the H/D isotope effect, which depended on the order of the reagents addition. This seminal paper was fundamental to an understanding of how the SmI2 works and enabled us to later explain various phenomena. For example, it was found that in a given reaction, when MeOH is used as a proton source, a spiro compound is obtained, while a bicyclic product is obtained when t-BuOH is used. Our contribution culminated in formulating guidelines for the rational use of proton donors in SmI2 reactions.The need to understand the complexity of the effect of additives on various processes is nicely demonstrated in photoinduced reactions. For example, hexamethylphosphoramide (HMPA) enhances the reduction of anthracene while hampering the reaction of benzyl chloride. The mechanistic understanding gained enabled us also to broaden the scope of photostimulated reactions from substrates reacting by a dissociative electron transfer mechanism to normal reductions, which are difficult to accomplish at the ground state. Harnessing the classical knowledge of proton transfer mechanisms to our SmI2 research enabled us to decipher an old conundrum: why does the combination of water and amine have such an enhancing effect on the reactivity of SmI2, which is not typical of these two when used separately. In our studies on the affinity of ligands to SmI2, we discovered that, in contradistinction to the accepted dogma, SmI2 is much more azaphilic than it is oxophilic. On the basis of the size difference between Sm3+ and Sm2+, we developed a simple diagnostic tool for the nature of the steps following the electron transfer. The reduction of imines showed that substrate affinity to SmI2 plays also a crucial role. In these reactions, new features such as autocatalysis and catalysis by quantum dots were discovered. Several studies of the ligand effect lead to a clear formulation of when an inner sphere or outer sphere electron transfer should be expected. In addition, several reactions where proton-coupled electron transfer (PCET) is the dominant mechanism were identified. Finally, the surprisingly old tool of NMR "shift reagents" was rediscovered and used to directly derive essential information on the binding constants of ligands and substrates to SmI2.

2.
J Org Chem ; 86(15): 10861-10865, 2021 Aug 06.
Artículo en Inglés | MEDLINE | ID: mdl-34318663

RESUMEN

While additives play an important role in the reactions of samarium iodide, ligand-SmI2 complexation constants are scarce. Here, VIS spectroscopy was harnessed along with NMR to determine the first complexation constant for most of the alcohols and amines used in SmI2 reactions. The second equilibrium constant was determined for selected ligands. In cases where both methods could be applied, in general, a good correlation between the equilibrium constants was obtained.

3.
Chemistry ; 26(30): 6846-6850, 2020 May 26.
Artículo en Inglés | MEDLINE | ID: mdl-32045042

RESUMEN

The method developed and introduced here enables for the first time (to the authors' knowledge), a quantitative assessment of the interaction of SmI2 with substrates prior to the electron transfer stage. As a proof of concept, equilibrium constants for some model substrates including carbonyl compounds and aromatic nuclei are reported here. In addition, the first equilibrium constants with some common ligands were also determined. The equilibrium constants range from approximately 0.07 m-1 for diisopropyl ketone to 2500 m-1 for hexamethylphosphoramide (HMPA). It is shown that the data acquired by this method, which is based on the concept of shift reagents, can shed light on the most intimate details of the reaction mechanism, and this method is a useful tool for planning a synthetic process.

4.
Chemistry ; 25(44): 10499-10504, 2019 Aug 06.
Artículo en Inglés | MEDLINE | ID: mdl-31150561

RESUMEN

The work described herein compares the effect of additives (HMPA, methanol, ethylene glycol, pinacol, N-methylethanolamine) on thermal and photochemical reactions of samarium diiodide (SmI2 ). In thermal reactions, additives that coordinate to SmI2 induce a significant increase in reaction rate. In photochemical reactions, the presence of an electronegative atom with a highly localized negative charge on the substrate leads to a rate deceleration. In order to benefit from the columbic interaction with the positively charged samarium cation, these substrates react preferentially by an inner sphere reduction mechanism. The addition of ligands prevents this close interaction causing rate retardation. Furthermore, studies demonstrate that excited state quenching of SmII by ethylene glycol and other additives indicate that it is unlikely to be the major cause for the observed rate retardation. This effect provides a simple diagnostic tool to distinguish between an inner and an outer sphere reduction mechanism.

5.
J Org Chem ; 84(4): 1994-1998, 2019 02 15.
Artículo en Inglés | MEDLINE | ID: mdl-30668106

RESUMEN

Ligation plays a multifaceted role in the chemistry of SmI2. Depending on the ligand, two of its major effects are increasing the reduction potential of SmI2, and in the case of a ligand, which is also a proton donor, it may also enhance the reaction by protonation of the radical anion generated in the preceding step. It turns out that the number of ligand molecules that are needed to maximize the reduction potential of SmI2 is significantly smaller than the number of ligand molecules needed for a maximal enhancement of the protonation rate. In addition to the economical use of the ligand, this information can also be utilized as a diagnostic tool for the reaction mechanism in differentiating between single and multistep processes. The possible pitfalls in applying this diagnostic tool to PCET and cyclization reactions are discussed.

6.
Chemistry ; 23(67): 17070-17077, 2017 Dec 01.
Artículo en Inglés | MEDLINE | ID: mdl-29024166

RESUMEN

Ligands that coordinate to SmI2 through oxygen are prevalent in the literature and make up a significant portion of additives employed with the reagent to perform reactions of great synthetic importance. In the present work a series of spectroscopic, calorimetric and kinetic studies demonstrate that nitrogen-based analogues of many common additives have a significantly higher affinity for Sm than the oxygen-based counterparts. In addition, electrochemical experiments show that nitrogen-based ligands significantly enhance the reducing power of SmI2 . Overall, this work demonstrates that the use of nitrogen-based ligands provides a useful alternative approach to enhance the reactivity of reductants based on SmII .

7.
Chemistry ; 21(50): 18394-400, 2015 Dec 07.
Artículo en Inglés | MEDLINE | ID: mdl-26525449

RESUMEN

The reaction of SmI2 with the substrates 3-methyl-2-butanone, benzyl chloride, p-cyanobenzyl chloride, and anthracene were studied in the presence of water and an amine. In all cases, the water content versus rate profile shows a maximum at around 0.2 M H2 O. The rate versus amine content profile shows in all cases, except for benzyl chloride, saturation behavior, which is typical of a change in the identity of the rate-determining step. The mechanism that is in agreement with the observed data is that electron transfer occurs in the first step. With substrates that are not very electrophilic, the intermediate radical anions lose the added electron back to samarium(III) relatively quickly and the reaction cannot progress efficiently. However, in a mixture of water/amine, the amine deprotonates a molecule of water coordinated to samarium(III). The negatively charged hydroxide, which is coordinated to samarium(III), reduces its electrophilicity, and therefore, lowers the rate of back electron transfer, which allows the reaction to progress. In the case of benzyl chloride, in which electron transfer is rate determining, deprotonation by the amine is coupled to the electron-transfer step.

8.
Chemistry ; 21(25): 9242-8, 2015 Jun 15.
Artículo en Inglés | MEDLINE | ID: mdl-25965697

RESUMEN

The paradigm that the cleavage of the radical anion of benzyl halides occurs in such a way that the negative charge ends up on the departing halide leaving behind a benzyl radical is well rooted in chemistry. By studying the kinetics of the reaction of substituted benzylbromides and chlorides with SmI2 in THF it was found that substrates para-substituted with electron-withdrawing groups (CN and CO2 Me), which are capable of forming hydrogen bonds with a proton donor and coordinating to samarium cation, react in a reversed electron apportionment mode. Namely, the halide departs as a radical. This conclusion is based on the found convex Hammett plots, element effects, proton donor effects, and the effect of tosylate (OTs) as a leaving group. The latter does not tend to tolerate radical character on the oxygen atom. In the presence of a proton donor, the tolyl derivatives were the sole product, whereas in its absence, the coupling dimer was obtained by a SN 2 reaction of the benzyl anion on the neutral substrate. The data also suggest that for the para-CN and CO2 Me derivatives in the presence of a proton donor, the first electron transfer is coupled with the proton transfer.

9.
J Org Chem ; 80(17): 8929-32, 2015 Sep 04.
Artículo en Inglés | MEDLINE | ID: mdl-26266694

RESUMEN

Pyridine and its derivatives are rather difficult to reduce, and the products often undergo a very fast reoxidation to regain aromaticity. The reduction of quinoline by SmI2 results in an instantaneous [2 + 3] cyclization reaction, forming a bridged seven-membered ring within a polycyclic system.

10.
J Org Chem ; 79(6): 2682-7, 2014 Mar 21.
Artículo en Inglés | MEDLINE | ID: mdl-24559471

RESUMEN

HMPA is known to increase the reduction potential of SmI2. However, in many cases, the transferred electron returns from the radical anion of the substrate back to the Sm(3+). This could be avoided by an efficient trapping of the radical anion: e.g., by protonation. However, bimolecular protonation by a proton donor from the bulk may be too slow to compete with the back electron transfer process. An efficient unimolecular protonation could be achieved by using a proton donor which complexes to SmI2, in which case the proton is unimolecularly transferred within the ion pair. A derivative of HMPA in which one of the methyl groups was substituted by a CH2CH2OH unit was synthesized. Cyclic voltammetry studies have shown that it resembles HMPA in its ability to enhance the reduction potential of SmI2, and reactivity studies show that it has also efficient proton shift capabilities. The various aspects of this additive were examined in the reactions of SmI2 with three substrates: benzyl chloride, methyl cinnamate, and anthracene.

11.
J Phys Chem A ; 117(33): 7737-41, 2013 Aug 22.
Artículo en Inglés | MEDLINE | ID: mdl-23886075

RESUMEN

Medium variations usually affect the shape of the bimolecular nucleophilic reaction profile at the reactants' and products' ends and, to a much lesser extent, the shape around the transition state. In water, the reactions of extended allylic systems such as F(-) + H-(CH=CH)n-CH2-F → F-CH2-(CH=CH)n-H + F(-) have been computationally shown (for n = 2) to have a single transition state. As the polarity is decreased the transition state is gradually transformed into a double-humped profile that then changes smoothly through a triple-well profile into a single-well profile where the symmetric structure of the transition state is retained. The depth of the well is ca. 16 kcal/mol for n = 2 and reaches 40 kcal/mol for n = 7, resembling the stability of a weak chemical bond. This is traced to electrostatic effects as well as to the effect of an intermediate VB configuration. In the analogous polyynes, a stable adduct is already formed at n = 1. This is attributed to the formation of the relatively stable vinylic carbanion. As the number of acetylene units increases, the vinylic geometry (a CCC angle of 123°) is gradually lost until at n = 5 the adduct attains a linear geometry.

12.
J Phys Chem A ; 117(24): 5023-7, 2013 Jun 20.
Artículo en Inglés | MEDLINE | ID: mdl-23705974

RESUMEN

Computational studies at the B3LYP/6-31+G* level were carried out on the addition of pyridine to polyynes (C6-C18) and on the protonation of polyynes by methyl ammonium fluoride under electric fields of 2.5 and 5 MV/cm. The electric field in each case was oriented along the polyyne axis in a direction that enhances the reaction by stabilizing the incipient dipole. It was found that the reaction of pyridine addition is endothermic with a late transition state. The longer the polyynes and the stronger the field, the electric field catalysis was more efficient. Extrapolation of the data to long polyynes shows that at 1000 nm an electric field of 50 000 V/cm will reduce the barrier by 10 kcal/mol. This reduction is equivalent to 7 orders of magnitude in rate enhancement. A similar barrier reduction could be achieved with a 2.5 MV/cm field at a polyyne length of 20 nm. Protonation reactions were found to be much more affected by the electric field. A reduction of the reaction barrier by 10 kcal/mol using a 2.5 MV/cm electric field could be achieved at a polyyne length of 10 nm. Thus the electric field along the long axis of a substrate could induce a gradient of reactivity which could, in principle, enable the barcoding of substrates by using a sequence of reactants having different reactivities.

13.
J Org Chem ; 77(20): 9199-204, 2012 Oct 19.
Artículo en Inglés | MEDLINE | ID: mdl-23025345

RESUMEN

The reaction of six substrates (diphenylacetylene, benzonitrile, methyl benzoate, phenylacetylene, naphthalene, and 1-chloro-4-ethylbenzene) with SmI(2) in the presence of MeOH or TFE was studied. The reactions were monitored under three different conditions: (a) irradiation, (b) irradiation in the presence of HMPA, and (c) reactions in the presence of HMPA in the dark. The combination of visible light and HMPA was found in some cases to be synergistic, in others to be additive, and in four cases to be inhibitive. The Marcus theory provides a good understanding of the synergistic and the additivity phenomena. The inhibitive effect is traced to the post electron transfer step in which Sm(3+) plays an important role. Once coordinated to HMPA, Sm(3+) is less capable of assisting in the protonation of the radical anion or the expulsion of the leaving group. Ranking according to the substrate's electron affinity shows that inhibition is manifested for the three least electrophilic substrates: phenylacetylene, naphthalene, and 1-chloro-4-ethylbenzene. Typical of these substrates is the short lifetime of their radical anions. Thus, if a step consecutive to electron transfer is slow and cannot compete successfully with the rapid back electron transfer, the benefit of having the electron transfer step enhanced is much reduced.


Asunto(s)
Hempa/química , Yoduros/química , Luz , Samario/química , Acetileno/análogos & derivados , Acetileno/química , Benzoatos/química , Estructura Molecular , Naftalenos/química , Nitrilos/química , Oxidación-Reducción
14.
J Org Chem ; 77(8): 4029-34, 2012 Apr 20.
Artículo en Inglés | MEDLINE | ID: mdl-22468753

RESUMEN

Despite their high electron-withdrawing strength, nitriles are not good electron acceptors and therefore are hard to reduce. In this work, using photostimulation in the visible region, we examined the reactivity of aliphatic and aromatic, mono- and dicyano compounds in reaction with SmI(2). A proton donor that complexes efficiently with SmI(2) must be used. Maximum yield was obtained at ca.0.2 M MeOH. Aromatic nitriles were more reactive than aliphatic nitriles, which exhibited negligible yields. Phenylacetonitrile presents an intermediate reactivity. The mechanism of the reaction involves coordination of the SmI(2) to the lone pair of the nitrile nitrogen followed by an inner sphere electron transfer. Surprisingly, m-dicyanobenzene was less reactive than the monocyano derivative benzonitrile. This was traced to the lower ability of the dicyano compound to coordinate to the SmI(2) due to, as was shown by quantum mechanical calculations, its lone pair having an energy significantly lower than that of benzonitrile. It is noteworthy that at the SmI(2) initial concentration used (0.04M), light penetrates only the 0.4 mm outer layer of the reaction mixture. Therefore the photostimulation effect observed was due to irradiation of only 4% of the total reaction volume, implying that under optimal conditions the effect should be 25 times larger.

15.
Phys Chem Chem Phys ; 14(39): 13684-91, 2012 Oct 21.
Artículo en Inglés | MEDLINE | ID: mdl-22968657

RESUMEN

The current study extends our work with spiropyran-merocyanines (SP-MC) as molecular photoswitches by delving into the effects of viscosity. This has led to the interesting finding of a dichotomy in viscosity dependence. Solutions of SP [6'-nitro-1,3,3-trimethylspiro(indolino-2,2'-benzopyran)] in a wide range of ethylene glycol-methanol (EG-MeOH) media (3.59 to 17.9 M in EG) were irradiated 90 s (365 nm). The absorbance at 90 s of MC (532 nm) formed photolytically varied with solvent. The least viscous medium yielded the highest concentration of MC and yields declined with increasing viscosity. Once irradiation ceased each system achieved thermal equilibrium. Molecular dynamics studies of typical thermal reactions governed by electronic and steric factors show that the transition state is achieved primarily after solvent reorganization has occurred to accommodate the new structure. It follows that in such thermal reactions viscosity may not cause any hindrance to the motion of atoms in molecules because solvent has already rearranged. In contrast, photochemical excitations occur at much higher rates (10(-15) s) than solvent reorganization, i.e. dielectric relaxation (10(-10) to 10(-12) s). The viscosity dependence of photochemical MC formation suggests that a major geometrical change is required for excited SP to be converted to MC. The dichotomy in dependence on viscosity is confirmed by the thermal equilibration of SP and MC. The equilibrium constant for the process increases three-fold (from 0.0535 to 0.158) as the EG content of the medium increases. However, the forward rate constant (SP → MC) is almost invariant with EG content or viscosity. The process is viscosity independent. The increase in the equilibrium constant with EG concentration is a result of a decline in the reverse rate constant for MC cyclisation to SP. This is attributed to special stabilisation of the MC that increases with increasing EG concentration. The present study, to our knowledge, is the first to dissect viscosity from solvent stabilisation factors in SP-MC systems. Further, the study highlights the fundamental difference between photolytic and thermal processes, providing another avenue of control for these SP-MC photoswitches.


Asunto(s)
Benzopiranos/química , Indoles/química , Nitrocompuestos/química , Temperatura , Glicol de Etileno/química , Metanol/química , Simulación de Dinámica Molecular , Estructura Molecular , Procesos Fotoquímicos , Viscosidad
16.
ACS Omega ; 7(44): 40021-40024, 2022 Nov 08.
Artículo en Inglés | MEDLINE | ID: mdl-36385862

RESUMEN

It has been recently shown that SmI2 is more azaphilic than oxophilic. Density functional theory calculations reveal that coordination of 1-3 molecules of ethylenediamine is more exothermic by up to 10 kcal/mol than coordination of the corresponding number of ethylene glycol molecules. Taking into account also hydrogen bonds between ligands and tetrahydrofuran doubles this preference. The intrinsic affinity parallels the order of basicity. The cooperativity with the hydrogen bonding makes SmI2 more azaphilic than oxophilic.

17.
J Am Chem Soc ; 133(37): 14795-803, 2011 Sep 21.
Artículo en Inglés | MEDLINE | ID: mdl-21848327

RESUMEN

The reduction of the three imines, N-benzylidene aniline (BAI), N-benzylidenemethylamine (BMI), and benzophenone imine (BPI), with SmI(2) gives the reduced as well as coupled products. The reactions were found to be autocatalytic due to the formation of the trivalent samarium in the course of the reaction. When preprepared SmI(3) was added to the reaction mixture, the reaction rate increased significantly. However, the kinetics were found to be of zero order in SmI(2). This type of behavior is typical of surface catalysis with saturation of the catalytic sites. Although no solids are visible to the naked eye, the existence of microcrystals was proven by light microscopy as well as by dynamic light scattering analysis. Although HRTEM shows the existence of quantum dots in the solid, we were unable to make a direct connection between the existence of the quantum dots and the catalytic phenomenon. In the uncatalyzed reaction, the order of reactivity is BPI > BMI > BAI. This order does not conform to the electron affinity order of the substrates but rather to the nitrogen lone pair accessibility for complexation. This conclusion was further supported by using HMPA as a diagnostic probe for the existence of an inner sphere electron transfer reaction.

18.
J Org Chem ; 76(22): 9438-43, 2011 Nov 18.
Artículo en Inglés | MEDLINE | ID: mdl-22004459

RESUMEN

The reduction of azobenzene by SmI(2) in THF to give hydrazobenzene was investigated. The kinetics are first order in the substrate and first order in SmI(2). The kinetic order in MeOH is ca. 0.56, and in TFE it is ca. 0.2. The fractional order in the proton donors is interpreted as being a result of their acting in two opposing manners. In one the proton donor enhances the reaction by protonation of the radical anion, and in the other it slows the reaction by binding to the lone pair electrons of the nitrogen in the azobenzene. This hampers the fast inner-sphere electron-transfer mode. Experiments conducted in the presence of low concentrations of HMPA show rate enhancement suggesting that the SmI(2), which is partly coordinated to HMPA molecules, has some free sites to bind to the substrate. When more HMPA is added, it prevents the fast inner-sphere mechanism and the rate decreases. In this system, the increase in the reduction potential of SmI(2) caused by HMPA is similar to the rate enhancement by an inner sphere mechanism. In general, the replacement of a skeletal carbon by a nitrogen atom causes a significant rate enhancement.

19.
J Org Chem ; 76(5): 1355-60, 2011 Mar 04.
Artículo en Inglés | MEDLINE | ID: mdl-21250712

RESUMEN

The endo/exo product ratio in the reactions of SmI(2) with norcamphor in the presence of various proton donors was determined. The effect of MeOH, EtOH, trifluoroethanol (TFE), ethylene glycol (EG), and water was investigated at various concentrations of these proton donors. At low concentrations of EtOH, TFE, and EG, an endo/exo ratio near unit was found. This ratio increased as the concentration of the proton donor increased. However, MeOH and water gave a U-type curve, in a plot of the endo/exo ratio vs proton donor concentration. The difference between the two groups of proton donors was shown not to result from differences in their acidities or polarity effects. It is suggested that at low MeOH and water concentrations, the second electron transfer takes place from the dimer of SmI(2) rather than from the monomer. This bulky electron donor approaches the radical anion preferentially from the exo direction giving rise to the high endo/exo ratio at the left arm of the U-shaped curve. Comparison of kinetic and product H/D isotope effects shows that protonation on carbon, the step that locks the stereochemistry, is a post rate determining step.


Asunto(s)
Etanol/química , Glicol de Etileno/química , Metanol/química , Norbornanos/química , Protones , Trifluoroetanol/química , Oxidación-Reducción , Estereoisomerismo , Agua/química
20.
J Am Chem Soc ; 131(23): 8280-4, 2009 Jun 17.
Artículo en Inglés | MEDLINE | ID: mdl-19507909

RESUMEN

The reduction of a series of alpha-cyanostilbenes with SmI(2) was studied in THF in the presence of various proton donors. No reaction occurred in the presence of the alcohols TFE, i-PrOH and t-BuOH. In the presence of MeOH, water and ethylene glycol the reactions occurred; however in the presence of water and ethylene glycol they were too fast for kinetic determinations (tau(1/2) < 1 ms). Reactions with MeOH were first order in SmI(2) and first order in the substrate. The order in MeOH varies as a function of its concentration and the plot of log k vs log [MeOH] is sigmoidal. Comparison of the kinetic isotope effect and the incorporation isotope effect suggests that, counterintuitively, protonation of the radical anion takes place on the carbon beta to the cyano group. It is concluded that proton donors that form complexes with SmI(2) expand the range of substrates that can be reduced by SmI(2). This is due to their proximity to the radical anion as it is formed. This short-lived radical anion cannot be efficiently trapped by a proton donor from the bulk medium. A protocol is herein suggested as to when proton donors which complex to SmI(2), e.g. MeOH, water and ethyleneglycol should be used, and when it is recommended to use noncomplexing proton donors, e.g. TFE, i-PrOH and t-BuOH, to induce reaction.

SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA