Tetrahedron
Volume 61, Issue 22,
30 May 2005
, Pages 5147-5208
Author links open overlay panel,
Graphical Abstract

Introduction
As mentioned in introductory organic chemistry textbooks, the ideal geometry for a double bond has the olefinic carbon atoms and the four atoms connected to them in the same plane. Thus, in the ground state of ethylene, all six atoms lie in one plane, the bond angles are near 120° and the carbon–carbon distance is 1.34Å. However, as stressed by Mislow as early as 1965, ‘planarity is not expected if the molecules does not have a plane of symmetry passing through the sp2 carbon atoms and all four corresponding ligands’.1 Cyclopentene is a good example, ab initio calculations showing that, when the molecule is allowed to adopt its equilibrium envelope conformation, the sp2 carbon atoms are pyramidalized.2 Pyramidalized alkenes are molecules containing carbon–carbon double bonds in which one or both of the doubly bonded carbon atoms do not lie in the plane defined by the three atoms attached to it. Deviations of the planarity for the vast majority of olefinic carbon atoms are small. However, when a carbon–carbon double bond is located at the bridgehead positions of a polycyclic structure, severe deviations of the usual planar geometry occur.3
In bridgehead alkenes, there are two different types of distortions, the out-of-plane bending and the in-plane bending. The latter is the major distortion observed in small cyclic alkenes (e.g., cyclopropene).4 In bridgehead alkenes, the out-of-plane bending is much more important and two major modes of distortion can be distinguished: twisting and pyramidalization.5 One extreme case is the pure twisting: the two olefinic carbon atoms stay fully sp2 hybridized and thus planar (Fig. 1a). As a consequence, the two p orbitals are misaligned, which weakens the π-component of the double bond. This is visualized by the twisting angle, τ, which is defined as the dihedral angle between the two p orbitals. In the other extreme situation, the syn pyramidalization, the carbon atoms are rehybridized by admixture of additional p character into the original sp2 σ-bonds; this makes the geometry around the carbon non-planar. The π bond is now formed from two p-orbitals with some added s-character; the alignment between the two orbitals is optimal (τ=0), but their orientation in the p-plane is no longer parallel and, for this reason, the distance between them increases and the net overlap is smaller. As we will discuss below, there is now an out-of-plane or flap angle, Ψ, defined as the dihedral angle between a plane containing two cis-substituents and the two olefinic carbon atoms and the plane containing the two other cis-substituents and the two olefinic carbon atoms and this is usually referred to as its supplementary ζ=180°−Ψ, (Fig. 1b). Usually, the situation encountered in anti-Bredt alkenes is intermediate: twisting and pyramidalization occur simultaneously (Fig. 1c).
Another form of pyramidalization, in which the π-type spn orbitals are oriented towards opposite sides of the double bond, is called anti pyramidalization. Although calculations have shown that anti pyramidalization of the carbon atoms of the double bond is energetically more favorable than syn pyramidalization,6 it is much less common and will only be mentioned briefly here.
Bridgehead olefins with torsionally strained double bonds (‘twisted or anti-Bredt alkenes’) have been studied extensively,7 but olefins in which the carbon atoms forming the double bonds are pyramidalized (‘pyramidalized alkenes’) have received much less attention. In this report, we deal with the synthesis and reactivity of untwisted, but pyramidalized, bridgehead olefins. Where appropriate, a mention of the physical and chemical properties will be made. Twisted alkenes, distorted aromatic compounds, such as cyclophanes, fullerenes and fullerene-related compounds are out of the scope of this report.
Some reviews on pyramidalized alkenes have appeared,8 the more comprehensive being the excellent survey by Borden in 1989,8b although more concise accounts have appeared later.9 The present report will cover the material that has appeared since 1989 up to mid 2004, although some previous aspects covered by Borden will be mentioned here.
As discussed in more detail later, pyramidalization changes the typical chemistry of the double bond dramatically. Dimerization of highly pyramidalized alkenes occurs rapidly, and pyramidalized alkenes that do not dimerize at room temperature react with atmospheric oxygen10 and with nucleophilic reagents.11
Pyramidalization allows the 2s atomic orbitals of the olefinic carbon atoms to be mixed into the π bond. The increase in 2s character stabilizes both the π and the π* molecular orbitals. However, this rehybridization also decreases the overlap of the two hybrid orbitals, as compared to the overlap of two parallel p atomic orbitals, by pointing the large lobes of the hybrid orbitals away from each other. The loss of the bonding overlap in the π molecular orbital destabilizes it. The effects of increased 2s character and reduced overlap tend to cancel, so that the energy of the π molecular orbital remains relatively constant. In contrast, a reduction of antibonding overlap in π* stabilizes this molecular orbital further. Hence, the π* molecular orbital drops rapidly in energy upon pyramidalization. Therefore, excitation of an electron from π to π* is made energetically less costly on increasing pyramidalization.12 The unusually low energy of π* is responsible for the ease of reduction of pyramidalized alkenes, for the ability of pyramidalized alkenes to form stable complexes with transition metals, such as Pt(0), that can donate electron density into this molecular orbital,13 and for the deshielding observed for the olefinic carbon atoms in 13C NMR.14
Section snippets
Measurements of pyramidalization
In studying pyramidalized alkenes, it is convenient to use a geometrical parameter to measure pyramidalization.15 More than 20 years ago, Borden et al. introduced the pyramidalization angle, Φ, as a measure of pyramidalization.12a Strictly speaking, the Φ angle is applicable only to those cases having C2v symmetry, with a mirror plane bisecting and perpendicular to the double bond and a mirror plane containing the double bond. As shown in Figure 2a, Φ is the angle between the plane containing
9,9′-Didehydrodianthracene
The title compound 3 has historic interest, because it was the first substantially pyramidalized alkene to be synthesized and it offered the first experimental evidence that pyramidalized alkenes are suceptible to nucleophilic addition reactions, as predicted by frontier orbital theory. Although Applequist et al. had suggested the formation of 3 by the reaction of 9-bromodianthracene 4 with strong bases,18 it was not until 1968 that Weinshenker and Greene described the successful isolation of 3.
Sesquinorbornenes and related alkenes
During the 1980s, the groups of Bartlett, Paquette and Watson synthesized and fully characterized several derivatives of syn- and anti-sesquinorbornene, 84 and 85, respectively, and syn- and anti-sesquinorbornadiene, 86 and 87, respectively (Fig. 3). Most of these derivatives are fairly stable compounds and several X-ray studies were carried out.
Although there are a few exceptions, near-planar alkene geometries have been the rule in the derivatives of 85 and 87. For the syn-sesquinorbornenes,
Bicyclo[3.3.0]oct-1(5)-ene derivatives
In the 1980s, Houk12, 17, 86 and Burkert,87 independently, proposed that pyramidalization occurs to relieve unfavorable torsional interactions, favoring staggering of the bonds at adjacent carbon atoms.9b–c Consistent with this proposal, a pyramidalized, C2v syn geometry of bicyclo[3.3.0]oct-1(5)-ene 201 was calculated to be lower in energy than a planar, C2h anti geometry (Fig. 9).9, 12
Ab initio calculations (HF/3-21G) carried out by Hrovat and Borden predicted a pyramidalization angle of 3.6°
Cubene
Cubane, pentacyclo[4.2.0.02,5.03,8.04,7]octane, is one of the most exciting and extensively studied cage compounds.138 Eaton and co-workers have worked for years on the synthesis of a myriad of cubane derivatives, including the highly pyramidalized 1,2-dehydrocubane (‘cubene’) 433.11, 117
According to ab initio calculations, cubene, with a pyramidalization angle of Φ=84.1° (HF/3-21G), is the most highly pyramidalized alkene yet known.12c Calculations also indicate that sufficient overlap exists
Unsaturated quadricyclanes
The facile access to tetracyclo[3.2.0.02,7.04,6]heptane (quadricyclane) through a photochemically induced [2+2] cycloaddition of the commercially available 2,5-norbornadiene149 made this framework very attractive for the installation of unusual double bonds (Fig. 18).150 As a result of some impressive work, Szeimies and co-workers have found evidence for three unsaturated quadricyclanes, 1(7)-quadricyclene 494,151 1(5)-quadricyclene 495,152 and 1(2)-quadricyclene, 496.153 The elusive anti-Bredt
Anti-pyramidalized alkenes
Anti-pyramidalization is calculated to be energetically less costly than syn-pyramidalization. However, few studies have been carried out in the area of anti-pyramidalized alkenes and, most importantly, there is still a lack of highly anti-pyramidalized alkenes.
According to its X-ray crystal structure and DFT theoretical calculations, heptafulvalene 688 has an anti-pyramidalized C2h structure. This conformation is lower in energy than the syn-pyramidalized C2v conformation.194
Probably, the best
Perspectives
Fifteen years have passed since the publication of the seminal review on pyramidalized alkenes by Borden in Chemical Reviews in 1989.8b In this period, many landmark achievements have been attained. The synthesis of several derivatives of the highly pyramidalized tricyclo[3.3.0.03,7]oct-1(5)-ene, the synthesis of polyunsaturated dodecahedranes, the generation of acepentalene and related compounds, the dimerization of cubene and the synthesis of cyclopropene-fused norbornene derivatives are
Acknowledgements
Financial support from Ministerio de Ciencia y Tecnologia and FEDER (Project No. PPQ2002-01080 and Ramón y Cajal fellowship to S. V.) and Comissionat per a Universitats i Recerca (Project No. 2001SGR00085) is gratefully acknowledged. The authors wish to express their deep gratitude to our co-workers, who have contributed significantly over the past 20 years with hard work, skill and enthusiasm to our contributions to the field of pyramidalized alkenes. We also thank the Centre de
Santiago Vázquez was born in Barcelona in 1968. He studied Pharmacy (1986–1991) at the Universitat de Barcelona. He obtained his PhD in Organic and Medicinal Chemistry at the same university in 1996 under the direction of Professor P. Camps. After spending 2 years (1998–1999) in the Christopher Ingold Laboratories (University College London) with Professor William B. Motherwell as a Marie Curie Research Fellow, he returned to Barcelona. In 2001 he took up his present position as ‘Investigador
First page preview
Click to open first page preview
References and notes (205)
- K.B. Wiberg et al.
J. Am. Chem. Soc.
(1984)
P.D. Bartlett et al.J. Org. Chem.
(1991)
J.M. Smith et al.J. Am. Chem. Soc.
(1993)
- S. Kammermeier et al.
Angew. Chem., Int. Ed. Engl.
(1996)
- N. Treitel et al.
Angew. Chem., Int. Ed.
(2003)
- G. Dyker et al.
Angew. Chem., Int. Ed. Engl.
(1995)
G. Dyker et al.Tetrahedron
(1996)
- T. Tsuji et al.
Chem. Commun.
(1997)
T. Tsuji et al.J. Am. Chem. Soc.
(1995)
- L.A. Paquette et al.
J. Am. Chem. Soc.
(1989)
L.A. Paquette et al.J. Am. Chem. Soc.
(1990)
- H. Can et al.
Eur. J. Org. Chem.
(2003)
R. Özen et al.See Also5.5: ホモリシス切断と結合解離エネルギーChemistry - Oxidation state of the sulfur atoms in the thiosulfate IonA new antagonism between hindered amine light stabilizers and acidic compounds including phenolic antioxidant6.8: Describing a Reaction - Bond Dissociation EnergiesJ. Chem. Crystallogr.
(2004)
- T. Kitamura et al.
J. Org. Chem.
(1999)
- M. Balci et al.
J. Chem. Crystallogr.
(1995)
- P.C. Kearney et al.
J. Am. Chem. Soc.
(1993)
J.C. Ma et al.Chem. Rev.
(1997)
Introduction to Stereochemistry
(1965)
J. Am. Chem. Soc.
(1992)
Top. Stereochem.
(1991)
J. SandströmI.V. KomarovRuss. Chem. Rev.
(2001)
H. Dodziuk(1995)
J. Am. Chem. Soc.
(1979)
J. Am. Chem. Soc.
(1973)
G.W. Wijsman et al.J. Org. Chem.
(2001)
Tetrahedron
(1980)
J.R. Wiseman et al.J. Am. Chem. Soc.
(1970)
Angew. Chem., Int. Ed. Engl.
(1971)
R. Keese et al.Angew. Chem., Int. Ed. Engl.
(1972)
P.M. WarnerChem. Rev.
(1989)
W. Grimme et al.Synlett
(1998)
P. Roach et al.Angew. Chem., Int. Ed.
(2003)
Chimia
(1981)
W.T. BordenChem. Rev.
(1989)
Synlett
(1996)
V.S. Mastryukov et al.Struct. Chem.
(2000)
V.S. Mastryukov et al.J. Phys. Chem. A
(2001)
H. HopfClassics in Hydrocarbon Chemistry
(2000)
J. Am. Chem. Soc.
(1988)
J. Schäfer et al.Tetrahedron Lett.
(1988)
J. Am. Chem. Soc.
(1979)
K.N. Houk et al.Isr. J. Chem.
(1983)
D.A. Hrovat et al.J. Am. Chem. Soc.
(1988)
J. Am. Chem. Soc.
(1990)
K. Morokuma et al.J. Am. Chem. Soc.
(1991)
A. Nicolaides et al.Organometallics
(1995)
J. Uddin et al.Organometallics
(1999)
B.F. YatesJ. Organomet. Chem.
(2001)
J. Chem. Soc., Perkin Trans. 2
(2002)
Struct. Chem.
(1991)
J. Am. Chem. Soc.
(1990)
R.C. HaddonJ. Am. Chem. Soc.
(1990)
R.C. HaddonScience
(1993)
J. Am. Chem. Soc.
(1974)
W.H. Watson et al.J. Am. Chem. Soc.
(1981)
R. Gleiter et al.Tetrahedron Lett.
(1982)
J. Spanget-Larsen et al.Tetrahedron
(1983)
K.N. Houk et al.J. Am. Chem. Soc.
(1983)
(f) See also Refs. 10,...J. Am. Chem. Soc.
(1959)
J. Am. Chem. Soc.
(1968)
J. Am. Chem. Soc.
(1974)
Angew. Chem., Int. Ed. Engl.
(1994)
Liebigs Ann.
(1995)
Liebigs Ann.
(1996)
Liebigs Ann./Recueil
(1997)
Angew. Chem., Int. Ed. Engl.
(1997)
Tetrahedron Lett.
(1964)
Nature
(2003)
T. Kawase et al.Angew. Chem., Int. Ed.
(2004)
Angew. Chem., Int. Ed. Engl.
(1997)
Angew. Chem., Int. Ed. Engl.
(1996)
R. Herges et al.Chem. Phys. Lett.
(2000)
Angew. Chem., Int. Ed.
(2003)
Org. Lett.
(2003)
Acta Crystallogr. Sect. C
(1988)
J. Org. Chem.
(1985)
Tetrahedron
(1986)
M.E. Jason et al.Inorg. Chem.
(1975)
J. Casanova et al.J. Am. Chem. Soc.
(1978)
V.V. Burlakov et al.Chem. Commun.
(2004)
J. Am. Chem. Soc.
(1984)
E. Honegger et al.J. Electron Spectra Relat. Phenom.
(1983)
J. Am. Chem. Soc.
(1996)
Y. He et al.J. Am. Chem. Soc.
(2003)
J. Am. Chem. Soc.
(1978)
K.B. Wiberg et al.Isr. J. Chem.
(1983)
H.A. BentChem. Rev.
(1961)
J. Phys. Chem. A
(2003)
Org. Lett.
(2003)
Strained Organic Molecules
(1978)
Cited by (79)
How long a C–C bond can be? An example of extraordinary long C–C single bond in 1,2-diarylamino-o-carborane
2019, Chinese Chemical Letters
How long a C
C bond can be? A question has long fascinated chemists. This work reports an example of extraordinary long C
C bond distance of 1.990(4) Å observed in single-crystal X-ray structure of 1,2-(NHMes)2-o-carborane (2; Mes = 2,4,6-trimethylphenyl). DFT calculations show that hyperconjugation of lone pairs of the nitrogen atoms into the empty σ* orbital of the cage C
C bond is the origin of the bond elongation. Such hyperconjugation can be suppressed if the two nitrogen atoms in 2 are linked to a Lewis acidic germanium (II) center.
Quantum chemical exploration of new Π-electron systems: Capsule-formed dimers of polycyclic aromatic hydrocarbons
2019, Chemical Physics Letters
We have found capsule-formed dimers of polycyclic aromatic hydrocarbons (cap-di-PAH) for coronene, ovalene, and circumcoronene by quantum chemical calculations. Two monomers are connected at the edge that gives a capsule form, holding the aromatic characters in the inner six-membered rings. The aromatic π-electron networks at the upper and lower parts of the dimers are not merged each other. Those cap-di-PAHs are located in sufficiently deep potential wells surrounded by high energy barriers, indicating the high possibility of existence.
Construction of carbocycles initiated by Cu-catalyzed radical reaction of Cl<inf>2</inf>C(CN)<inf>2</inf>
2017, Tetrahedron
A Cu-catalyzed radical reaction of Cl2C(CN)2 was utilized for stereoselective conversion of unsaturated molecules into functionalized carbocycles. Chloromalononitrile radicals, generated by treating Cl2C(CN)2 with catalytic amounts of CuCl and dppf in refluxing dioxane, intermolecularly reacted with the unsaturated bonds of acyclic or 10-membered compounds. The resultant carboradicals then intramolecularly added to another unsaturated bond to produce 1,2-disubstituted cyclopentane or trans-decalin derivatives. The chemo- and stereoselectivities of these radical cascade reactions are discussed in detail.
Synthesis of a novel diene from a cyclobutane precursor: An entry to 2,9-disubstituted [2]diadamantanes
2013, Tetrahedron
Citation Excerpt :
Noradamantene (1, n=1), the second member of the homologous series of pyramidalized alkenes 1,1,2 dimerizes to the heptacyclic cyclobutane derivative 2, n=1, even at temperatures as low as −78 °C.3 We are interested in using noradamantene as a building block for the heptacyclic ketone 3, a potential precursor for the interesting carbene 4.4 For this reason we studied experimentally and the formation of diene 5, n=1 from cyclobutane 2, n=1.
In the gas-phase thermolysis (350°C) of heptacyclic cyclobutane 2, n=1, (the [2+2] dimer of noradamantene) two isomeric dienes (the symmetric 5, n=1 and the asymmetric 6a) and the reduction product [2]diadamantane (7-H) are formed. Computations suggest that at 350°C a diradical intermediate (8, n=1, derived from C–C bond cleavage of 2, n=1) gives rise to 5, n=1. Diradical 8, n=1 is also likely to abstract allylic hydrogens from 5, n=1 affording 7-H, and also to isomerize 5, n=1 to give the diene 6b from which 6a can result as the more stable diene of a Cope rearrangement. The gas-phase thermolysis of 2, n=1, in the presence of iodine affords 7-H at 350°C and the new compound 2,9-diiodo-[2]diadamantane (7-I2) at 150°C. 7-I2 can be easily prepared at room temperature by the iodination of 2, n=1 in CH2Cl2 and affords in a synthetically useful way diene 5, n=1 by reduction with Na/Hg at room temperature.
Cross metathesis of several methylenecyclopentane derivatives
2013, Tetrahedron
The cross metathesis (CM) of several methylenecyclopentane derivatives using Hoveyda–Grubbs second generation catalyst 4 (5–10mol%) has been studied. Medium to good yields of tetrasubstituted alkenes have been obtained. In the case of 8-methyl-2,5-dimethylene-2,3,5,6-tetrahydro-1H,4H-3a,6a-(methanoiminomethano)pentalene-7,9-dione 2 products from single, double and triple CM were formed. With 8-methyl-5-methylene-5,6-dihydro-3a,6a-(methanoiminomethano)pentalene-2,7,9(1H,3H,4H)-trione 3 a good yield of the CM product was obtained working at 140°C in xylene for 3d, showing the high thermal stability of this catalyst. In the CM of diene 2 and enone 3, the main products were always the anti-stereoisomers.
Tricyclo[3.3.1.0<sup>3,7</sup>]nonane-3,7-diyl bis(4- methylbenzenesulfonate)
2013, Acta Crystallographica Section E: Structure Reports Online
Recommended articles (6)
Research article
One pot synthesis of biologically active pregnane derivatives, their single crystal structures, spectroscopic characterization and theoretical calculations
Journal of Molecular Structure, Volume 1052, 2013, pp. 76-84
One pot allylic oxidation of 3β-acetoxypregna-5,16-diene-20-one (2) and nucleophilic addition at C-16 position of 3β-hydroxypregna-5,16-diene-20-one (3) yielded 3β-acetoxypregna-5,16-diene-7,20-dione (4) and 3β-hydroxy-16α-(5′-hydroxypentyloxy)-pregn-5-ene-20-one (5) respectively in high yield. A detailed theoretical study supported by X-ray analysis of compounds 4 and 5 has been carried out. Conformational analysis of compounds 4 and 5 was done with the help of crystal structure, which crystallize out in orthorhombic form having P212121 space group. Structural characterization of compounds 4 and 5 was done with the aid of 1H, 13C NMR, IR, UV, ESI-MS and ESI-HRMS. The molecular geometries and vibrational frequencies for compounds 4 and 5 in the ground state were calculated using the Density functional theory (DFT) with 6-31G(d,p) basis set and compared with experimental data. 1H and 13C nuclear magnetic resonance magnetic shifts of 4 and 5 were calculated using GIAO method and compared with the experimental data. UV–Vis spectra of both the compounds were recorded and electronic properties such as HOMO–LUMO energies were calculated by time dependent TD-DFT approach. The compounds were screened for their anti-hyperlipidemic and anti-oxidant activity.
Research article
Eu3+/Sm3+ hybrids based with 8-hydroxybenz[de]anthracen-7-one organically modified mesoporous silica SBA-15/16
Solid State Sciences, Volume 50, 2015, pp. 9-17
A series of organic–inorganic hybrid materials were prepared by linking lanthanide (Eu3+, Sm3+) complexes to mesoporous SBA-15/SBA-16 through 8-hydroxybenz[de]anthracen-7-one modified silane (HBA-Si) as linker. The physical characterizations of these hybrids revealed that they all have high surface area, uniformity in mesostructure. The luminescence properties of these covalently bonded materials (denoted as Ln(HBA-SBA-15)3phen and Ln(HBA-SBA-16)3phen) were compared with ternary complexes (Ln(HBA)3phen) (Ln=Eu, Sm). Eu(HBA-SBA-15(16))3phen hybrids display better thermal stability, whose luminescent lifetimes and quantum efficiencies were matchable with those of Eu(HBA)3phen complex in spite of its much lower effective condensation of Eu3+ species. In addition, the luminescent performance of functionalized SBA-15 hybrids was more favorable than that of functionalized SBA-16 hybrids, revealing that SBA-15 was a better host material for lanthanide complex than mesoporous silica SBA-16.
Research article
Comment on “A new equation based on ionization energies and electron affinities of atoms for calculating of group electronegativity” by S. Kaya and C. Kaya [Comput. Theoret. Chem. 1052 (2015) 42–46]
Computational and Theoretical Chemistry, Volume 1083, 2016, pp. 72-74
Kaya and Kaya’s equation for calculating group electronegativity by global equalization of atomic electronegativities is neither new, nor reliable. Critical comments are presented concerning: (i) the dramatic failure of global ground-state electronegativity equalization (ENE) and the reasons thereof, (ii) the importance of the Wigner–Witmer symmetry laws to obtain accurate ENE in localized bonds, (iii) the electron self-interaction error causing unreasonably small partial charges, (iv) the fact that the “new equation” and global ENE generally yield identical partial charges on all atoms-in-the-molecule of the same element independent of their neighbors, (v) the electron donating power of alkyl groups. Tools and models to correct for the failures and improve the accuracy of local ENE and the resulting partial atomic charges are indicated.
Research article
Role of free-radical chain reactions and silylene chemistry in using methyl-substituted silane molecules in hot-wire chemical vapor deposition
Thin Solid Films, Volume 635, 2017, pp. 42-47
The reaction chemistry of four methyl-substituted silane molecules, including monomethylsilane (MMS), dimethylsilane (DMS), trimethylsilane (TriMS), and tetramethylsilane (TMS), has been studied in the hot-wire (catalytic) chemical vapor deposition (HWCVD) process, both on heated tungsten or tantalum filaments and in the gas phase. The four molecules dissociate catalytically on hot W or Ta surfaces to form ·CH3, ·H and H2. In a HWCVD reactor setup where the pressure is relatively high at 8–533Pa, two reaction mechanisms involving silylene and free radical intermediate, respectively, operate with the four precursor gases. For MMS, its reaction chemistry is characterized by the exclusive involvement of methylsilylene intermediate. Reaction products when using DMS as a precursor gas come from silylene chemistry and free-radical chain reactions. Filament temperature and pressure affect strongly the competition between the two mechanisms. Specifically, silylene chemistry dominates at low temperatures of 1200–1300°C and a low pressure of 16Pa, and free-radical chain reactions take over at high temperatures and pressures. With the increasing methyl substitutions in TriMS and TMS, free-radical chain reactions become predominant. A clear transition in the operating reaction mechanism from silylene chemistry to free-radical chain reactions has been shown when the number of Si
CH3 bonds increases in the methyl-substituted silane molecules.
Research article
Off–on switchable chemosensor based on rhodamine armed with morpholine moiety
Journal of Luminescence, Volume 168, 2015, pp. 283-287
We have synthesized a novel morpholine functionalized rhodamine derivative RBM which specifically binds to Fe3+ in the presence of large excess of other competing metal ions. The chemosensor is highly selective for Fe3+ over other cations. Meanwhile, the distinct color changes and rapid switch-on fluorescence also provide “naked eyes” detection for Fe3+ over a broad pH range. The chemosensor also displays 1:1 complex formation with Fe3+ ion, with a detection limit of 2.1µM in aqueous CH3CN solution, showing that it may offer potential as a chemosensor for the detection of submillimolar Fe3+ ions in physiological environments.
Research article
Synthesis of hydrogenated 2-benzazepin-1-ones by the addition of aryllithium intermediates to isocyanates
Tetrahedron: Asymmetry, Volume 26, Issues 5–6, 2015, pp. 276-280
Two routes for the synthesis of (2-bromophenyl)propionaldehyde derivatives 5 and 8–10 were followed. In the first route, aldehyde 4 was homologated by a Wittig reaction and the resulting α,β-unsaturated acetal 5 was hydrogenated to give 8. According to the second route a Heck reaction of bromoiodobenzene 6 with allyl alcohol and subsequent acetalization provided the propionaldehyde acetals 8–10. After halogen/metal exchange at 5, 8, and 9, the addition to homochiral isocyanate 11 provided the enantiomerically pure benzamides 12–14. Depending on the reaction conditions dihydro- and tetrahyro-2-benzazepines 15 and 16 were prepared, which are potential building blocks for the synthesis of differently substituted homochiral 2-benzazepines.
Santiago Vázquez was born in Barcelona in 1968. He studied Pharmacy (1986–1991) at the Universitat de Barcelona. He obtained his PhD in Organic and Medicinal Chemistry at the same university in 1996 under the direction of Professor P. Camps. After spending 2 years (1998–1999) in the Christopher Ingold Laboratories (University College London) with Professor William B. Motherwell as a Marie Curie Research Fellow, he returned to Barcelona. In 2001 he took up his present position as ‘Investigador Ramón y Cajal’ at Universitat de Barcelona. His scientific interests include polycyclic cage compounds, drug synthesis and free radical and computational chemistry.
Pelayo Camps graduated in chemistry in the Universitat de Barcelona (UB). His PhD studies were carried out at the Organic Chemistry Institute (CSIC, Barcelona) under the direction of Prof. Dr. José Pascual, completing his doctoral thesis and obtaining the PhD degree from the same University in 1972. This year started his docent carrier at the Universitat Autònoma de Barcelona (UAB) as Assistant Professor. In 1978, after a postdoctoral stage at the University of Aix-Marseille-III with Prof. Dr. José Elguero, he was promoted to Associate Professor of the Faculty of Pharmacy of the Universidad de Valencia (UV), where he stayed for a year, returning to the UAB in 1979. Two years later obtained a Full Professor position in the Faculty of Chemistry (San Sebastian) of the Universidad del País Vasco where he spent 2 years. After five more years in the Faculty of Pharmacy of the UV, in 1988, he moved to his actual position as a Full Professor in the Faculty of Pharmacy of the UB where he is currently the Head of the Pharmaceutical Chemistry Unit. His research interest has been always related with the organic synthesis of different kind of compounds: polycyclic and cage compounds via highly pyramidalized alkenes, acetylcholinesterase inhibitors for the treatment of Alzheimer's disease, and the use of chiral auxiliaries for the asymmetric synthesis of drugs and related compounds. He has previously held a visiting professorship at the University of Bordeaux.
Copyright © 2005 Elsevier Ltd. All rights reserved.