Next Article in Journal
An Advanced Approach for MgZnAl-LDH Catalysts Synthesis Used in Claisen-Schmidt Condensation
Next Article in Special Issue
Iminosugar-Phosphines as Organocatalysts in the [3 + 2] Cycloaddition of Allenoates and N-Tosylimines
Previous Article in Journal
In Situ H2 Reduction of Al2O3-Supported Ni- and Mo-Based Catalysts
Previous Article in Special Issue
Recent Advances in the Synthesis of Five-Membered Cyclic Carbonates and Carbamates from Allylic or Propargylic Substrates and CO2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Organic Chemistry and Synthesis Rely More and More upon Catalysts

1
Institute of Chemical Sciences and Engineering, Swiss Institute of Technology in Lausanne (EPFL), 1015 Lausanne, Switzerland
2
Department of Chemistry and Biochemistry, University of California (UCLA), Los Angeles, CA 90095-1569, USA
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(7), 758; https://doi.org/10.3390/catal12070758
Submission received: 1 June 2022 / Revised: 20 June 2022 / Accepted: 21 June 2022 / Published: 8 July 2022
(This article belongs to the Special Issue Commemorative Issue in Honor of Professor Pierre Vogel)

Abstract

:
A few months before the COVID-19 pandemic, Pierre Vogel and Kendall N. Houk published with a new textbook Wiley-VCH, “Organic Chemistry: Theory, Reactivity, and Mechanisms in Modern Synthesis”, with a foreword from the late Roberts H. Grubbs. The book demonstrates how catalytic processes dominate all fields of modern organic chemistry and synthesis, and how invention combines thermodynamics, kinetics, spectroscopy, quantum mechanics, and thermochemical data libraries. Here, the authors present a few case studies that should be of interest to teachers, practitioners of organic and organometallic chemistry, and the engineers of molecules. The Vogel–Houk book is both textbook and reference manual; it provides a modern way to think about chemical reactivity and a powerful toolbox to inventors of new reactions and new procedures.

1. Introduction

Our planet and its people face new problems that require new solutions. Many of us are concerned by climate change, by the loss of biodiversity, by human population growth, by the ageing of people, by neurodegenerative diseases and cancer that are expanding, by the threat of new infectious diseases recently demonstrated so vividly, by the continuously increasing cost of medical care in general, by drinking water availability and its contamination, by air and ocean pollution, and by the degradation of the soil that nourishes us. These problems should not make us fall into a depressed funk. Most scientists are stimulated by these problems and can propose solutions that represent new opportunities for engineers, investors, and business. In the search for these solutions, new molecules and new devices are invented. The first industrial commodity produced by organic chemistry was Perkin’s purple (mauveine) in 1856 [1,2,3]. The first synthetic drug was chloral hydrate (Cl3CCH(OH)2 introduced on the market in 1869, a compound prepared first by Liebig in 1832 (chlorination of ethanol) [4,5,6]. At that time, all dyestuffs and drugs were extracted from minerals, plants, and animals [7,8,9]. Only 300 organic compounds were known in 1800; the first edition of the Beilstein Handbook of Organic Chemistry described 20,000 organic compounds in 1882. Now, more than 80 million organic compounds are known. Nowadays, most dyestuffs are synthetic; most drugs are chemicals obtained through synthesis or are biologics derived from cell cultures [10,11]. At the end of the 19th century, most materials and commodities were made from minerals, plants, wood, animals, and metals. Today many of them are made out of polymers derived from fossil carbon. Sooner or later, engineers will find economical ways to obtain these commodities, and better ones, from left-overs of agriculture and forestry, and from urban waste [12]. The molecules and devices we need now are more and more sophisticated and must be invented at a faster and faster pace. Scientists are exploring a larger and larger chemical space; they are engineers of the molecules who understand their structures from atoms, their dynamics (how they can be transformed), and the way they generate assemblies (macromolecules).
With increasing molecular complexity and sophistication of the procedures available, the study of chemistry represents a fantastic challenge. Scientists with experience and especially teachers are responsible for finding suitable models and explanations the beginners can apply as learning tools. This is done in a beautiful manner in the latest textbooks of organic chemistry [13,14,15] and for transition metal-catalyzed reactions in the book of Hartwig and colleagues [16]. In our textbook, we intend to complement these manuals critically, reviewing selected fundamental concepts and theories and using them to discuss a selection of reactions and processes that have been pivotal in the development of concepts and procedures, and which will hold even greater importance for sciences of matter in the future [17]. This textbook is a guide to quantitative and qualitative models that rely upon thermochemical data and kinetics, and shows how to grasp reactivity and molecular complexity such as asymmetric synthesis and catalysis. Here, we shall not summarize the whole book, but present a few study cases.

2. Invention Is Quicker if Physical Organic Chemistry Principles Are Integrated in Preparative Chemistry

The study of physical organic chemistry is a prerequisite for becoming a modern molecular scientist [18,19]. The main feature of physical organic chemistry is to associate chemical reactions that define chemical functions with quantitative measurements (equilibrium constants, heat of reaction, rate constants, isotopic effects). Chemical informatics estimate that 166.4 billion possible molecules can be made with up to 17 atoms of C, N, O, S, and halogens based on known or estimated stability of compounds and their possible reactions (known feasibility) [20]. When criteria of medicinal chemistry and pharmacokinetics are used to filter the above database, Reymond and co-workers estimate that about 10 million small molecules are potential drugs [21]. With the assistance of robots, these molecules can be obtained quickly and one can generate large libraries of new compounds, most of which are made by applying robust reactions (typically amide formation, Suzuki–Miyaura C–C-coupling, Buchwald–Hartwig nucleophilic aromatic substitutions, azide + alkyne cycloadditions) [22,23,24,25]. We think this is not optimal today; the chemical space exploration must add many less-common reactions, especially recently developed catalytic processes [26]. Moreover, reactions permitting access to enantio-pure compounds (asymmetric synthesis) must be developed and applied. This requires not only imitation of literature procedures, but the invention of new and better ones. Nature remains a source of inspiration; complicated bioactive compounds such as alkaloids, polyphenols, polyketides, peptides, nucleic acids, oligosaccharides, and glycoproteins, as well as their mimics, should be available through synthetic chemistry. Synthesis can produce analogs and eventually better drugs than those offered by nature. For that, knowledge of physical chemistry, physics, mechanics (when designing micro-reactors and automatic systems, for instance), quantum chemistry, and the application of artificial intelligence can be quite beneficial. Molecular scientists must have a profound understanding of chemical reactivity and must be able to predict whether a planned reaction is possible or not, and at what rate it will occur under the conditions chosen. For more than 100 years, successes of synthetic chemistry have been the result of trial and error and serendipity [27]. This has turned many talented students away from synthesis and from chemistry. Today, synthetic chemistry, and especially organic and organometallic chemistry in homogeneous systems, is a reliable science with reproducible procedures based on thermodynamics, kinetics, and availability of highly performing analytical tools. Scientists using all this knowledge and techniques have more productive and inventive intuition [28]. The relationships between molecule dynamics and energy (heat, light) are established on a quantitative basis. With the help of quantum mechanical calculations, one can invent models that can be applied to almost any reaction. Literature and data banks (e.g., NIST WebBook of Chemistry [29]) provide us with quantitative and accurate thermodynamic data. The latter must be used to predict chemical reactivity (equilibrium constant and reaction rates) and the properties of the new compounds targeted. This is one of the messages that the Vogel–Houk book tries to pass to students, chemists, biochemists, and engineers. The domestication of fire one million years ago has permitted humans to continuously improve their quality of life. The treatment of matter with heat has created new objects and new knowledge (metallurgy, pottery, ceramics). The understanding of how matter and heat are interrelated is the basis of the sciences of matter. This should be kept in mind when teaching or/and applying chemistry, for instance.
Because the transformation of matter should not produce waste, atom economical reactions should be preferred to reactions producing co-products [30,31,32]. Catalytic processes should be chosen, as they save significant energy; furthermore, they should be run in inexpensive, non-toxic, and recoverable solvents, if any.
Amide formation is one of the most important reaction in cells and in the laboratory. The direct amidification of carboxylic acids by amines requires the elimination of water [33]. New catalysts have been invented that achieve direct amidification of carboxylic acids with amines in an aqueous solvent (DMSO, 70 °C) [34].
CH3COOH + MeNH2 ⇄ CH3CONHMe + H2O
Thermochemical data for Equilibrium (1) in the gas phase tells us that this direct amidification is exothermic, with ΔrH°(1) = −8 ± 1.5 kcal mol−1. Gas phase experimental entropies are not available for Me2NH and CH3CONMe2. The entropy variation ΔrS°(1) of Equilibrium (1) under 1 atmosphere and at 25 °C can be estimated from the variation of translation entropies [35], which depend on the molecular weights of reactants and products, according to Equation (2) for equilibrium A + BP + Q:
ΔrS°(A + BP + Q) ≅ ΔrS°trans = 1.5 Rln(MP·MQ/MA·MB)
For Equilibrium (1) in the gas phase, ΔrS°(1) ≅ −1 eu (entropy unit = cal mol−1 K−1). This gives an entropy cost of –TΔSr° ≅ 300 cal mol−1 = 0.3 kcal mol−1 at 25 °C. This indicates that the reaction must be exergonic (K298K(1) > 1): ΔrG° = ΔrH° –TΔSr° ≅ −7.7 ± 1.5 kcal mol−1 at room temperature. In solution, this value will be affected by differential solvation effects, and, when using larger carboxylic acids than acetic acid and different amines than methyl amine, the entropy variation will be more negative (e.g., for a carboxylic acid and an amine having both a molecular weight of 200, one obtains with relationship (2) ΔrS° ≅ −5 eu, and an entropy cost –TΔSr° ≅ 1.5 kcal mol−1). Thermochemical data for model Equilibrium (1) in the gas phase should encourage chemists to look for more catalysts that can induce the direct formation of amides at room temperature without a co-solvent such as DMSO, or with a co-solvent that can be recovered more readily, as well as for carboxylic acids and amines that contains other functions (e.g., ketone, aldehyde, ester, etc.). We hope that this example will convince chemists, biochemists, and engineers to use thermochemical data available for compounds in the gas phase, even though the planned reaction must be run in solution and at a temperature different than 25 °C.

3. Thermochemical Data Establish Fundamental Concepts of Reactivity

Under thermodynamic control, substitutions of alkyl halides (e.g., Equilibrium (3), Pr = prop-1-yl) and of alkenyl halides (e.g., Equilibrium (4), vinyl = CH2=CH) by other halides generally favor the formation of H-F in the gas phase:
Pr-F + HBr ⇄ Pr-Br+ H-F
vinyl-F + HBr ⇄ vinyl-Br + H-F
In contrast, Equilibria (5) that exchange the fluoride of acetyl fluoride (Ac-F = CH3CO-F) with chloride, bromide, or iodide with the corresponding hydrogen halide H-X disfavor the formation of H-F, meaning that fluorine prefers to be bonded to an acyl carbon atom rather than to a hydrogen atom. In contrast, equilibria of acetyl chloride (Ac-Cl) with the corresponding bromide and iodide are nearly thermoneutral in the gas phase.
Ac-F + H-X ⇄ Ac-X + H-F
ΔrHo(5) = 4.6 (X = Cl), 4.0 (X = Br), 4.0 (X = I) kcal mol−1
Why does fluorine prefer the left side (C center instead of the H atom) of Equilibria (5)? Donation of the non-bonding electrons of the oxygen atom of the carbonyl group stabilizes the ionic form of the acetyl-halide bond. This negative hyperconjugation effect (n(C=O:)/σ* interaction) involves the interaction of the non-bonding or lone pair, orbitals n(CO:) of the carbonyl group, and the anti-bonding, empty orbital σ*(C-F) of the C-F bond. This interaction is not possible in alkyl, alkenyl, and hydrogen halides, which do not possess lone pair electrons. Of all acyl halides, this hyperconjugative interaction is strongest in acyl fluorides, where the difference in electronegativity between carbon and fluorine is larger than in the other acyl halides. Because of the large electronegativity difference between F and C, σ*(C-F) is the best σ-acceptor of all C-X bonds. Furthermore, the conjugation n(X:)/π*(C=O) (donation from the non-bonded electron pairs of X: to the carbonyl double bond), which stabilizes the carbonyl compound, is the weakest for X=F, and strongest for amino groups (Figure 1). The infrared carbonyl stretching frequencies of acyl derivatives (νC=O) increase with the C=O bond strength, as shown below.
The thermochemical data presented above demonstrate that a carbonyl group is a π-withdrawing and a σ-donating group [36,37,38]. This explains the regioselectivity of the electrophilic additions (6) to the alkene moiety of bicyclo[2.2.1]hept-5-en-2-one (1) to be the opposite of those of its synthetic precursors, the corresponding cyanoacetate (4) (reactions (7), Scheme 1) [39,40,41]. Quantum mechanical calculations find that the 6-oxobicyclo[2.2.1]hept-2-yl (2, E=H) cation is much more stable than its 5-oxo isomer [42,43]. The former carbocation is stabilized by n(CO)/σ(C(1)-C(6))/π(C(6)) hyperconjugation, what is not possible in the latter isomer. This type of interaction allows keto [44] and ester groups [45] to migrate more quickly in Wagner–Meerwein and pinacolic rearrangements than alkyl and phenyl groups.
Carbon monoxide (C=O) has a dipole moment that places a partial negative charge on its carbon atom [46,47]. In contrast, carbon dioxide (O=C=O) has partial negative charges on its two oxygen atoms and a partial positive charge on its carbon atom. Carbon monoxide is in fact an electron-rich carbene because of the electron donation by the non-bonding electron of its oxygen atom. In carbonylmetal complexes, CO is a σ-donor and a π-attractor.
M=C=O ↔ M(−)-C≡O(+)

4. Mnemonic Devices Are Very Useful to Learn Chemistry, but They Must Be Applied Critically

We now discuss Markovnikov’s rule about the orientation of the additions of H-X to alkenes. The rule states that when water, a carboxylic acid, or a hydrogen halide adds to an unsymmetrical alkene, the hydrogen atom joins the carbon atom bearing the largest number of hydrogen atoms and the nucleophile (X) bonds to the vicinal center that is more substituted. The classic explanation invokes a two-step mechanism in which the proton adds first to the alkene forming a carbenium ion intermediate in the rate determining step, which is then quenched by the nucleophile irreversibly. This hypothesis implies kinetic control (the rate of the two competing regioselective additions are different and the major regioisomer results from the fastest reaction). The stability sequence for alkyl cations is tertiary > secondary > primary [48,49,50]. However, in addition, in the cases where R-X are alcohols, amines, thiols, and alkyl halides, secondary derivatives are more stable than their primary isomers (see Equilibria (8)). Similarly, tertiary systems are more stable than their secondary isomers [51]. The formulation of Markovnikov’s rule as a kinetic control is not always valid [52,53,54,55]. Additions to alkenes are exothermic but have negative entropies (condensations) that can cause the reactions to be reversible (with standard Gibbs energy variation ΔrGT = ±1 kcal/mol). For example, additions of water to unstrained alkenes in the gas phase are exothermic by ca. −12 kcal/mol, a value very similar to the entropy cost of the addition at 25 °C. For instance, in the gas phase, the standard heat of reaction ΔrH°(2-methylpropene + H2O ⇄ t-butanol) ≅ −12.6 kcal/mol, and the standard entropy variation of the equilibrium ΔrS°(2-methylpropene + H2O ⇄ t-butanol) ≅ −37 eu; at 25 °C, the entropy cost −TΔrS° amounts to −298·(−37 cal mol−1K−1) ≅ 11.0 kcal mol−1 [29,56]. Thus, one predicts that the regioselectivity of the additions of water, alcohols, and carboxylic acids to acyclic alkenes might be thermodynamically rather than kinetically controlled above 25 °C. As shown with Equilibria (8) in the gas phase, the linear adducts are less stable than their “branched” isomers. Additions of HX to terminal alkenes or 1,2-dialkylethylenes avoid the generation of secondary carbenium intermediates and follow other mechanisms that do not involve carbenium ion intermediates. The reverse reactions, that is, eliminations, also may follow concerted mechanisms avoiding carbenium ion intermediates. Even for such reactions, Markovnikov’s rule is generally followed. This is because of the Dimroth principle, which was enunciated in 1933 [57]. If one or a set of reactants can undergo two competitive one-step reactions that follow the same mechanism and produce two different isomers, the favored product under conditions of kinetic control is the most stable one. The energy barrier is the lowest for the most exothermic reaction (Bell–Evans–Polanyi theory established in 1936–1938 for radical exchange reactions such as R-X + Y → R + X-Y, and proton transfers: ΔH = αΔrH + β) [58,59]. Deviations to this principle are observed when differential steric factors and solvation effects affect the reactions.
Prop-1-yl-X ⇄ Prop-2-yl-X
X =MeEtn-PrOHSHNH2FClBrI
ΔrH°(8):−2.0−1.6−1.8−4.2−4.0−3.2−1.8−3.1−3.0−2.4 kcal mol−1
Mnemonic tools such as the Markovnikov’s rule and the ever-popular arrow-pushing are very useful when one learns chemistry [60]. At some point, we need to remember how they have been constructed and on what measurements they are based upon. Scientists should not be afraid to be critical of the models proposed to them and should ask for more experimental proofs.

5. Models Must Be Based on Measurements and Thermochemical Data

We next discuss the mechanisms of some pericyclic reactions. Among them, cycloadditions, cheletropic additions, and ene-reactions all permit the construction of C-C and C-hetero bonds in an atom economical fashion, sometimes with high chemo-, regio-, stereo-, and enantioselectivity. Isomerizations such as electrocyclic ring opening or closing reactions and sigmatropic rearrangements are also very interesting synthetic tools and will continue to play an important role in synthesis. Most pericyclic reactions tolerate a large variety of substituents and permit access to a large molecular diversity. Furthermore, these reactions can be catalyzed and run under very smooth conditions, saving energy. Before 1965, concerted pericyclic processes were called “no-mechanism reactions” to imply that there are no detectable intermediates and everything happens in one step. Since then, the study of their mechanisms has revealed many subtleties and has contributed enormously to our understanding of chemical reactivity. Depending upon the nature of the reactants (type of substituents, number of substituents, etc.) and reaction conditions (temperature, solvent, presence of additives and of catalysts), the pericyclic reactions may not follow concerted, one-step mechanisms, but can be multi-step processes.
The Woodward–Hoffmann rules explain the stereoselectivity of concerted pericyclic reactions [61]. The rules and other related theories such as the Longuet–Higgins correlation of electronic configurations between reactants and products rely upon symmetry in the transition structures [62]. Perfectly symmetrical transition structures are very rare, especially when non-symmetrical reactants are engaged (as demonstrated by kinetic isotopic effects). The activation enthalpy (ΔH) of a concerted reaction is the sum of repulsive steric effects and skeletal distortion energies, minus stabilization energies resulting from stabilizing orbitals interactions in the transition state [63]. The FMO theory applied to transition structures of pericyclic reactions also predicts their stereochemistry, as well as their aromatic character [64,65] (Evans’ rule) [66]. FMO theory provides estimates of substituent and additive effects on the rate and the chemo-, regio-, and stereoselectivity of the reactions without requiring symmetrical transition structures [67]. Another approach takes advantage of the fact that non-symmetrical transition structures of pericyclic reactions are diradicaloids, i.e., species with very narrow HOMO–LUMO gaps, or in valence bond terminology, diradical Catalysts 12 00758 i008 zwitterion species. Thus, without having to calculate molecular orbitals, by using just thermochemical data (homolytical bond dissociation enthalpies: DH°(R/H) or DH°(R/X)), the substituent effect on the relative stability of radicals, cations and anions, ionization energies (IE(R) [68], and electron affinities of radicals (–EA(R)) [69] can be used to predict relative rates and chemo-, regio-, and stereoselectivities of concerted, one-step pericyclic reactions. The diradicaloid model, advocated by Dewar for Diels–Alder reactions [70,71], is simple because it is based on thermochemical data (measurements) and it can be applied to catalyzed pericyclic reactions.
For the cyclodimerization of ethylene into cyclobutane (reaction (9), gas phase, 25 °C, 1 atm.), one estimates the temperature at which cyclobutane equilibrates with ethylene. The equilibrium constant KT(9) = 1 (ΔrG° = ΔrH° − TΔrS° = 0) is realized at a relatively low temperature, as given by ΔrH°/ΔrS° ≈ 18,400/41.6 ≈ 442 K = 169 °C. Above this temperature, the cyclodimerization is endergonic (KT < 1). The cycloreversion of cyclobutane into ethylene occurs between 350 and 450 °C with a measured barrier ΔH = 61.1 kcal mol−1. One calculates ΔfH° (tetramethylene diradical) = ΔfH° (cyclobutane) + DH° (Et/Et) – ring strain of cyclobutane = 6.6 + 86.0 − 26.5 = 66.1 kcal mol−1. This gives an enthalpy difference ΔrH°(cyclobutane ⇄ tetramethylene diradical) = 66.1 − 6.6 = 59.5 kcal mol−1. With a measured barrier ΔH(cyclobutane → 2 ethylene) = 61.1 kcal mol−1, the thermochemical data support the hypothesis of a diradical mechanism for the cycloreversion (9), which has been demonstrated through femtochemistry by Zewail and co-workers [72]. The reaction is not concerted or pericyclic; there is no assistance between the σ-bond breaking and π-bond forming processes (Scheme 2). The thermally “allowed” concerted [π2s + π2a] cycloreversion is not followed, since it involves too severe distortions of cyclobutane to reach the twisted transition structures predicted by the Longuet–Higgins theory and the Woodward–Hoffmann rules. The entropy of activation ΔS of this condensation is that of the reaction ΔrS° corrected for the three free rotations about the three single C-C bonds of the 1,4-diradical (3 times 5 eu). This leads to the estimated ΔS = −41.6 + 15 ≈ −27 eu. One obtains finally the Eyring activation free enthalpy ΔG = ΔHTΔS = 41 − 442·(−0.027) ≈ 53 kcal mol−1 (with ΔH = Eyring activation enthalpy). Using equation lnk = −ΔG/RT + lnT + 23.76 = −53,000/1.987·442 + ln(442) + 23.76 ≈ −30.7, one obtains a rate constant k(9) ≈ 4.6 × 10−14 dm3 mol−1 s−1, or a half-life (time for 50 % conversion) τ½ = 1/k(9) [initial concentration of ethylene = 1] ≈ 682,400 years! This reaction is too slow to be observed at 442 K.
If one considers the substituent effects on the relative stability of radicals, one can predict the relative rate and the regioselectivity of all (2 + 2)-cycloadditions of alkenes. Thermochemical analysis analogous to that given for the cyclodimerization (9) of ethylene into cyclobutane suggests that the octa-1,7-diene-3,6-diyl diradical is also an intermediate of the (2 + 2)-cyclodimerization of butadiene. When the alkenes are substituted with polar groups, their cycloadditions are solvent dependent because zwitterions can be formed in place of diradicals.

6. Diradicaloids Are Transition Structures of Concerted Cycloadditions

Figure 2 represents the possible paths of a Diels Alder reaction. For the same orientation (regioselectivity), there are three limiting mechanisms; two of them involve the formation of (Z)-hex-5-en-1,4-diyl diradicals, and the third one corresponds to the synchronous one-step reaction, as implied by the Woodward–Hoffmann rules. Between these extremes are the concerted but two-stage mechanism proposed in 1959 by Woodward and Katz [73]. For the retro-Diels–Alder reaction of norbornene into cyclopentadiene and ethylene, Zewail and co-workers proposed from their femtosecond real-time studies that both concerted and non-concerted trajectories are possible; however, they decomposed norbornene by an intense laser pulse and generated the diradical by a photochemical process [74]. Were (E)-hex-5-en-1,4-diyl diradicals to be formed, they could not cyclize, but would induce polymerization of the cycloaddends, as already discussed by Littmann in 1936 [75]. Polymerization often accompanies cycloadditions of relatively unreactive dienes unless a radical scavenging agent is added to the reaction mixture. In between the three limiting mechanisms defined above, a spectrum of non-synchronous one-step mechanisms are generally thought to occur, as supported on the basis of kinetic isotope effects [76].
Thermochemical data tell us that the preferred diradical intermediate is the one for which the substituents stabilize the radicals the best, and this implies, as a rule, that this diradical results from combining the least substituted centers of the diene and dienophile. In other words, the most stable isomeric diradical intermediate is the one that has generated the strongest bond between the two reactants (for instance D1 or D1’, Figure 3). Since many Diels−Alder reactions have measured activation enthalpies ΔH that are lower than the enthalpy difference ΔrH° between the diradical and the cycloaddends, the difference ΔrH° − ΔH can be taken as Doering’s energy of concert (Figure 3). The transition state of such reactions can be readily described in terms of frontier molecular orbital interaction theory [77,78]. An alternative valence-bond representation is that transition structures can be represented by diradicaloids, i.e., diradicals that exchange electron between the two radical entities. Within the valence bond theory, this electron exchange can be represented by a (Z)-diradical and one or two zwitterions, as shown in Figure 3 with D1 Catalysts 12 00758 i008 Z1 Catalysts 12 00758 i008 Z2.
The diradicaloid model explains the chemo-, regio-, and stereoselectivity (e.g., endo rule of the Diels–Alder reactions) of all cycloadditions and cheletropic additions. If one of the substituents A or D can be coordinated selectively to an additive, a catalytical effect can be observed, which is explained also by the diradicaloid model.

7. Diradicaloid Transition Structures of Concerted Sigmatropic Rearrangements

Sigmatropic rearrangements and especially the (3,3)-sigmatropic rearrangements such the 3,4-diaza-Cope rearrangement in the Fischer’s indole synthesis, the Claisen rearrangements and its various variants, the aza-Claisen, the Overman, and the Cope rearrangements have played an important role in synthesis, and will continue to do so. Here, also one considers a spectrum of mechanisms (Figure 4) in which there are two limiting paths involving diradical intermediates, and in between these two limits, concerted one-step mechanisms. One limiting mechanism is dissociation with the formation of two allyl radicals 7 + 8; the other is the associative formation of cyclohexa-1,4-diyl diradical intermediate of type D3 (Grob’s hypothesis).
Gajewski and Conrad mapped out the positions of transition states on such diagrams through their measurements of secondary deuterium kinetic isotope effects and comparisons to equilibrium isotope effects [79]. In the concerted one-step mechanisms, the bond forming process assists the bond breaking process. Their transition states can be represented by transition structures belonging to two types of diradicaloids (Figure 5). The two allyl radicals 7 + 8 exchange an electron, forming ion-pairs IP1 or/and IP2. On its side, diradical D3 is stabilized by electron-exchange leading to zwitterions Z3 or/and Z4. Substituent effects on the relative stability of radicals, cations, and anions allow one to predict the reactivity of these concerted reactions. This model explains substituent effects on rates and the accelerating effects of catalysts. For instance, any additive that can stabilize one or the other of the charge-transferred limiting structures will catalyze the (3,3)-sigmatropic rearrangement.

8. Diradicaloid Transition Structures in Concerted Ene-Reactions and Related Processes

The Alder−ene reaction is the thermal reaction of two alkenes that are condensed into another alkene with the formation of a new σ(C-C) bond and the transfer of a hydrogen atom between them. A large number of analogous reactions are known for heteroatoms containing unsaturated reactants, such as the aldol reaction (10) and the carbonyl-ene reaction (11) (Figure 6). In the absence of Lewis or Brønsted acid, the transition structure of the enol aldol condensation can be represented by a diradicaloid (1,4-diradical Catalysts 12 00758 i0081,4-zwitterion) of type D4 Catalysts 12 00758 i008 Z5 if it follows a concerted one-step mechanism (Figure 6, reaction (10)). The role of the acid catalyst A is to stabilize the zwitterionic limiting structure Z5 by equilibrating with zwitterion Z5’. As a consequence, the reaction may follow a multi-step mechanism in which Z5’ is one reaction intermediate that can adopt quasi-cyclic or acyclic conformations. Similar diradicaloid models (D5 Catalysts 12 00758 i008 Z6) can be considered for metalla-carbonyl-ene reactions promoted by a Lewis acid (Figure 6, reaction (11)). Very electrophilic ketones such as hexafluoroacetone react with allylsilanes without catalyst [80]. The Hosomi–Sakurai reaction promoted by TiCl4 (Figure 6, reaction (11)) is not only regiospecific, but also stereospecifically anti with respect to the silyl group, which leads to the formation of a (E)-homoallylic alcohol [80,81,82]. ((E)-crotyl)trimethylsilane and ((E)-cinnamyl)trimethylsilane react with aldehydes in the presence of TiCl4 to give homoallyl alcohols with over 93% syn-selectivity (Figure 6, reactions (12) and (13)). Lower syn-selectivity is observed for the reactions of (Z)-crotylsilanes [83,84]. The stereoselectivity is generally accepted to be determined by chair-like six-membered (“closed”) transition states [85,86,87,88,89,90].

9. Valence Bond Theory Can Be Used to Explain Structural and Medium Effects on the Rates of Displacement Reactions

The diradicaloid model is not limited to the transition structures of concerted pericyclic reactions. Concerted displacement reactions (14) are quite important in synthesis.
R-X + Y:/M+ ⇄ R-Y + X:/M+
Shaik and Pross, as well as others, have described the SN2 reaction transition states in terms of valence bond theory [91,92,93,94,95,96,97,98]. The transition structure of a SN2 reaction (Figure 7) engaging a negatively charged nucleophile can be represented by the limiting structures:
[Y:/R/X ↔ Y:/R+/X: ↔ Y/R/X: ↔ Y/R:/X].
This representation allows one to predict that the weaker the R-X bond (DH°(R/X) = ΔfH°(R) + ΔfH°(X) - ΔfH°(R-X), ΔfH° = standard heat of formation); the faster the nucleophilic displacement. Furthermore, the higher the stability of cation R+ and of nucleofugal group X:, or/and the higher the stability of the carbanion R: the faster the reaction is. Nucleophilicity of Y: depends on how much the anion is dissociated from its counter-ion M+, and this depends upon the medium polarity and/or catalyst that interact specifically with M+. The lower the ionization energy (EI(Y:)) of the nucleophile Y:, the faster the displacement reaction is. The higher the electron affinity (−EA(R-X)) of the electrophile R-X, the faster the reaction is. Thermochemical data such as ionization energies and electron affinities permit one to evaluate the power of the pushing effect of the arrow associated with the attack of the electrophile by the nucleophile, and of the power of the pulling effect of the arrow associated with the departure of the nucleofugal group X:. Additives that can combine with the nucleofugal group X (and not with the nucleophile Y) are catalysts of the displacement reaction.

10. Hydrocarbation of Unsaturated Compounds by R-H Reagents Do Not Produce Waste

Classically, for the formation of single C-C bonds, chemists have relied upon displacement reactions by matching a nucleophile with an electrophile (e.g., Nu-H + R-X + base → Nu-R + base-H+/X, or Nu-metal + R-X → Nu-R + metal+/X; R = alkyl, alkenyl, aryl, heteroaryl, alkynyl) [99] and on electrophilic substitutions (Ar-H + E-X ⇄ Ar-E + HX; e.g.,: Friedel–Crafts alkylation and acylation) [100]. Catalyzed or photo-catalyzed free-radical additions to C-C multiple bond are applied more and more in fine organic synthesis, not only to generate polymers (e.g., R-X (X= I, Br) + BCH=CHA → R-CH(B)-CH(X)-A) [101,102,103,104,105,106]. Aldehyde and ketone olefinations are also powerful methods to create double C=C bonds. All these reactions are not atom economical, as they produce co-products. The same can be said for the very much used additions of organometallic reagents to unsaturated systems (Nu-metal + A=E ⇄ Nu-A-E-metal, with A, E = carbon or heteroelement; = is a double or triple bond) [107]. Direct additions (15) of R-H to unsaturated compounds are more atom economical [108,109]. One-step concerted [π2s+σ2s]-cycloadditions are predicted to be difficult reactions (symmetry forbidden, small LUMO/HOMO overlap, and large energy gaps between these reactant MO’s). Conformational strain required for the theoretically allowed concerted [π2a+σ2s]- or [π2s+σ2a]-cycloadditions renders these one-step reactions impossible.
R-H + X=Y ⇄ R-X-Y-H
Many of them are base-catalyzed and involve the conjugate bases R of R-H as intermediate. Examples are the aldol (e.g., R1CHO + R2CH2COR3 ⇄ R1CH(OH)-CH(R2) –COR3) and the Claisen condensation (e.g., R1CH2COOR2 + R3CHO ⇄ R1CH(COOR2) –CH(OH)R3), the Henry reaction (e.g., R1CH2NO2 + R2CHO ⇄ R1CH(NO2)-CH(OH)R2), and the Michael addition (e.g., R1CH2COR2 + CH2=CHCOR3 → R1CH(COR2)-CH2–CH2COR3) [110]. Under acidic conditions, the most common reactions are the Mannich reaction (e.g., R1COCH3 + CH2O + R2NH → R1COCH2-CH2-NR2 + H2O), the Alder ene-reaction (A-CH=CH2 + R2C=CH-C(R’2)H ⇄ A-CH2-CH2-C(R2)-CH=C(R’2), the alkylation of arenes by alkenes (e.g., Ar-H + CH2=CHR ⇄ ArCH(Me)R), the alkylation of alkenes with alkanes (e.g., the alkyl process, synthesis of isooctane from isobutene and isobutane), the hydroxyalkylation of alkenes by aldehydes (e.g., the Prins reaction RCH=CH2 + CH2=O ⇄ RCH=CH-CH2OH) [111,112], and the hydroxyalkylation of aromatic compounds with aldehydes. Other processes are catalyzed by a nucleophile (KCN, amine, phosphine, diaminocarbene) such as the benzoin condensation (e.g., 2 RCHO → RCH(OH)-COR), the Stetter reaction (e.g., R1CHO + R2-CH=CH-COR3 → R1CO(R2)CH-CH2-COR3), the Rahut-Currier reaction (e.g., ArCOCH=CHCOOR + CH2=CHCOMe → ArCOCH2CH(COOR)C(=CH2)COMe), and the Morita-Baylis-Hillman reaction (e.g., R1CHO + R2CH=CHA → R1CH(OH)-C(A)=CHR2 [113]. Since the discovery of the Fischer–Tropsch reactions (e.g., nCO + (2n+1)H2 → nH2O + CnH(2n+2)) and the Roelen hydroformylation of alkenes (RCH=CH2 + CO + H2 → RCH2CH2CHO + RCH(CHO)Me), transition metal-catalyzed hydrogenative C-C coupling reactions are taking more and more importance in fine organic synthesis [114,115,116]. Because it can be applied to a large variety of reactants and be asymmetric, Krische’s hydrogenative coupling of dienes (or alkynes) with carbonyl compounds and imines (e.g., CH2=CH(R1)CH=CH2 + H2 + R2CHO → CH2=CH-C(Me)(R1)-CH(OH)-R2) is a very powerful synthetic tool [117,118,119,120,121]. Alternatively, transition metal complexes can dehydrogenate primary alcohols and thus provide the corresponding aldehydes as an intermediates and, formally, H2 necessary for the hydrogenative coupling of alkenes and aldehydes. Thus, one can use alcohols in the direct hydro(hydroxycarbation) of alkenes, 1,3-dienes, allenes and alkynes. The concept developed by Krische and co-workers is therefore a formal hydrocarbation (16) of unsaturated compounds by primary alcohols, a reaction that can simultaneously generate up to two new contiguous stereogenic centers [122,123,124].
H2C=CR1(R2) + R3CH2-OH → CH3-C(R1)(R2)-CH(R3)-OH
Krische defines the process as being a hydro(hydroxycarbation) via transfer hydrogenation. An example is given with Reaction (17) (Scheme 3). The alcohol equilibrates with its aldehyde and an iridium hydride that adds chemoselectively to the alkene generating a methyl(propargyl)iridium intermediates, which, in turn, adds to the intermediate aldehyde. When the transition metal catalyst is coordinated to an enantiomerically pure ligand, the hydro(hydroxycarbation) can be highly stereo-(dr = diastereoisomeric ratio) and enantioselective (ee = enantiomeric excess) [125,126].
There have been many more exploratory studies on transition metal-catalyzed reactions than mechanistic studies. Thermochemical and kinetic data for transition metal complexes and their reactions are not as numerous as for organic compounds, but are being recorded more and more frequently. Nevertheless, because the structures and the reactions of transition metal complexes can be modeled by the structures and reactions of organic compounds, simple models are available to “classify” the reactions catalyzed by transition metal species.
In classical chemistry, direct hydrocarbations of unsaturated compounds X=Y are non-concerted processes. If R is an allyl moiety, a concerted ene-reaction can occur (e.g., CH2=CH-CH2R’ + X=Y → H-X-Y-CH2-CH=CHR’). If the catalyst is a base B:, it deprotonates R-H forming the conjugate base R- as an intermediate, which then adds to the alkene (or alkyne), generating an anionic adduct. The latter is then quenched by the conjugate acid BH+ of the basic catalyst (nucleophilic addition (18)) or by R–H, realizing a chain process. Concurrently, the anionic adduct intermediate can add to another alkene (alkyne) and start the polymerization of the latter (anionic polymerization).
A second mechanism involves a radical intermediate R in which the catalyst is a radical or a species that can abstract a hydrogen atom from R-H. The addition of radical R to the alkene (alkyne) produces another radical that abstracts a hydrogen atom from H-X (radical addition (19)) or from R-H (chain process). Concurrently, the intermediate radical resulting from the initial radical addition can, in its turn, add to another alkene (alkyne) and lead to the polymerization of the latter (radical polymerization).
A third mechanism involves a cationic intermediate R+ resulting from a hydride abstraction by a catalyst capable of abstracting a hydride from R-H reversibly. Such a mechanism corresponds to an electrophilic addition (20). Alternatively, the alkene can be protonated given a carbenium ion intermediate that adds to another alkene generating carbocationic adduct. At this stage, the latter is reduced by a hydride transfer from the alkane (e.g., the alkylate technology for the production of high octane number gasoline from a mixture of butene and isobutane) [127,128,129,130,131]. Concurrently, the first cationic adduct can add to another alkene, leading to its polymerization (cationic polymerization). Formally, instead of initiating the reaction by forming a carbenium ion intermediate R+, one can imagine a catalyst M+ that combines with R-H forming an organometallic species R-M and a proton. The protonation of the alkene (alkyne) follows, forming a carbenium ion that is quenched by R-M (Reaction (21), Figure 8), or that adds to another alkene (alkyne), starting its polymerization [132,133,134,135,136,137,138,139]. Mechanisms that involve a single electron transfer are also possible [140,141,142].

11. Transition Metal-Catalyzed Hydrocarbations

Transition metal-catalyzed processes follow mechanisms in which several intermediates equilibrate with reactants and products. Thus, thermodynamics control the outcome of the process in many instances. Other transition-metal-catalytic reactions that are not reversible have rate-determining steps that determine the chemo-, regio-, stereo-, and enantioselectivity. The latter can be guessed, or quantum chemical calculations can help, at least in rationalizing the observations. A number of rules are available from thermochemical data; for instance, the M-C bonds are weaker than the M-H bonds, and the C-C bonds are weaker than the C-H bonds for comparable systems (same type of substitution). Scheme 4 with reaction (22) is generally retained for transition metal-catalyzed direct hydrocarbation of unsaturated compounds.
Unsaturated transition metal complexes (with 16 or fewer valence electrons) are able to add across the σ(C-H) of R-H and form adducts of type R-[M]-H (can be an equilibrium of geometrical isomers) in which a σ(M-R) and a σ(M-H) bond are formed. In this adduct, the metal atom has gained two more valence electrons (VE) [143,144,145,146,147]. This is called an oxidative addition [148]. The oxidative addition (23) of a transition metal complex ML to a C-H bond is modeled in organic chemistry with the insertion of a carbene into a single C-H bond to give an alkane. For instance, methane adds to methylene carbene (reaction (24)), giving ethane (ΔfH° = standard heat of formation in the gas phase). The carbene has a carbon atom with a sextet of VE, meaning that it is unsaturated, whereas the carbon center in the product (alkane) has an octet of VE, meaning it is saturated. Another comparison is the reaction of a methyl cation with methane that equilibrates with an ethonium ion (C2H7+ = protonated ethane) in the gas phase [149]. Dihydrogen (H2) reacts with methyl cations (which has a sextet of VE on its carbon atom), giving a methonium ion (CH5+ = protonated methane, carbon atom with an octet of VE) as shown (Scheme 5) with Equilibrium (25) [150,151,152,153,154].
Transition metal complexes are good catalysts because they can exchange their ligand at relatively low temperature and assemble two reactants that must combine around the metal atom (template effect). Because bonds around the metallic center have bond angles that can be changed with little deformation energy, reactions that combine two reactants have relatively low energy barriers. Furthermore, the metallic center is highly polarizable, which means that it can take or give electrons very readily and at each step of the process.
The first and last steps of a transition-metal-catalytic process are ligand exchanges, as illustrated with the hydrocarbation reaction (22). The exchange of a ligand L (carbon monoxide, amine, phosphine, arsine, ether, aldehyde, ketone, alkene, alkyne, arenes, H2, etc.) with an alkene (Reaction (26)) is modeled by the cyclopropanation of alkenes with methylene generated by photolysis of ketene (Reaction (27)), for instance, (Scheme 6). Another reaction to consider as a model is the formation of a H-cycloproponium ion by the addition of a methyl cation to an alkene (reaction (28)). H-cycloproponium ions are intermediates (species longer-lived than transition structures) in Wagner–Meerwein rearrangements (see Table 1) [155].
One of the processes that generates a bond between two reactants is the β-insertion (Reaction (30), the opposite reaction being a β-elimination). For the insertion of a hydrogen atom, the model reaction is the corner-to-corner migration in H-cycloproponium ions that corresponds to a 1,3-hydrogen shift, as shown with reaction (29) [156]. The rearrangement involves a transition structure that is a C-cycloproponium ion or edge-protonated cyclopropane (Scheme 7). A priori, any R group can do such migration (β-insertion). After the β-insertion, the metallic center has lost two VE. For the reaction to occur, a ligand L must coordinate the unsaturated metallic center.
After the β-insertion, a reductive elimination (Reaction (31)), occurs, forming the final product of hydrocarbation and an unsaturated metallic species that can return as the active catalyst for another oxidative addition with reactant R-H, or be stabilized by the coordination of a ligand molecule L of a solvent molecule (Scheme 8). The reductive elimination is the opposite reaction of the oxidative addition (e.g., Reaction (23)). If the 16-VE metallic adduct intermediate is an allylmetal species, it may add to the unsaturated system in a metalla-ene reaction [157,158].
The other processes (α-insertion/α-elimination, α-cycloinsertion/α-cycloelimination, β-cycloinsertion/β-cycloelimination) induced by transition metal complexes are also modeled by reactions of carbocations, as summarized in Table 1 [156].

12. Review of the Fundamental Reactions of Simple Carbocations

The mass spectrum of methane (CH4) has peaks with mass/charge ratios m/Q = 12, 13, 14, 15, and 16 corresponding to the cations C•+, HC+, H2C•+, H3C+, and H4C•+, respectively. The parent cation H4C•+ (or molecular ion) can dissociate into hydrogen radical H and methyl cation H3C+, and so on. If there is sufficient pressure in the ionization chamber, the transfer of a proton (32) can take place (bimolecular reaction):
H4C•+ + CH4 → H3C + CH5+
In this case, a peak is observed in the mass spectrum of methane with a mass/charge ratio m/Q = 17, which corresponds to methonium ion, CH5+, whose carbon atom is surrounded by an octet of valence electrons. It is much more stable (by 45 kcal mol−1, Equilibrium (25)) than methyl cation, H3C+, whose carbon atom is surrounded by a sextet of valence electrons. Methonium ion is the prototype of carbonium ions. Methyl cation is the prototype of carbenium ions. The notation H4C+ describes the methanium radical-cation, which has seven valence electrons [159,160].
The salts of methyl and ethyl cations cannot be observed as persistent species in solution. When CD3CH2F is allowed to react with an excess of SbF5/SO2 at −78°C, complete deuterium/hydrogen scrambling in the ethyl group of ethyl fluoride is observed by 1H-NMR spectroscopy. However, when CH3CH2F is added to DSO3F/SbF5 or DF/SbF5 in SO2 at −78°C, no detectable incorporation of deuterium into the ethyl group occurs [161]. This demonstrates that scrambling takes place intramolecularly in the ethyl cation intermediate (Scheme 9). Quantum calculations predict similar stabilities for classical ethyl cations and the bridged πH-ethenium ions (Equilibrium (33)) [162,163], which has been confirmed by infrared spectroscopy of the ion in the gas phase [164,165]. The πH-ethenium ion is the smallest possible π-complex of an alkene.
The simplest all-carbon and hydrogen carbenium ion prepared as a stable salt in solution is the i-propyl cation (Scheme 10). The 1,1,1-trideuterio-2-propyl cation (generated by treatment of 1,1,1-trideuterio-2-chloropropane in SbF5/SO2ClF at −110°C) rearranges (Equilibrium (34)) at −60°C into isomeric ions with Arrhenius activation parameters Ea = 16.4 ± 0.4 kcal mol−1 and logA = 13.2 ± 0.3 [166]. The [2-13C]prop-2-yl cation generated from [2-13C]prop-2-yl chloride at low temperature was found to undergo intramolecular carbon scrambling between −90 and −60°C (Equilibrium (35)). This requires a process involving a corner-protonated cyclopropane or an edge-protonated cyclopropane intermediate (Equilibrium (36)) [167].
The edge-protonated cyclopropane is somewhat higher in energy than the corner-protonated cyclopropane, although in the gas phase, fast scrambling (Equilibrium (36)) of hydrogen in protonated cyclopropane has been observed (Scheme 11) [168]. Infrared absorption features of gaseous isopropyl cations showed that corner-protonated cyclopropane is 8 kcal mol−1 less stable than isopropyl cations [169].
The 1H-NMR spectrum of the but-2-yl cation (s-Bu+) in SbF5/SO2ClF solution measured below −120°C indicates a very fast proton interchange due to Equilibrium (37) [170,171]. Between −100 and −40°C, all the hydrogen atoms in s-Bu+ are scrambled with an activation energy Ea =7.5 ± 0.1 kcal mol−1 (logA = 12.3 ± 0.1). If a series of 1,2-hydride shifts were responsible for the proton interchange, formation of a n-butyl (but-1-yl) cation intermediate (n-Bu+) would be required (Scheme 12). However, by analogy with the isopropyl cation (Equilibria (34) and (35)), the formation of n-Bu+ would be expected to require an activation energy higher than 15 kcal mol−1. Hence, a more facile process involving formation of a protonated methylcyclopropane and degenerate hydrogen shifts via an edge-protonated cyclopropane intermediate or transition state is more probable (Equilibrium (38)). When heated to −40 °C, the but-2-yl cation rearranges irreversibly into a t-butyl cation with an activation energy of 18 kcal mol−1. The similarity of this energy barrier with that observed for the isopropyl ⇄ n-propyl cation rearrangement (Ea ≅ 16 kcal mol−1) indicates that the reaction takes place via the primary carbenium ion intermediate isobutyl cation, i-Bu+, although this species is undoubtedly bridged by H or methyl in its minimum energy structure.
Gas phase measurements indicate that the protonated cyclobutane is less stable than s-Bu+, protonated methylcyclopropane and n-Bu+ (Scheme 13) [172].
Two different types of hydrogen exchange processes have been observed in the 1H-NMR spectrum of the t-amyl cation (2-methylbut-2-yl cation) (Scheme 14) [173]. One involves the interchange of the two types of methyl group protons, not affecting the methylene group (Equilibrium (40)). On the NMR time scale, this process occurs at temperatures above 0 °C and has an activation energy of Ea ≅ 15 kcal mol−1 (logA ≅ 13.2). The rearrangement first involves a 1,2-sigmatropic shift of hydrogen generating a secondary alkyl cation intermediate (3-methylbut-2-yl cation). This is followed by a 1,2-sigmatropic shift of a methyl group (Wagner–Meerwein rearrangement) that generates another secondary alkyl cation intermediate (3-methylbut-2-yl cation). The latter undergoes a quick 1,2-sigmatropic shift of a hydrogen-producing t-amyl cation via a protonated dimethylcyclopropane (as intermediate or transition state). Since degenerate Wagner–Meerwein shifts of both hydrogen and methyl groups are very fast (Ea < 4 kcal/mol, see e.g., Equilibrium (37)) when no change in the degree of chain branching of the carbenium ion occurs, the major part of the activation energy for this rearrangement (15 kcal mol−1) must be due to the formation of the secondary ion intermediate (3-methylbut-2-yl cation) from an t-amyl cation (2-methylbut-2-yl cation) [173].
The second process (41) occurs above 80 °C on the NMR time scale. It causes the scrambling of all the hydrogen atoms in a t-amyl cation. It has an activation energy Ea = 18.8 ± 1 kcal mol−1 (logA = 13.2 ± 0.5). A mechanism involving reversible hydride shifts to primary carbenium ion intermediates (e.g., 2-methylbut-1-yl cation) can be rejected on the basis that such species would have energies well above (ca. 30 kcal mol−1) that of the starting tertiary cation (t-amyl cation). In the alternative mechanism shown with Equilibrium (41), the hydrogen atom interchange involves corner-to-corner hydrogen shifts in the protonated dimethylcyclopropane intermediates. Thus, this process, which involves an overall change in the degree of chain branching, has an energy barrier 3 kcal mol−1 higher than those for related processes not involving such a change [155,174]. X-ray diffraction studies of crystals of a salt of 2-norbornyl cations ([C7H11]+] [Al2Br7].CH2Br2) at 40 K demonstrated this secondary alkyl cation adopting a Cs symmetrical structure, that is, a H-nortricyclonium cation (a corner-protonated cyclopropane) (Scheme 15) [175].
The secondary cyclopentyl cation is stable in SbF5/SO2ClF solution up to -20°C [176]. It undergoes a fast, degenerate 1,2-hydride shift (equilibrium (42), Ea < 4.5 kcal mol-1), but no ring contraction (equilibrium (43)) has ever been detected (Scheme 16). Considering an energy difference of 15 kcal mol−1 for the isomerization of a secondary alkyl cation to a primary isomer and the ring strain difference of 20 kcal mol−1 between cyclopentane and cyclobutene, one estimates the cyclobutylmethyl cation to be less stable than the cyclopentyl cation by ca. 35 kcal mol−1. Contrary to the cyclopropylmethyl cation, which is highly stabilized [177,178], the cyclobutylmethyl cation does not enjoy the same hyperconjugative stabilization [179,180].
In the gas phase, the doubly 13C-labelled cyclopentyl cations 11 and 12 generated from the doubly 13C-labelled bromides 710 undergo degenerate carbon scrambling (Equilibrium (44)) prior to elimination of ethylene (Fragmentation (45)). The complete carbon scrambling can, in principle, proceed via either the cyclobutylmethyl cation intermediate (Equilibrium (43)) or the “non-classical” pyramidal cation 13 (Equilibrium (44)), as calculations have suggested that they have similar energies [181,182]. Cation 13 can be viewed as a π-complex between ethylene and a cyclopropyl cation (corner-protonated spiropentane) or as a double π-complex between two ethylene units and the methylidynium ion, :CH+ (Scheme 17). The fragmentation of the cyclopentyl cation into the allyl cation and ethylene (Fragmentation (45)) may also involve cation 13 as an intermediate. The dissociation of an ethylene unit from 13 via transition state 14 is expected to give an allyl cation and ethylene. This process would be analogous to the gas phase fragmentation of the corner-protonated cyclopropane into molecular hydrogen and allyl cations. Mechanisms similar to those presented with Equilibrium (44) have been proposed to rationalize the complete hydrogen and carbon atom scrambling in the gaseous cyclohexyl ⇄ methylcyclopentyl cation equilibrium that occurs prior to monomolecular elimination of ethylene [183]. Cation 13 is also formed upon protonation of spiropentane. In the gas phase, its lifetime exceeds 7 × 10−9 s. Cation 13 is separated from the cyclopentyl cation by 30 kcal mol−1 [184,185].

13. Conclusions

The inventor of new processes can predict equilibrium constants and reaction rates by applying simple rules of thermodynamics and using published thermochemical data. Organic chemistry today offers a fantastic panoply of new reactions, especially catalytic reactions, that render the exploration of chemical space faster and more facile. Very important and sustainable processes are available, especially when atom economical processes are engaged, such as pericyclic reactions and direct hydrocarbation of unsaturated compounds. Catalysts can be invented that permit these reactions to be run under smooth conditions (no cooling, no heating, no solvent or in harmless solvents such as water, ethanol, ethyl acetate, or 2-methyltetrahydrofuran). There is a catalyst for any kind of reaction that permits energy savings and reduces the co-production of waste. The Vogel–Houk book has been written to help engineers and inventors realize a more sustainable world and to quickly generate a large number of new molecules that our civilization needs or will need. It also illustrates how fundamental concepts of chemical reactivity are based on thermodynamics. In this review, we have extracted a few models that help to show the analogies that exist between reactions of organic compounds and transition organometallic compounds. They help to understand a large number of catalyzed processes.

Author Contributions

Conceptualization, P.V. and K.N.H.; writing—original draft preparation, P.V.; writing—review and editing, P.V. and K.N.H. All authors have read and agreed to the published version of the manuscript.

Funding

The writing of this review was supported by the Basic Science Faculty of the “Ecole Polytechnique Fédérale de Lausanne” (EPFL) and by the Department of Chemistry and Biochemistry, University of California (UCLA), Los Angeles.

Acknowledgments

The authors thanks all their past and present co-workers and colleagues for their critical and fruitful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perkin, W.H. LXXIV. On mauveine and allied colouring matters. J. Chem. Soc. Trans. 1879, 35, 717–732. [Google Scholar] [CrossRef] [Green Version]
  2. Travis, A.S. Perkin’s Mauve: Ancestor of the Organic Chemical Industry. Technol. Cult. 1990, 31, 51–82. [Google Scholar] [CrossRef]
  3. Hubner, K. History–150 Years of mauveine. Chem. Unserer Zeit 2006, 40, 274–275. [Google Scholar] [CrossRef]
  4. Liebig, J. Ueber die Zersetzung des Alkohols durch Chlor. Ann. Der Pharm. 1832, 1, 31–32. [Google Scholar] [CrossRef] [Green Version]
  5. Butler, T.C. The introduction of chloral hydrate into medical practice. Bull. Hist. Med. 1970, 44, 168–172. [Google Scholar]
  6. Jones, A.W. Early drug discovery and the rise of pharmaceutical chemistry. Drug Test. Anal. 2011, 3, 337–344. [Google Scholar] [CrossRef]
  7. Ng, R. Drugs: From Discovery to Approval, 3rd ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 1–560. ISBN 1118907272/9781118907276. [Google Scholar]
  8. Inglis, L. Milk of Paradise: A History of Opium; Pan MacMillan: London, UK, 2019; pp. 1–440. ISBN 9781760782276. [Google Scholar]
  9. Lev, E. Traditional healing with animals (zootherapy): Medieval to present-day Levantine practice. J. Ethnopharmacol. 2003, 85, 107–118. [Google Scholar] [CrossRef]
  10. Bade, R.; Chan, H.-F.; Reynisson, J. Characteristics of known drug space, natural products, their derivatives and synthetic drugs. Eur. J. Med. Chem. 2010, 45, 5646–5652. [Google Scholar] [CrossRef]
  11. Li, J.J. History of drug discovery. In Drug Discovery; Li, J.J., Corey, E.J., Eds.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2013; pp. 1–42. [Google Scholar]
  12. Vogel, P. Sustainable Development: The Roles of Carbon and Bio-Carbon; an Introduction to Molecular Sciences; World Scientific Publishing: Hackensack, NJ, USA; London, UK; Singapore, 2022; pp. 1–585. ISBN 978-9811240485/9811240485. [Google Scholar]
  13. Clayden, J.; Greeves, N.; Warren, S. Organic Chemistry, 2nd ed.; Oxford University Press: Oxford, UK, 2012; pp. 1–1234. ISBN 978-0199270293/0199270295. [Google Scholar]
  14. Smith, M.B. March’s Advanced Organic Chemistry: Reactions, Mechanisms, and Structure, 7th ed.; John Wiley & Sons. Inc.: Hoboken, NJ, USA, 2013; pp. 1–2047. ISBN 0470462590/1118854683/9788126556588/9780470462591/9781118854686/9788126556588. [Google Scholar]
  15. Bruice, P.Y. Organic Chemistry, 8th ed.; Pearson Education: Upper Saddle River, NJ, USA, 2016; pp. 1–1344. ISBN 978-0134042282/013404228X. [Google Scholar]
  16. Hartwig, J.F. Organotransition Metal Chemistry. From Bonding to Catalysis; University Science Books: Melville, NY, USA, 2010; pp. 1–1160. ISBN 978-1891389535. [Google Scholar]
  17. Vogel, P.; Houk, K.N. Organic Chemistry: Theory, Reactiviy and Mechanisms in Modern Synthesis; Wiley-VCH Verlag GmbH & Co., KGaA: Weinheim, Germany, 2019; pp. 1–1352. ISBN 9783527345328/3527345329. [Google Scholar]
  18. Anslyn, E.V.; Dougherty, D.A. Modern Physical Organic Chemistry; University Science Books: Sausalito, CA, USA, 2006; pp. 1–1095. ISBN 1891389319/9781891389313. [Google Scholar]
  19. Wang, Z.; Wille, U.; Juaristi, E. (Eds.) Encyclopedia of Physical Organic Chemistry; John Wiley & Sons, Inc.: New York, NY, USA, 2017; Volume 6, pp. 1–4464. ISBN 978-1-118-47045-9. [Google Scholar]
  20. Reymond, J.-L. The chemical space project. Acc. Chem Res. 2015, 48, 727–730. [Google Scholar] [CrossRef] [Green Version]
  21. Awale, M.; Sirockin, F.; Stiefl, N.; Reymond, J.-L. Medicinal chemistry database GDBMedChem. Mol. Inform. 2019, 38, 1900031. [Google Scholar] [CrossRef] [Green Version]
  22. Boström, J.; Brown, D.G. Stuck in a rut with old chemistry. Drug Discov. Today 2016, 21, 701–703. [Google Scholar] [CrossRef] [PubMed]
  23. Brown, D.G.; Boström, J. Analysis of past and present synthetic methodologies on medicinal chemistry: Where have all the new reactions gone? J. Med. Chem. 2016, 59, 4443–4458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Jiang, X.; Hao, X.; Jing, L.; Wu, G.; Kang, D.; Liu, X.; Zhan, P. Recent applications of click chemistry in drug discovery. Expert Opin. Drug Discov. 2019, 14, 779–889. [Google Scholar] [CrossRef] [PubMed]
  25. Takayama, Y.; Kusamori, K.; Nishikawa, M. Click chemistry as a tool for cell engineering and drug delivery. Molecules 2019, 24, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Dreher, S.D. Catalysis in medicinal chemistry. Reaction Chem. Eng. 2019, 4, 1530–1535. [Google Scholar] [CrossRef]
  27. Siddiqui, S.; Sharma, S.; Sharma, B.; Siddiqui, A.A. Role of serendipity in drug discovery. J. Pharm. Res. 2010, 9, 49–55. [Google Scholar] [CrossRef] [Green Version]
  28. Pedreira, J.G.B.; Franco, L.S.; Barreiro, E.J. Chemical intuition in drug design and discovery. Curr. Top. Med. Chem. 2019, 19, 1679–1693. [Google Scholar] [CrossRef]
  29. NIST Chemistry WebBook, NIST Standard Reference Database Number 69, Last Update to Data 2022. Available online: https://webbook.nist.gov/chemistry/ (accessed on 15 May 2022).
  30. Hendrickson, J.B. Systematic synthesis design. 4. Numerical codification of construction reactions. J. Am. Chem. Soc. 1975, 97, 5784–5800. [Google Scholar] [CrossRef]
  31. Newhouse, T.; Baran, P.S.; Hoffmann, R.W. The economies of synthesis. Chem. Soc. Rev. 2009, 38, 3010–3021. [Google Scholar] [CrossRef]
  32. Gaich, T.; Baran, P.S. Aiming for the ideal synthesis. J. Org. Chem. 2010, 75, 4657–4673. [Google Scholar] [CrossRef]
  33. Morawetz, H.; Otaki, P.S. Kinetics and equilibria of amide formation in aqueous media. J. Am. Chem. Soc. 1963, 85, 463–468. [Google Scholar] [CrossRef]
  34. de Azambuja, F.; Parac-Vogt, T.N. Water-tolerant and atom economical amide bond formation by metal-substituted polyoxometalate catalysts. ACS Catal. 2019, 9, 10245–10252. [Google Scholar] [CrossRef]
  35. Perrot, P. A to Z of Thermodynamics; Oxford University Press: Oxford, UK; New York, NY, USA; Tokyo, Japan, 1998; pp. 1–336. ISBN 0198565569/9780198565567. [Google Scholar]
  36. Kernen, P.; Vogel, P. The homoconjugated electron-releasing carbonyl group of 1-methyl-7-oxabicyclo[2.2.1]hept-5-en-2-one. Regioselective syntheses of 5-chloro- and 6-chloro-1-methyl-7-oxabicyclo[2.2.1]hept-5-en-2-one. Helv. Chim. Acta 1993, 76, 2338–2343. [Google Scholar] [CrossRef]
  37. Vogel, P. The electron-releasing carbonyl group. Part 1: Discovery and theoretical studies. Chim. Oggi 1997, 15, 18–25. [Google Scholar]
  38. Vogel, P. The electron-releasing carbonyl group. Part 2. Spectroscopic detection and synthetic applications. Chim. Oggi 1997, 15, 37–43. [Google Scholar] [CrossRef]
  39. Carrupt, P.-A.; Vogel, P. Regioselective additions of electrophiles to olefins remotely perturbed–The carbonyl groups as a homoconjugated electron-donating substituent. Tetrahedron Lett. 1982, 23, 2563–2566. [Google Scholar] [CrossRef]
  40. Carrupt, P.-A.; Vogel, P. The carbonyl group as homoconjugated electron-releasing substituent. Regioselective electrophilic additions at bicyclo[2.2.1]hept-5-en-2-one, bicyclo[2.2.2]oct-5-en-2-one, and derivatives. Helv. Chim. Acta 1989, 72, 1008–1028. [Google Scholar] [CrossRef]
  41. Ruggiu, A.A.; Lysek, R.; Moreno-Clavijo, E.; Moreno-Vargas, A.J.; Robina, I.; Vogel, P. The regioselectivity of the addition of benzeneselenyl chloride to 7-azanorborn-5-ene-2-yl derivatives is controlled by the 2-substituent: New entry into 3- and 4-hydroxy-5-substituted prolines. Tetrahedron 2010, 66, 7309–7315. [Google Scholar] [CrossRef]
  42. Carrupt, P.-A.; Vogel, P. The carbonyl group as homoconjugated electron-donating substituent–Ab initio STO3G MO calculations. Tetrahedron Lett. 1984, 25, 2879–2882. [Google Scholar] [CrossRef]
  43. Sordo, J.A.; Varela-Alvarez, A.; Giani, S.; Vogel, P. Quantum calculations on the acid catalyzed rearrangements of norborn-5-en-2-one, 7-oxanorborn-5-en-2-one and 7-azanorborn-5-en-2-one. Appl. Catal. A General 2008, 336, 72–78. [Google Scholar] [CrossRef]
  44. Le Drian, C.; Vogel, P. Acid-catalyzed rearrangement of 5,6-exo-epoxy-7-oxabicyclo[2.2.1]hept-2-yl derivatives–Migratory apitudes of acyl vs. alkyl groups in Wagner-Meerwein transpositions. Helv. Chim. Acta 1987, 70, 1703–1720. [Google Scholar] [CrossRef]
  45. Berner, D.; Dahn, H.; Vogel., P. Unambiguous proof for alcoxycarbonyl group migration in Wagner-Meerwein rearrangements. Helv. Chim. Acta 1980, 63, 2538–2553. [Google Scholar] [CrossRef]
  46. Scuseria, G.E.; Miller, M.D.; Jensen, F.; Geertsen, J. The dipole moment of carbon monoxide. J. Chem. Phys. 1991, 94, 6660–6663. [Google Scholar] [CrossRef]
  47. De Proft, F.; Martin, J.M.L.; Geerlings, P. On the performance of density functional methods for describing atomic populations, dipole moments and infrared intensities. Chem. Phys. Lett. 1996, 250, 393–401. [Google Scholar] [CrossRef]
  48. Abboud, J.L.M.; Alkorta, I.; Davalos, J.Z.; Müller, P.; Quintanilla, E. Thermodynamic stabilities of carbocations. Adv. Phys. Org. Chem. 2002, 37, 57–135. [Google Scholar] [CrossRef]
  49. Aue, D.H. Carbocations. Wiley Interdiscip. Rev.-Comput. Mol. Sci. 2011, 1, 487–508. [Google Scholar] [CrossRef]
  50. Klump, D.A. Carbocations. In Organic Reaction Mechanisms Series; Knipe, A.C., Ed.; Wiley Online Library, John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2020; Chapter 6; pp. 337–367. [Google Scholar] [CrossRef]
  51. Wodrich, M.D.; Wannere, C.S.; Mo, Y.; Jarowski, P.D.; Houk, K.N.; Schleyer, P.V.R. The concept of protobranching and its many paradigm shifting implications for energy evaluations. Chem. Eur. J. 2007, 13, 7731–7744. [Google Scholar] [CrossRef]
  52. Aizman, A.; Contreras, R.; Galvan, M.; Cedillo, A.; Santos, J.C.; Chamorro, E. The Markovnikov regioselectivity rule in the light of site activation models. J. Phys. Chem. A 2002, 106, 7844–7849. [Google Scholar] [CrossRef]
  53. Hughes, P. Was Markovnikov’s rule an inspired guess? J. Chem. Educ. 2006, 83, 1152–1154. [Google Scholar] [CrossRef]
  54. Kerber, R.C. Markovnikov’s rule. J. Chem. Educ. 2007, 84, 1109. [Google Scholar] [CrossRef] [Green Version]
  55. Ilich, P.P. Markovnikov’s rule-Replies. J. Chem. Educ. 2007, 84, 1109. [Google Scholar] [CrossRef] [Green Version]
  56. Zavitsas, A.A.; Rogers, D.W.; Matsunaga, N. Heats of formation of organic compounds by simple calculation. J. Org. Chem. 2010, 75, 6502–6515. [Google Scholar] [CrossRef] [PubMed]
  57. Dimroth, O. Beziehungen zwischen Affinität und Reaktiongeschwindichkeit. Angew. Chem. 1933, 52, 571–576. [Google Scholar] [CrossRef]
  58. Evans, M.G.; Polanyi, M. Further considerations on the thermodynamics of chemical equilibria and reaction rates. Trans. Faraday Soc. 1936, 32, 1333–1359. [Google Scholar] [CrossRef]
  59. Bell, R.P. The kinetics of proton transfer reactions. Trans. Faraday Soc. 1938, 34, 0229–0236. [Google Scholar] [CrossRef]
  60. Levy, D.E. Arrow-Pushing in Organic Chemistry, 2nd ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2017; pp. 1–400. [Google Scholar] [CrossRef]
  61. Woodward, R.B.; Hoffmann, R. Conservation of orbital symmetry. Angew. Chem. Int. Ed. 1969, 8, 781–853. [Google Scholar] [CrossRef]
  62. Longuet-Higgins, H.C.; Abrahamson, E.W. The electronic mechanism of electrocyclic reactions. J. Am. Chem. Soc. 1965, 87, 2045–2046. [Google Scholar] [CrossRef]
  63. Bickelhaupt, F.M.; Houk, K.N. Analyzing reaction rates with the distortion/interaction-activation strain model. Angew. Chem. Int. Ed. 2017, 56, 10070–10088. [Google Scholar] [CrossRef] [Green Version]
  64. Dewar, M.J.S. Aromaticity and pericyclic reactions. Angew. Chem. Int. Ed. 1971, 10, 761–776. [Google Scholar] [CrossRef]
  65. Zimmerman, H.E. Möbius-Hückel concept in organic chemistry-Application to organic molecules and reactions. Acc. Chem. Res. 1971, 4, 272–280. [Google Scholar] [CrossRef]
  66. Evans, M.G. The activation energies of reactions involving conjugated systems. Trans. Faraday Soc. 1939, 35, 0824–0834. [Google Scholar] [CrossRef]
  67. Houk, K.N. Frontier molecular orbital theory and organic reactions. Nature 1977, 266, 662. [Google Scholar] [CrossRef]
  68. Lias, S.G.; Bartmess, J.E.; Liebman, J.F.; Holmes, J.L.; Levin, R.D.; Mallard, W.G. Gas-Phase Ion and Neutral Thermochemistry. J. Phys. Chem. Reference Data 1988, 17 (Suppl. 1), 861. [Google Scholar] [CrossRef]
  69. Rienstra-Kiracofe, J.C.; Tschumper, G.S.; Schaefer, H.F.I.; Nandi, S.; Ellison, G.B. Atomic and molecular electron affinities: Photoelectron experiments and theoretical computations. Chem. Rev. 2002, 102, 231–282. [Google Scholar] [CrossRef]
  70. Dewar, M.J.S.; Pyron, R.S. Nature of transition state in some Diels-Alder reactions. J. Am. Chem. Soc. 1970, 92, 3098–3103. [Google Scholar] [CrossRef]
  71. Dewar, J.J.S.; Pierini, A.B. Mechanism of the Diels-Alder reaction. Studies of the addition of maleic anhydride to furan and methylfurans. J. Am. Chem. Soc. 1984, 106, 203–208. [Google Scholar] [CrossRef]
  72. Pedersen, S.; Herek, J.L.; Zewail, A.H. The validity of the diradical hypothesis-Direct femtosecond studies of the transition-state structures. Science 1994, 266, 1359–1364. [Google Scholar] [CrossRef]
  73. Woodward, R.B.; Katz, T.J. The mechanism of the Diels-Alder reaction. Tetrahedron 1959, 5, 70–89. [Google Scholar] [CrossRef]
  74. Horn, B.A.; Herek, J.L.; Zewail, A.H. Retro-Diels-Alder femtosecond reaction dynamics. J. Am. Chem. Soc. 1996, 118, 8755–8756. [Google Scholar] [CrossRef]
  75. Littmann, E.R. The mechanism of the diene synthesis. J. Am. Chem. Soc. 1936, 58, 1316–1317. [Google Scholar] [CrossRef]
  76. Houk, K.N.; Gonzalez, J.; Li, Y. Pericyclic reaction transition-states-Passions and punctilios, 1935–1995. Acc. Chem. Res. 1995, 28, 81–90. [Google Scholar] [CrossRef]
  77. Houk, K.N. The frontier molecular orbital theory of cycloaddition reactions. Acc. Chem. Res. 1975, 8, 361–369. [Google Scholar] [CrossRef]
  78. Chen, P.-P.; Ma, P.; He, X.; Svatunek, D.; Liu, F.; Houk, K.N. Computational exploration of ambiphilic reactivity of azides and Sustmann’s paradigmatic parabola. J. Org. Chem. 2021, 86, 5792–5804. [Google Scholar] [CrossRef] [PubMed]
  79. Gajewski, J.J.; Conrad, N.D. Variable transition-state structure in the Cope rearrangement as deduced from secondary deuterium kinetic isotope effects. J. Am. Chem. Soc. 1978, 100, 6269–6270. [Google Scholar] [CrossRef]
  80. Hosomi, A.; Sakurai, H. Syntheses of γ,δ-unsaturated alcohols from allylsilanes and carbonyl compounds in presence of Ti tetrachloride. Tetrahedron Lett. 1976, 17, 1295–1298. [Google Scholar] [CrossRef]
  81. Hosomi, A. Characteristics in the reactions of allylsilanes and their applications to versatile synthetic equivalents. Acc. Chem. Res. 1988, 21, 200–206. [Google Scholar] [CrossRef]
  82. Biamonte, M.A. Sakurai allylation. In Name Reactions for Homologation, Part I; Li, J.J., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009; pp. 539–575. ISBN 0470487011/9780470487013. [Google Scholar]
  83. Hayashi, T.; Kabeta, K.; Hamachi, I.; Kumada, M. Erythro-selectivity in addition of γ-substituted allylsilanes to aldehydes in the presence of Ti tetrachloride. Tetrahedron Lett. 1983, 24, 2865–2868. [Google Scholar] [CrossRef]
  84. Deleris, G.; Dunoguès, J.; Calas, R. Synthèse d’alcools γ,δ-éthyléniques ou propargyliques par voie organosilicique. Tetrahedron Lett. 1976, 28, 2449–2450. [Google Scholar] [CrossRef]
  85. Hoffmann, R.W.; Zeiss, H.J. Stereoselective synthesis of alcohols. 8. Diastereoselective synthesis of β-methylhomoallyl alcohols via crotylboronates. J. Org. Chem. 1981, 46, 1309–1314. [Google Scholar] [CrossRef]
  86. Collum, D.B.; McDonald, J.H.; Still, W.C. Synthesis of the polyether antibiotic Monensin. 2. Preparation of intermediates. J. Am. Chem. Soc. 1980, 102, 2118–2120. [Google Scholar] [CrossRef]
  87. Sato, F.; Iida, K.; Iijima, S.; Moriya, H.; Sato, M. Threo-selective synthesis of β-methylhomoallyl alcohols via but-2-enyltitanium compounds. J. Chem. Soc. Chem. Commun. 1981, 1140–1141. [Google Scholar] [CrossRef]
  88. Yamamoto, Y.; Maruyama, K. Crotylzirconium derivatives as a new reagent for the threo-selective synthesis of β-methylhomoallyl alcohols. Tetrahedron Lett. 1981, 22, 2895–2898. [Google Scholar] [CrossRef]
  89. Hiyama, T.; Kimura, K.; Nozaki, H. Chromium(II) mediated threo-selective synthesis of homoallyl alcohols. Tetrahedron Lett. 1981, 22, 1037–1040. [Google Scholar] [CrossRef]
  90. Yamamoto, Y.; Yatagai, H.; Naruta, Y.; Maruyama, K. Erythro-selective addition of crotyltrialkyltins to aldehydes regardless of the geometry of the crotyl unit-Stereoselection independent of the stereochemistry of precursors. J. Am. Chem. Soc. 1980, 102, 7107–7109. [Google Scholar] [CrossRef]
  91. Shaik, S.S.; Pross, A. SN2 Reactivity of CH3X derivatives-A valence bond approach. J. Am. Chem. Soc. 1982, 104, 2708–2719. [Google Scholar] [CrossRef]
  92. Pross, A. Relationship between rates and equilibria and the mechanistic significance of the Brønsted parameter-A qualitative valence-bond approach. J. Org. Chem. 1984, 49, 1811–1818. [Google Scholar] [CrossRef]
  93. Shaik, S.S.; Schlegel, H.B.; Wolfe, S. Theoretical Aspects of Physical Organic Chemistry. The SN2 Mechanism; John Wiley & Sons Inc.: New York, NY, USA, 1992; pp. 1–285. ISBN 0471840416. [Google Scholar]
  94. Shaik, S.S.; Ioffe, A.; Reddy, A.C.; Pross, A. Is the avoided crossing state a good approximation for the transition state of a chemical reaction-An analysis of Menschutkin and ionic SN2 reactions. J. Am. Chem. Soc. 1994, 116, 262–273. [Google Scholar] [CrossRef]
  95. Pross, A. Theoretical & Physical Principles of Organic Reactivity; John Wiley & Sons, Inc.: New York, NY, USA, 1995; pp. 1–312. ISBN 0471555991/9780471555995. [Google Scholar]
  96. Anglada, J.M.; Besalú, E.; Bofill, J.M.; Crehuet, R. Prediction of approximate transition states by Bell-Evans-Polanyi principle: I. J. Comput. Chem. 1999, 20, 1112–1129. [Google Scholar] [CrossRef]
  97. Bofill, J.M.; Anglada, J.M.; Besalú, E.; Crehuet, R. Quantum chemical reactivity: Beyond the study of small molecules. In Fundamentals of Molecular Similarity; Springer: Boston, MA, USA, 2001; pp. 125–141. [Google Scholar] [CrossRef]
  98. Wester, R.B.A.E.; Davis, A.V.; Neumark, D.M. Time-resolved study of the symmetric SN2-reaction I-+CH3I. J. Chem. Phys. 2003, 119, 10032–10039. [Google Scholar] [CrossRef]
  99. Smith, M.B. Aliphatic Substitution, Nucleophilic and Organometallic. In March’s Advanced Organic Chemistry, Reactions, Mechanisms, and Structure; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013; Chapter 10; pp. 373–568. [Google Scholar] [CrossRef]
  100. Crampton, M.R. Electrophilic aromatic substitution. In Organic Reaction Mechanisms Series; Knipe, A.C., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2014; Chapter 6; pp. 257–283. [Google Scholar] [CrossRef]
  101. Rowlands, G.J. Radicals in organic synthesis: Part 2. Tetrahedron 2010, 66, 1593–1636. [Google Scholar] [CrossRef]
  102. Tebben, L.; Studer, A. Nitroxides: Applications in synthesis and in polymer chemistry. Angew. Chem. Int. Ed. 2011, 50, 5034–5068. [Google Scholar] [CrossRef] [PubMed]
  103. Quiclet-Sire, B.; Zard, S.Z. Fun with radicals: Some new perspectives for organic synthesis. Pure Appl. Chem. 2011, 83, 519–551. [Google Scholar] [CrossRef] [Green Version]
  104. Lapointe, G.; Kapat, A.; Weidner, K.; Renaud, P. Radical azidation reactions and their application in the synthesis of alkaloids. Pure Appl. Chem. 2012, 84, 1633–1641. [Google Scholar] [CrossRef] [Green Version]
  105. Sebren, L.J.; Devery, J.J.; Stephenson, C.R. Catalytic radical domino reactions in organic synthesis. ACS Catal. 2014, 4, 703–716. [Google Scholar] [CrossRef] [Green Version]
  106. Denes, F.; Pichowicz, M.; Povie, G.; Renaud, P. Thiyl radicals in organic synthesis. Chem. Rev. 2014, 114, 2587–2693. [Google Scholar] [CrossRef]
  107. Smith, M.B. Addition to carbon-carbon multiple bonds. In March’s Advanced Organic Chemistry, Reactions, Mechanisms, and Structure; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013; Chapter 15; pp. 859–902. [Google Scholar] [CrossRef]
  108. Trost, B.M. The atom economy—A search for synthetic efficiency. Science 1991, 254, 1471–1477. [Google Scholar] [CrossRef]
  109. Shin, I.; Montgomery, T.P.; Krische, M.J. Catalytic C-C bond formation and the Hendricksonian ideal: Atom- and redox-economy, stereo- and site-selectivity. Aldrichimica Acta 2015, 48, 15. [Google Scholar]
  110. Smith, M.B. March’s Advanced Organic Chemistry: Reactions, Mechanisms, and Structure, 8th ed.; John Wiley & Sons. Inc.: Hoboken, NJ, USA, 2022; Chapters 15 & 16; pp. 891–1272. ISBN 978-1119371809/1119371805. [Google Scholar]
  111. Arundale, E.; Mikeska, L.A. The olefin aldehyde condensation-The Prins reaction. Chem. Rev. 1952, 51, 505–555. [Google Scholar] [CrossRef]
  112. Overman, L.E.; Velthuisen, E.J. Scope and facial selectivity of the Prins-pinacol synthesis of attached rings. J. Org. Chem. 2006, 71, 1581–1587. [Google Scholar] [CrossRef]
  113. Li, C.J. The development of catalytic nucleophilic additions of terminal alkynes in water. Acc. Chem. Res. 2010, 43, 581–590. [Google Scholar] [CrossRef]
  114. Engle, K.M.; Mei, T.S.; Wasa, M.; Yu, J.Q. Weak coordination as a powerful means for developing broadly useful C-H functionalization reactions. Acc. Chem. Res. 2012, 45, 788–802. [Google Scholar] [CrossRef] [PubMed]
  115. Colby, D.A.; Tsai, A.S.; Bergman, R.G.; Ellman, J.A. Rhodium-catalyzed chelation-assisted C-H bond functionalization reactions. Acc. Chem. Res. 2012, 45, 814–825. [Google Scholar] [CrossRef] [PubMed]
  116. Rogge, T.; Kaplaneris, N.; Chatani, N.; Kim, J.; Chang, S.; Punji, B.; Schafer, L.L.; Musaev, D.G.; Wencel-Delord, J.; Roberts, C.A.; et al. C-H activation. Nat. Rev. Methods Primers 2021, 1, 43. [Google Scholar] [CrossRef]
  117. Jang, H.Y.; Huddleston, R.R.; Krische, M.J. Reductive generation of enolates from enones using elemental hydrogen: Catalytic C-C bond formation under hydrogenative conditions. J. Am. Chem. Soc. 2002, 124, 15156–15157. [Google Scholar] [CrossRef] [PubMed]
  118. Jang, H.Y.; Krische, M.J. Catalytic hydrogen-mediated cross-coupling of enones and carbonyl compounds: Aldol condensation by hydrogenation. Eur. J. Org. Chem. 2004, 2004, 3953–3958. [Google Scholar] [CrossRef]
  119. Ngai, M.Y.; Kong, J.R.; Krische, M.J. Hydrogen-mediated C-C bond formation: A broad new concept in catalytic C-C coupling. J. Org. Chem. 2007, 72, 1063–1072. [Google Scholar] [CrossRef]
  120. Nguyen, K.D.; Park, B.Y.; Luong, T.; Sato, H.; Garza, V.J.; Krische, M.J. Metal-catalyzed reductive coupling of olefin-derived nucleophiles: Reinventing carbonyl addition. Science 2016, 354, 300. [Google Scholar] [CrossRef]
  121. Holmes, M.; Schwartz, L.A.; Krische, M.L. Intermolecular metal-catalyzed reductive coupling of dienes, allenes, and enynes with carbonyl compounds and imines. Chem. Rev. 2018, 118, 6026–6052. [Google Scholar] [CrossRef]
  122. Ngai, M.Y.; Rucas, E.; Krische, M.J. Ruthenium-catalyzed C-C bond formation via transfer hydrogenation: Branch-selective reductive coupling of allenes to paraformaldehyde and higher aldehydes. Org. Lett. 2008, 10, 2705–2708. [Google Scholar] [CrossRef] [Green Version]
  123. Shibahara, F.; Bower, J.F.; Krische, M.J. Ruthenium-catalyzed C-C bond forming transfer hydrogenation: Carbonyl allylation from the alcohol or aldehyde oxidation level employing acyclic 1,3-dienes as surrogates to preformed allyl metal reagents. J. Am. Chem. Soc. 2008, 130, 6338–6339. [Google Scholar] [CrossRef] [Green Version]
  124. Sam, B.; Breit, B.; Krische, M.J. Paraformaldehyde and methanol as C1 feedstocks in metal-catalyzed C-C couplings of -unsaturated reactants: Beyond hydroformylation. Angew. Chem. Int. Ed. 2015, 54, 3267–3274. [Google Scholar] [CrossRef] [PubMed]
  125. Geary, L.M.; Woo, S.K.; Leung, J.C.; Krische, M.J. Diastereo- and enantioselective Ir-catalyzed carbonyl propargylation from the alcohol or aldehyde oxidation level: 1,3-Enynes as allenylmetal equivalents. Angew. Chem. Int. Ed. 2012, 51, 2972–2976. [Google Scholar] [CrossRef] [PubMed]
  126. Nguyen, K.D.; Herkommer, D.; Krische, M.J. Ruthenium-BINAP catalyzed alcohol C-H tert-prenylation via 1,3-enyne transfer hydrogenation: Beyond stoichiometric carbanions in enantioselective carbonyl propargylation. J. Am. Chem. Soc. 2016, 138, 5238–5241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Speith, J.G. The Chemistry and Technology of Petroleum; CRC Press, Francis & Taylor Group: Boca Raton, FL, USA, 2006; pp. 1–984. ISBN 1420008382/9781420008388. [Google Scholar]
  128. Corma, A.; Martínez, A. Chemistry, catalysts, and processes for isoparaffin–olefin alkylation: Actual situation and future trends. Rev. Sci. Eng. 1993, 35, 483–570. [Google Scholar] [CrossRef]
  129. Li, X.; Nagoaka, K.; Simon, L.J.; Olindo, R.; Lercher, J.A. Mechanism of butane skeletal isomerization on sulfated zirconia. J. Catal. 2005, 232, 456–466. [Google Scholar] [CrossRef]
  130. Wang, P.; Zhang, J.; Wang, G.; Li, C.; Ang, C. Nature of active sites and deactivation mechanism for n-butane isomerization over alumina-promoted sulfated zirconia. J. Catal. 2016, 338, 124–134. [Google Scholar] [CrossRef]
  131. Olah, G.A.; Farooq, O.; Husain, A.; Ding, N.; Trivedi, N.J.; Olah, J.A. Superacid FSO3H/HF catalyzed butane isomerization. Catal. Lett. 1991, 10, 239–247. [Google Scholar] [CrossRef]
  132. Kennedy, J.P. Living cationic polymerization of olefins. How did the discovery come about. J. Polym. Sci. Part A Polym. Chem. 1999, 37, 2285–2293. [Google Scholar] [CrossRef]
  133. Puskas, J.E.; Peng, H.H. Kinetic simulation of living carbocationic polymerizations. I. Simulation of living isobutylene polymerization. Polym. React. Eng. 1999, 7, 553–576. [Google Scholar] [CrossRef]
  134. Rudin, A.; Choi, P. The Elements of Polymer Science and Engineering, 3rd ed.; Academic Press: New York, NY, USA; Elsevier: Amsterdam, The Netherlands, 2013; pp. 1–563. ISBN 978-0-12-382178-2. [Google Scholar]
  135. Vasilenko, I.V.; Yeong, H.Y.; Delgado, M.; Ouardad, S.; Peruch, F.; Voit, B.; Ganachaud, F.; Kostjuk, S.V. A catalyst platform for unique cationic (co)polymerization in aqueous emulsion. Angew. Chem. Int. Ed. 2015, 54, 12728–12732. [Google Scholar] [CrossRef]
  136. Ouardad, S.; Wirotius, A.L.; Kostjuk, S.; Ganachaud, F.; Peruch, F. Carbocationic polymerization of isoprene using cumyl initiators: Progress in understanding side reactions. RSC Adv. 2015, 5, 59218–59225. [Google Scholar] [CrossRef]
  137. Goriainow, W.; Butlerow, A. Ueber die Polyolene und die Umwandlung von Aethylen in Aethylalkohol. Liebigs Ann. Chem. 1873, 169, 146–149. [Google Scholar] [CrossRef] [Green Version]
  138. Kostjuk, S.V. Recent progress in the Lewis acid co-initiated cationic polymerization of isobutylene with 1,3-dienes. RSC Advances 2015, 5, 13125–13144. [Google Scholar] [CrossRef]
  139. Rajasekhar, T.; Singh, G.; Kapur, G.S.; Ramakumar, S.S.V. Recent advances in catalytic chain transfer polymerization of isobutylene: A review. RSC Adv. 2020, 10, 18180–18191. [Google Scholar] [CrossRef]
  140. Mattay, J. Photo-induced electron-transfer in organic synthesis. Synthesis 1989, 1989, 233–256. [Google Scholar] [CrossRef]
  141. Garcia, H.; Roth, H.D. Generation and reactions of organic radical-cations in zeolites. Chem. Rev. 2002, 102, 3977–4007. [Google Scholar] [CrossRef] [PubMed]
  142. Broggi, J.; Rollet, M.; Clément, J.-L.; Canard, G.; Terme, T.; Gigmes, D.; Vanelle, P. Polymerization initiated by organic electron donors. Angew. Chem. Int. Ed. 2016, 55, 5994–5999. [Google Scholar] [CrossRef]
  143. Elschenbroich, C. Organometallics, 3rd Completely Revised and Extended Edition; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2006; pp. 1–817. ISBN 978-3-527-29390-2. [Google Scholar]
  144. Ishii, Y.; Tsutsui, M. Organotransition-Metal Chemistry; Springer: Boston, MA, USA, 1975; pp. 1–398. [Google Scholar] [CrossRef]
  145. Crabtree, R.H. An organometallic future in green and energy chemistry? Organometallics 2011, 30, 17–19. [Google Scholar] [CrossRef]
  146. Mathey, F. Transition Metal Organometallic Chemistry; SpringerBriefs in Molecular Sciences; Springer: Singapore, 2013; pp. 1–100. ISBN 978-981-4451-09-3. [Google Scholar]
  147. Frenking, G. Chemical bonding in transition metals. In The Chemical Bond: Chemical Bonding Across the Periodic Table; Frenking, G., Sason, S., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2014; pp. 175–218. [Google Scholar] [CrossRef]
  148. Tolman, C.A. 16 and 18 electron rule in organometallic chemistry and homogeneous catalysis. Chem. Soc. Rev. 1972, 1, 337–353. [Google Scholar] [CrossRef]
  149. Carneiro, J.W.d.M.; Schleyer, P.V.R.; Saunders, M.; Remington, R.; Schaefer, H.F., III; Rauk, A.; Sorensen, T.S. Protonated ethane. A theoretical investigation of C2H7+ structures and energies. J. Am. Chem. Soc. 1994, 116, 3483–3493. [Google Scholar] [CrossRef]
  150. Smith, D.; Adams, N.G.; Alge, E. Isotope exchange and collisional association in the reactions of CH3+ and its deuterated analogs with H2, HD, and D2. J. Chem. Phys. 1982, 77, 1261–1268. [Google Scholar] [CrossRef]
  151. White, E.T.; Tang, J.; Oka, T. CH5+: The infrared spectrum observed. Science 1999, 284, 135–137. [Google Scholar] [CrossRef] [PubMed]
  152. Huang, X.C.; Johnson, L.M.; Bowman, J.M.; McCoy, A.B. Deuteration effects on the structure and infrared spectrum of CH5+. J. Am. Chem. Soc. 2006, 128, 3478–3479. [Google Scholar] [CrossRef]
  153. Hinkle, C.E.; McCoy, A.B. Isotopic effects on the dynamics of the CH3+ + H2 → CH5+ → CH3+ + H2 reaction. J. Phys. Chem. A 2012, 116, 4687–4694. [Google Scholar] [CrossRef] [PubMed]
  154. Asvany, O.; Yamada, K.M.T.; Bruenken, S.; Potapov, A.; Schlemmer, S. Experimental ground-state combination differences of CH5+. Science 2015, 347, 1346–1349. [Google Scholar] [CrossRef] [PubMed]
  155. Saunders, M.; Vogel, P.; Hagen, E.L.; Rosenfeld, J. Evidence for protonated cyclopropane intermediates from studies of stable solutions of carbonium ions. Acc. Chem. Res. 1973, 6, 53–59. [Google Scholar] [CrossRef]
  156. Vogel, P. Carbocation Chemistry. In Studies in Organic Chemistry; Elsevier: Amsterdam, The Netherlands, 1985; pp. 1–596. ISBN 0444425225. [Google Scholar]
  157. Cabrera, J.M.; Krische, M.J. Total synthesis of clavosolide A via asymmetric alcohol-mediated carbonyl allylation: Beyond protecting groups or chiral auxiliaries in polyketide construction. Angew. Chem. Int. Ed. 2019, 58, 10718–10722. [Google Scholar] [CrossRef]
  158. Della-Felice, F.; Sarotti, A.M.; Krische, M.J.; Pilli, R.A. Total synthesis and structural validation of phosdiecin A via asymmetric alcohol-mediated carbonyl reductive coupling. J. Am. Chem. Soc. 2019, 141, 13778–13782. [Google Scholar] [CrossRef]
  159. McLafferty, F.W. Mass Spectrometry of Organic Ions; eBook; Academic Press: London, UK; New York, NY, USA, 1963; p. 700. ISBN 9780323142779. [Google Scholar]
  160. Gerlich, D. Probing the structure of CH5+ ions and deuterated variants via collisions. Phys. Chem. Chem. Phys. 2005, 7, 1583–1591. [Google Scholar] [CrossRef]
  161. Collins, C.J. The pinacol rearrangement. Quart. Rev. 1960, 14, 357–377. [Google Scholar] [CrossRef]
  162. Dewar, M.J.S.; Rzepa, H.S. Gaseous ions. 4. MINDO-3 calculations for some simple organic cations and for their hydrogen elimination-reactions. J. Am. Chem. Soc. 1977, 99, 7432–7439. [Google Scholar] [CrossRef]
  163. Lischka, H.; Köhler, H.-J. Structure and stability of carbocations C2H3+ and C2H4X+, X = H, F, Cl, Cl, and CH3 –ab initio investigations including electron correlation and comparison with MINDO-3 results. J. Am. Chem. Soc. 1978, 100, 5297. [Google Scholar] [CrossRef]
  164. Andrei, H.-S.; Solca, N.; Dopfer, O. IR spectrum of the ethyl cation: Evidence for the nonclassical structure. Angew. Chem. Int. Ed. 2008, 47, 395–397. [Google Scholar] [CrossRef] [PubMed]
  165. Ricks, A.M.; Douberly, G.E.; Schleyer, P.v.R.; Duncan, M.A. Infrared spectroscopy of protonated ethylene: The nature of proton binding in the non-classical structure. Chem. Phys. Lett. 2009, 480, 17–20. [Google Scholar] [CrossRef]
  166. Saunders, M.; Hagen, E.L. Rearrangement reactions of secondary carbonium Ions. Isopropyl cation. J. Am. Chem. Soc. 1968, 90, 6881–6882. [Google Scholar] [CrossRef]
  167. Olah, G.A.; White, A.M. Stable carbonium ions. XCI: Carbon-13 nuclear magnetic resonance spectroscopy study of carbonium ions. J. Am. Chem. Soc. 1969, 91, 5801–5810. [Google Scholar] [CrossRef]
  168. McAdoo, D.J.; McLafferty, F.W.; Bente, P.F., III. Ion cyclotron resonance spectroscopy in structure determination. II. Propyl ions. J. Am. Chem Soc. 1972, 94, 2027–2033. [Google Scholar] [CrossRef]
  169. Chiavarino, B.; Crestoni, M.E.; Fornarini, S.; Lemaire, J.; Mac Aleese, L.; Maître, P. Infrared absorption features of gaseous isopropyl carbocations. ChemPhysChem 2004, 5, 1679–1685. [Google Scholar] [CrossRef]
  170. Saunders, M.; Hagen, E.L.; Rosenfeld, J. Rearrangement reactions of secondary carbonium ions. Protonated cyclopropane intermediates formed from sec-butyl cation. J. Am. Chem. Soc. 1968, 90, 6882–6884. [Google Scholar] [CrossRef]
  171. Lee, C.C.; Zohdi, H.F. Solvolytic studies with 2-methyl-[1-13C]-1-propyl tosylate. Can. J. Chem. 1983, 61, 2092–2094. [Google Scholar] [CrossRef]
  172. Fiaux, A.; Smith, D.L.; Futrell, J.H. Ion-molecule reaction of CH3+, C2H5+, C3H5+ and C3H7+ with unsaturated C2, C3 and C4 hydrocarbons. Int. J. Mass Spectrom. Ion Phys. 1977, 25, 281–294. [Google Scholar] [CrossRef]
  173. Saunders, M.; Hagen, E.L. Rapid rearrangements in t-amyl cation and relative sign of coupling constants. J. Am. Chem. Soc. 1968, 90, 2436–2437. [Google Scholar] [CrossRef]
  174. Saunders, M.; Budiansky, S.P. Computer modeling of multistep carbonium-ion rearrangements-Applications to methylcyclopentyl and tert-amyl. Tetrahedron 1979, 35, 929–932. [Google Scholar] [CrossRef]
  175. Scholz, F.; Himmel, D.; Heinemann, F.W.; Schleyer, P.V.R.; Meyer, K.; Krossing, I. Crystal structure determination of the nonclassical 2-norbornyl cation. Science 2013, 341, 62–64. [Google Scholar] [CrossRef] [PubMed]
  176. Olah, G.A.; Lukas, J. Stable carbonium ions. 54. Protonation of and hydride ion abstraction from cycloalkanes and polycycloalkanes in fluorosulfonic acid-antimony pentafluoride. J. Am. Chem. Soc. 1968, 90, 933–938. [Google Scholar] [CrossRef]
  177. Olah, G.A.; Reddy, V.P.; Prakash, G.K.S. Long-lived cyclopropylcarbinyl cations. Chem. Rev. 1992, 92, 69–95. [Google Scholar] [CrossRef]
  178. Mo, Y.; Schleyer, P.v.R.; Jiao, H.; Lin, Z. Quantitative evaluation of hyperconjugation in the cyclopropylcarbinyl cation and in cyclopropylborane. Chem. Phys. Lett. 1997, 280, 439–443. [Google Scholar] [CrossRef]
  179. Prakash, G.K.S.; Reddy, V.P.; Rasul, G.; Casanova, J.; Olah, G.A. The search for persistent cyclobutylmethyl cations in superacidic media and observation of the cyclobutyldicyclopropylmethyl cation. J. Am. Chem. Soc. 1998, 120, 13362–13365. [Google Scholar] [CrossRef]
  180. Reddy, V.P.; Rasul, G.; Prakash, G.K.S.; Olah, G.A. Structural studies of nonclassical cyclobutylmethyl cations by the ab initio method. J. Org. Chem. 2007, 72, 3076–3080. [Google Scholar] [CrossRef]
  181. Franke, W.; Schwarz, H.; Thies, H.; Chandrasekhar, J.; Schleyer, P.V.R.; Hehre, W.J.; Saunders, M.; Walker, G. An experimental and theoretical-study of the degenerate carbon skeleton isomerization of the cyclopentyl cation in the gas-phase. Angew. Chem. Int. Ed. 1980, 19, 485–487. [Google Scholar] [CrossRef]
  182. Franke, W.; Schwarz, H.; Thies, H.; Chandrasekhar, J.; Schleyer, P.V.R.; Hehre, W.J.; Saunders, M.; Walker, G. Entartete Isomerisierung via Kohlenstoff-Platzwechsel beim Cyclopentyl-Kation in der Gasphase. Experimenteller und theoretischer Nachweis der Existenz eines pyramidalen C5H9+-Kations bei der unimolekularen Ethylen-Eliminierung. Chem. Ber. Rec. 1981, 114, 2808–2824. [Google Scholar] [CrossRef]
  183. Franke, W.; Frenking, G.; Schwarz, H.; Wolfschütz, R. Zur Kohlenstoff-Äquilibrierung in cyclischen C6H11+-Kationen in der Gasphase und zum Mechanismus der unimolekularen Ethylen-Abspaltung. Chem. Ber. Rec. 1981, 114, 3878–3895. [Google Scholar] [CrossRef]
  184. Cecchi, P.; Pizzabiocca, A.; Renzi, G.; Grandinetti, F.; Sparapani, C.; Buzek, P.; Schleyer, P.V.R.; Speranza, M. Gas-phase protonation of spiropentane-A novel entry into the C5H9+ potential-energy surface. J. Am. Chem. Soc. 1993, 115, 10338–10347. [Google Scholar] [CrossRef]
  185. Szabo, K.J.; Cremer, D. Route to a kinetically stabilized protonated spirocyclopentane with a pentacoordinated carbon-atom-the missing link between bicyclo[3.2.0]hept-3-yl and 7-norbornyl cation. J. Org. Chem. 1995, 60, 2257–2259. [Google Scholar] [CrossRef]
Figure 1. Hyperconjugation in acyl halides (donation from the carbonyl group n(CO:) non-bonded electron-pairs to the σ(C-X) bond) competes with the n(X:)/π(C=O) conjugation. This competition exists also in carboxylic esters and carboxamides. Reprinted with permission from Ref. [17]. Copyright 2019 Wiley.
Figure 1. Hyperconjugation in acyl halides (donation from the carbonyl group n(CO:) non-bonded electron-pairs to the σ(C-X) bond) competes with the n(X:)/π(C=O) conjugation. This competition exists also in carboxylic esters and carboxamides. Reprinted with permission from Ref. [17]. Copyright 2019 Wiley.
Catalysts 12 00758 g001
Scheme 1. Homoconjugated carbonyl group can be electron-releasing. Norborn-5-en-2-one (1) adds to electrophiles E-X giving the corresponding adducts 3 with high stereo- and regioselectivity arising from the quenching of cationic intermediates 2 by the nucleophiles X-. 2-Cyanonorborn-5-en-2-yl acetate (4) adds to electrophiles E-X with opposite regioselectivity giving the corresponding adducts 6 arising from the quenching of cationic intermediate 5 by the nucleophiles X-. In cationic intermediates 2 the positively charged center C(6) is homoconjugated with the carbonyl group whereas in cationic intermediates 5 the positive charge prefers to reside on C(5), away from the two electron-withdrawing substituents CN and AcO.
Scheme 1. Homoconjugated carbonyl group can be electron-releasing. Norborn-5-en-2-one (1) adds to electrophiles E-X giving the corresponding adducts 3 with high stereo- and regioselectivity arising from the quenching of cationic intermediates 2 by the nucleophiles X-. 2-Cyanonorborn-5-en-2-yl acetate (4) adds to electrophiles E-X with opposite regioselectivity giving the corresponding adducts 6 arising from the quenching of cationic intermediate 5 by the nucleophiles X-. In cationic intermediates 2 the positively charged center C(6) is homoconjugated with the carbonyl group whereas in cationic intermediates 5 the positive charge prefers to reside on C(5), away from the two electron-withdrawing substituents CN and AcO.
Catalysts 12 00758 sch001
Scheme 2. Many (2 + 2)-cycloadditions and (2 + 2)-cycloreversions involve 1,4-diradical intermediates.
Scheme 2. Many (2 + 2)-cycloadditions and (2 + 2)-cycloreversions involve 1,4-diradical intermediates.
Catalysts 12 00758 sch002
Figure 2. Spectrum of mechanisms for the Diels already Alder reaction. In polar medium and for polar substituents, the diradical intermediates D1 and D2 can be zwitterions. ‡ for transition states.
Figure 2. Spectrum of mechanisms for the Diels already Alder reaction. In polar medium and for polar substituents, the diradical intermediates D1 and D2 can be zwitterions. ‡ for transition states.
Catalysts 12 00758 g002
Figure 3. The valence-bond diradicaloid model for transition states of concerted Diels–Alder reactions. The lower is the sum EI(diene) + (−EA(dienophile)), or EI(dienophile) + (−EA(diene)); the faster is the cycloaddition; the higher is the degree of concert.
Figure 3. The valence-bond diradicaloid model for transition states of concerted Diels–Alder reactions. The lower is the sum EI(diene) + (−EA(dienophile)), or EI(dienophile) + (−EA(diene)); the faster is the cycloaddition; the higher is the degree of concert.
Catalysts 12 00758 g003
Figure 4. More O’Ferrall–Jencks diagram for Cope and Claisen rearrangements. The horizontal coordinate represents the shortening of the C(1)-C(6) distance; the vertical coordinate represents the lengthening of the σ(C(3)-C(4)) bond; the enthalpy is perpendicular to the plane made by the two former coordinates. The position of the transition states (‡) in this diagram are deduced according to Gajewski from the comparison of kinetic and equilibrium deuterium isotopic effects.
Figure 4. More O’Ferrall–Jencks diagram for Cope and Claisen rearrangements. The horizontal coordinate represents the shortening of the C(1)-C(6) distance; the vertical coordinate represents the lengthening of the σ(C(3)-C(4)) bond; the enthalpy is perpendicular to the plane made by the two former coordinates. The position of the transition states (‡) in this diagram are deduced according to Gajewski from the comparison of kinetic and equilibrium deuterium isotopic effects.
Catalysts 12 00758 g004
Figure 5. Diradicaloid model for the variable transition states of the (3,3)-sigmatropic rearrangements.
Figure 5. Diradicaloid model for the variable transition states of the (3,3)-sigmatropic rearrangements.
Catalysts 12 00758 g005
Figure 6. The diradicaloid model illustrated with Equilibria (10) and (11) permits to explain the catalytic effect of added protic or Lewis acids (A) to enolate aldol condensations, to metalla-carbonyl-ene reactions, and the carbonyl-ene-reaction.
Figure 6. The diradicaloid model illustrated with Equilibria (10) and (11) permits to explain the catalytic effect of added protic or Lewis acids (A) to enolate aldol condensations, to metalla-carbonyl-ene reactions, and the carbonyl-ene-reaction.
Catalysts 12 00758 g006
Figure 7. Classical representation of the transition state of a SN2 displacement reaction.
Figure 7. Classical representation of the transition state of a SN2 displacement reaction.
Catalysts 12 00758 g007
Scheme 3. Example of a regio-, stereo-, and enantioselective hydro(hydroxycarbation) of an alkene.
Scheme 3. Example of a regio-, stereo-, and enantioselective hydro(hydroxycarbation) of an alkene.
Catalysts 12 00758 sch003
Figure 8. Classical general mechanisms for catalyzed hydrocarbations of unsaturated compounds. Polymerization of the unsaturated compound is a concurrent reaction.
Figure 8. Classical general mechanisms for catalyzed hydrocarbations of unsaturated compounds. Polymerization of the unsaturated compound is a concurrent reaction.
Catalysts 12 00758 g008
Scheme 4. Transition metal-catalyzed hydrocarbation of unsaturated systems.
Scheme 4. Transition metal-catalyzed hydrocarbation of unsaturated systems.
Catalysts 12 00758 sch004
Scheme 5. 14, 16 VE transition metal complexes undergo oxidative additions to σ(X-H) bonds, as do carbene or carbenium ions (unsaturated 6 VE species).
Scheme 5. 14, 16 VE transition metal complexes undergo oxidative additions to σ(X-H) bonds, as do carbene or carbenium ions (unsaturated 6 VE species).
Catalysts 12 00758 sch005
Scheme 6. 14, 16 VE transition complexes form π-complexes with unsaturated systems that can be seen as metallacyclopropanes, just as carbenes and carbenium ions (6 VE species) do with alkenes, generating cyclopropanes and protonated cyclopropanes, respectively.
Scheme 6. 14, 16 VE transition complexes form π-complexes with unsaturated systems that can be seen as metallacyclopropanes, just as carbenes and carbenium ions (6 VE species) do with alkenes, generating cyclopropanes and protonated cyclopropanes, respectively.
Catalysts 12 00758 sch006
Scheme 7. β-insertions and β-eliminations in transition metal complexes are modeled by corner-to-corner migration in protonated cyclopropanes.
Scheme 7. β-insertions and β-eliminations in transition metal complexes are modeled by corner-to-corner migration in protonated cyclopropanes.
Catalysts 12 00758 sch007
Scheme 8. β-Eliminations of transition metal complexes generate unsaturated species that are stabilized by coordination to an adequate ligand or solvent molecule.
Scheme 8. β-Eliminations of transition metal complexes generate unsaturated species that are stabilized by coordination to an adequate ligand or solvent molecule.
Catalysts 12 00758 sch008
Scheme 9. Ethyl cations and πH-ethenium ions (π-complex of a proton with ethylene) have similar stabilities.
Scheme 9. Ethyl cations and πH-ethenium ions (π-complex of a proton with ethylene) have similar stabilities.
Catalysts 12 00758 sch009
Scheme 10. Both the hydrogen- (Equilibrium (34)) and carbon-atom scrambling (Equilibrium (35)) are observed in isopropyl salt in a super-ionizing media (SbF5/SO2ClF).
Scheme 10. Both the hydrogen- (Equilibrium (34)) and carbon-atom scrambling (Equilibrium (35)) are observed in isopropyl salt in a super-ionizing media (SbF5/SO2ClF).
Catalysts 12 00758 sch010
Scheme 11. Edge-protonated cyclopropane is slightly less stable than corner-protonated cyclopropane.
Scheme 11. Edge-protonated cyclopropane is slightly less stable than corner-protonated cyclopropane.
Catalysts 12 00758 sch011
Scheme 12. Hydrogen- and carbon-atom scrambling in butyl salts in a super-ionizing medium (SbF5/SO2ClF).
Scheme 12. Hydrogen- and carbon-atom scrambling in butyl salts in a super-ionizing medium (SbF5/SO2ClF).
Catalysts 12 00758 sch012
Scheme 13. Protonated cyclobutane is much less stable than protonated methylcyclopropane.
Scheme 13. Protonated cyclobutane is much less stable than protonated methylcyclopropane.
Catalysts 12 00758 sch013
Scheme 14. Hydrogen- and carbon-atom scrambling in 2-methylbut-2-yl (tertioamyl) salt in a super-ionizing medium (SbF5/SO2ClF).
Scheme 14. Hydrogen- and carbon-atom scrambling in 2-methylbut-2-yl (tertioamyl) salt in a super-ionizing medium (SbF5/SO2ClF).
Catalysts 12 00758 sch014
Scheme 15. The 2-norbornyl cation is a stable corner-protonated cyclopropane derivative (non-classical carbonium ion).
Scheme 15. The 2-norbornyl cation is a stable corner-protonated cyclopropane derivative (non-classical carbonium ion).
Catalysts 12 00758 sch015
Scheme 16. The cyclobutylmethyl cation is not a stable primary carbenium ion.
Scheme 16. The cyclobutylmethyl cation is not a stable primary carbenium ion.
Catalysts 12 00758 sch016
Scheme 17. In the gas phase, the cyclopentyl cation equilibrates with the corner-protonated spiropentane intermediate, which permits its carbon-atom scrambling and fragmentation into ethylene and an allyl cation.
Scheme 17. In the gas phase, the cyclopentyl cation equilibrates with the corner-protonated spiropentane intermediate, which permits its carbon-atom scrambling and fragmentation into ethylene and an allyl cation.
Catalysts 12 00758 sch017
Table 1. Fundamental reactions of transition metal complexes and comparison with organic reactions. Tolman’s rules are applied to define (conventionally) the oxidation number x of the metal, M (VE: number of valence electrons). [M] = M(x) for a metal bearing other ligands eventually, including solvent molecules. Reprinted with permission from Ref. [17]. Copyright 2019 Wiley.
Table 1. Fundamental reactions of transition metal complexes and comparison with organic reactions. Tolman’s rules are applied to define (conventionally) the oxidation number x of the metal, M (VE: number of valence electrons). [M] = M(x) for a metal bearing other ligands eventually, including solvent molecules. Reprinted with permission from Ref. [17]. Copyright 2019 Wiley.
1. Ligand exchange
Catalysts 12 00758 i001
No change of valence electron count and oxidation number between the metal center of reactant [M]=CO and product [M](π-alkene) by convention.
2. Oxidative addition/reductive elimination
Catalysts 12 00758 i002
Oxidative addition requires an unsaturated metallic species (with 16, 14 VE). The metal oxidation number is increased by two and the metal of the adduct has two more VE than the metal in the starting complex. Note that unsaturated [M] can add to any kinds of X-Y bonds reversibly. The addition of carbenes R2C: (C with a sextet) to X-H bonds generates stable compounds R2C(X)-H with an octet carbon center. The opposite reaction is an α-elimination or 1,1-elimination.
3. α-Insertion/α-elimination
Catalysts 12 00758 i003
The number of VE is reduced by two when going from the reactant to the product in an α-insertion; the oxidation number of the metal is not changed (Z=O, NR’, C(X)Y; R=H, alkyl, alkenyl, aryl, R’O, Hal, etc.).
4. β-Insertion/β-elimination
Catalysts 12 00758 i004
The number of VE is reduced by two when going from reactant to product in the β-insertion; the oxidation number of the metal is not changed (R = H, alkyl, alkenyl, aryl, R’O, Hal, etc.).
5. α-cycloinsertion/α-cycloelimination
Catalysts 12 00758 i005
The number of VE of the metal is reduced by two when going from reactant to product in the α-cycloinsertion; the oxidation number of the metal is increased by two by convention.
6. Oxidative cyclization/reductive fragmentation (β-cycloinsertion/β-cycloelimination)
Catalysts 12 00758 i006
The number of VE of the metal is reduced by two when going from reactant to product in the oxidative cyclization and the oxidation number of the metal increases by two by convention. The metal-catalyzed cycloinsertions can combine other unsaturated reactants besides alkenes.
7. Reactions at the coordinated ligands
Catalysts 12 00758 i007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vogel, P.; Houk, K.N. Organic Chemistry and Synthesis Rely More and More upon Catalysts. Catalysts 2022, 12, 758. https://doi.org/10.3390/catal12070758

AMA Style

Vogel P, Houk KN. Organic Chemistry and Synthesis Rely More and More upon Catalysts. Catalysts. 2022; 12(7):758. https://doi.org/10.3390/catal12070758

Chicago/Turabian Style

Vogel, Pierre, and Kendall N. Houk. 2022. "Organic Chemistry and Synthesis Rely More and More upon Catalysts" Catalysts 12, no. 7: 758. https://doi.org/10.3390/catal12070758

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop