Next Article in Journal
MicroRNA miR-491-5p Targeting both TP53 and Bcl-XL Induces Cell Apoptosis in SW1990 Pancreatic Cancer Cells through Mitochondria Mediated Pathway
Previous Article in Journal
One-Pot Ugi/Aza-Michael Synthesis of Highly Substituted 2,5-Diketopiperazines with Anti-Proliferative Properties
Previous Article in Special Issue
Palladium-Catalyzed Multicomponent Synthesis of 2-Imidazolines from Imines and Acid Chlorides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

New Advances in Titanium-Mediated Free Radical Reactions

1
Department of Chemistry, Materials, and Chemical Engineering "Giulio Natta", Politecnico di Milano, P.zza Leonardo da Vinci 32, 20133 Milano, Italy
2
INSTM (National Consortium for Materials Science and Technology) Local Unit, Politecnico di Milano, 20133 Milan, Italy
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(12), 14700-14732; https://doi.org/10.3390/molecules171214700
Submission received: 31 October 2012 / Revised: 28 November 2012 / Accepted: 6 December 2012 / Published: 11 December 2012
(This article belongs to the Special Issue Transition Metals Catalysis)

Abstract

:
Titanium complexes have been widely used as catalysts for C-C bond-forming processes via free-radical routes. Herein we provide an overview of some of the most significant contributions in the field, that covers the last decade, emphasizing the key role played by titanium salts in the promotion of selective reactions aimed at the synthesis of multifunctional organic compounds, including nucleophilic radical additions to imines, pinacol and coupling reactions, ring opening of epoxides and living polymerization.

1. Introduction

Since the pioneering work of Barton et al. [1] of 1960, the potential of radical reactions has been extensively explored, thanks also to the growth of advanced techniques for the detection of radicals and the definition of their structure and reactivity. This ongoing research has allowed the development of innovative selective free-radical reactions, occurring under mild conditions and tolerated by common functional groups, making such processes suitable for the synthesis and functionalization of organic compounds [2,3]. In the last decade, important improvements in this filed have been achieved by combining radical chemistry with transition-metal catalysis, due to the key role of transition metal complexes both in promoting and highly controlling sophisticated free-radical synthetic routes [4,5,6,7]. This approach has been adopted for different applications, ranging from oxidative [8,9,10] and reductive [11] radical processes to the selective C-C and C-N bond formation [12,13,14,15,16,17,18,19], passing through living polymerization [20,21,22].
Among the transition metals widely employed for these purposes, in the present review we focus on titanium and address the recent advances in selective free-radical processes mediated by Ti(III) and Ti(IV) complexes. Titanium, the seventh most abundant metal on Earth, is one of the cheapest transition metals and a lot of titanium compounds are nontoxic and environmentally friendly. The high added value of using titanium salts to promote radical synthetic routes not only relies on their selective intervention in the initiation and the termination steps of the radical chain, but also on their remarkable coordinating role shown, due to their Lewis acid character.
Limiting our overview to the results reported in the literature in the past twelve years, Section 2 describes the potential of Ti(III)/PhN2+ and Ti(III)/hydroperoxide systems in promoting one-pot, multicomponent, nucleophilic radical additions to aldimines and ketimines generated in situ. Section 3 deals with the most significant examples achieved in pinacol and coupling reactions while, in Section 4, ample discussion is devoted to the ring-opening of epoxides mediated and catalyzed by titanocenes. Finally, in Section 5, we briefly disclose the recent results reported in the field of living polymerization.
It is noteworthy that titanium dioxide is also able to promote free-radical processes by photoactivation at suitable wavelengths, but this chemistry is not the object of the present review. The combined action of light and TiO2, under oxidative conditions, finds interesting applications in the field of pollutant photodegradation, selective organic photosynthesis and antibacterial activity. For a comprehensive overview on the TiO2 photocatalytic activity, we refer the reader to some very recent reviews [23,24,25,26].

2. Nucleophilic Radical Addition to Imines

Nucleophilic radical addition to imines, iminium salts or hydrazones, mediated by transition metal salts such as manganese [27], copper [28,29,30], and zirconium [31,32,33], represents a valuable alternative to the classical ionic routes for the selective C-C bond formation and the synthesis of high added value polifunctional molecules [12,13,14,15,16,17]. This approach becomes even more intriguing if the transition metal is able to promote, in one-pot, the in situ formation of the imine followed by nucleophilic addition, in accordance with the multicomponent reaction (MCR) strategy. In the last decade TiCl3, because of this particular specificity, has attracted the increasing interest of our research group, enabling us to develop quite a number of free-radical MCRs.

2.1. TiCl3/ArN2+ System

The selective arylative amination of aromatic aldehydes using the TiCl3/ArN2+ system (Scheme 1) was first reported by Clerici and Porta in 1990 [34]. According to this procedure, Ti(III) promotes the aryldiazonium salt decomposition to an aryl radical. Simultaneously, the Ti(IV) formed, due to its oxophilicity, favors the in situ formation of aldimines from anilines and aldehydes. Furthermore, as a strong Lewis acid, Ti(IV) enhances, via N-complexation, the electrophilicity of the C=N bond carbon atom towards nucleophilic attack by the aryl radical.
This reaction occurs in an aqueous environment, where it is well known that imines, easily hydrolysable, are quite unstable and less prone to a nucleophilic attack. Later, in 2005, this approach was successfully extended to a wider range of nucleophilic radical sources by introducing a fourth component in the cascade reaction. In fact, in the presence of alkyl iodides [35], the phenyl radical, deriving from the corresponding diazonium salt decomposition, promotes the fast iodide-atom abstraction, generating an alkyl radical which, in turn, undergoes a nucleophilic attack onto aldimines generated in situ (Scheme 2).
According to a similar procedure, the nucleophilic addition of ethers onto aldimines was performed via α-H atom abstraction via the same phenyl radical source (Scheme 2) [36], providing the first radical version of the Mannich reaction. The resulting iminium radical is then reduced by a second equivalent of TiCl3, rendering the overall mechanism irreversible towards the formation of the desired product.

2.2. TiCl3/ROOH and TiCl4-Zn/ROOH Systems

In 2006, we reported a series of new one-pot multicomponent reactions mediated by TiCl3 or TiCl4-Zn hydroperoxide systems, which overcome some limits of the previously disclosed approach. The choice of the right hydroperoxide allows the formation of more practical, efficient and selective precursors of the radical cascade. As a consequence, the new procedure can be extended to a wider range of nucleophilic sources and can be applied even onto ketimines generated in situ, allowing us to create different building blocks for bioactive molecules (Scheme 3). Moreover, this reaction occurs under mild conditions in 30–60 min. Finally, the proposed approach requires neither the pre-formation of the imine nor the protection of the amino group and may easily be conducted under aqueous conditions.

2.2.1. TiCl3/ROOH

An alternative route to the radical aminoalkylation of ethers was proposed by Porta and co-workers in 2006 [37]. The TiCl3-tert-butylhydroperoxide (t-BuOOH) system was revealed to be a more practical, efficient and selective precursor of α-alkoxyalkyl radicals from ethers. Accordingly, TiCl3/t-BuOOH readily assemble an aniline, an aldehyde, and an ether, leading to 1,2-aminoethers (Scheme 4).
In agreement with the general proposed mechanism (Scheme 5), the dropwise addition of TiCl3 into the reaction medium promotes the decomposition of the hydroperoxide, leading to the formation of tert-butoxyl radical (t-BuO·). The latter undergoes an α-H abstraction from the ether, with higher selectivity with respect to the phenyl radical generated by the TiCl3/PhN2+ system (no β-hydrogen abstraction was observed). Once formed, the nucleophilic ketyl radical adds to the aldimine generated in situ, in analogy with the mechanism indicated in Scheme 2.
The reaction goes like a titration and can be easily monitored by considering the different color of the species involved: Ti(III) (violet) and Ti(IV) (orange). Thus, the addition of the TiCl3 solution stops when a pale blue color is barely maintained to ensure the complete decomposition of the peroxide. This methodology was then extended to the selective carbamoylation of aldimines in formamide [38], providing a radical version of the Strecker synthesis of α-amino amides (Scheme 6a), and to the aminoalkylation of methanol [39], leading to the formation of α,β-amino alcohols (Scheme 6b).
When the same protocol is used in the presence of an alcoholic solvent different from methanol, a domino reaction competes with the one-pot process (Scheme 7) [40].
In fact, in this case the ketyl radical formed behaves as a strong reducing agent [41] and competes with Ti(III) in the reduction of t-BuOOH, leading to the formation of the corresponding aldehyde. The latter, when in the presence of an aniline, promotes the formation of an aldimine, which is prone to nucleophilic attack by a second ketyl radical. As a consequence, when the reaction is carried out in the absence of aldehydes, it proceeds to the formation of β-amino alcohols by in situ generation of an aldehyde.

2.2.2. TiCl4-Zn/ROOH

Since ketimines are less stable in aqueous medium and less reactive than aldimines, studies involving the reductive radical addition to ketimines are scant in comparison with those conducted on their aldimine counterparts. Moreover, in almost all cases the preformation of ketimines is required. It is because of this low stability under non-anhydrous conditions that the TiCl3/ROOH system resulted ineffective when an aniline, a ketone and formamide were assembled in one-pot with the intention of promoting the formation of quaternary α-amino amides. Nevertheless, we recently reported an optimization of the previously discussed protocol, consisting in the replacement of the aqueous TiCl3 with an anhydrous solution of TiCl4 in CH2Cl2 in combination with Zn powder (Scheme 8a) [42]. The same protocol was successfully applied in methanol solvent for the selective hydroxymetylation of ketimines, affording the corresponding β,β-disubstituted-β-amino alcohols (Scheme 8b) [43].
The mechanism is analogous to that previously discussed and Zn metal has the unique role of sacrificial reductant for the in situ generation of Ti(III) from Ti(IV) (Scheme 9).
Surprisingly, the TiCl4-Zn/t-BuOOH combination was also revealed to be even more effective and selective with respect to the TiCl3/t-BuOOH approach in promoting domino reactions. 1,2-Amino alcohols were obtained in high yields by simply combining two molecules of ethanol or propanol in the presence of aromatic and aliphatic amines, without requiring the addition of aldehydes or ketones (Scheme 10) [44]. Even more interesting is that an analogous reaction occurred by operating in cyclic ether solvents [44].
According to the proposed mechanism (Scheme 11), it was possible to synthetize amino alcohols by electrophilic amination of ketyl radicals, promoting the formation of the imine via ring opening of the ether cyclic and the subsequent nucleophilic radical addition by a second ketyl radical.

2.3. Cp2TiCl2/Zn Systems

In 2011, Saha and Roy reported a selective radical-induced allylation of pre-formed aldimines for the synthesis of homoallyl amines [45]. The combination of titanocene(IV) dichloride and Zn dust in THF under argon leads to the formation of Cp2TiCl in situ. The latter acts as radical initiator, promoting the formation of the allyl radical from the corresponding allyl bromide (Scheme 12). This approach was successfully applied to the synthesis of deoxy aza-disaccharides and for the preparation of the bicyclic skeleton of alkaloids.

3. Coupling Reactions

Radical coupling mediated by transition metal salts represents one of the pillars of free-radical synthesis. Once again, simple titanium salts and titanium complexes bearing specific organic ligands result particularly effective in promoting intramolecular as well as intermolecular coupling with high stereo-, diastero- and even enantioselectivity [46,47]. Among the several protocols developed, pinacolization, allylation and cross-coupling reactions have attracted particular attention in the last decade.

3.1. Pinacol Reactions

The pinacolization of aldehydes is one of the most studied reactions in the area of radical coupling for C-C bond formation. This interest arises from the usefulness of hydrobenzoins as chiral auxiliaries and ligands for stereoselective organic synthesis. Moreover, the pinacol reaction represents an ideal tool for studying the enatioselective catalyzed formation of radicals [48].
This reaction can be accomplished with a variety of one- or two-electron metal reducing agents. Our group has widely investigated the role that TiCl3 and TiCl4 play in promoting pinacolization of aromatic aldehydes. In particular, a dl-stereodirecting pinacolization was accomplished by using TiCl3 in a strictly anhydrous non-coordinating solvent (CH2Cl2), favored by Ti(III)-complexation of the oxygen atom of the carbonyl group, which lowers the reduction potential of the aromatic aldehyde (Scheme 13) [49].
Later on, in the same CH2Cl2 solvent, an analogous reactivity was observed when TiCl4 is coupled with diisopropylethylamine (DIPEA). Porta and co-workers [50] revealed how, in this case, the TiCl4/DIPEA complexation is the driving force of the overall process, causing the instantaneous formation of Ti(III) by an inner-sphere electron transfer from Ti(IV) to DIPEA. Once generated, Ti(III) coordinates the carbonyl oxygen of an added aromatic aldehyde, promoting its reductive dimerization.
More recently we have unexpectedly found that, despite the non-anhydrous conditions, the aqueous TiCl3/t-BuOOH redox system promotes the hydrodimerization of aromatic aldehydes, when used in combination with an alcoholic co-solvent like ethanol [41].
In accordance with the proposed mechanism (Scheme 14), the explanation of the paradox relies in the reducing power of the α-hydroxyalkyl radicals generated in situ via H-atom abstraction from the alcoholic solvent by tert-butoxy radical. These ketyl radicals, instead of Ti(III), are the actual reductants of the aromatic aldehydes.
Pinacolization of aromatic ketones like acetophenone can be performed in good yields and high stereoselectivity (dl/meso ratio = 9/1) in the presence of Cp2TiCl, generated in situ by combining Cp2TiCl2 with an excess of Mn dust [51].
Moreover, new families of Ti(IV) complexes bearing chiral ligands (1, 2 and 3, Figure 1) have been developed in the last decade. These derivatives, when employed in combination with co-reducing agents, revealed to be ideal precursors of chiral Ti(III) catalysts for the enantioselective pinacol coupling of aldehydes.
In 2001 Riant and co-workers [52] reported the combined use of stoichiometric amounts of the titanium-Schiff base complex 1 and Mn to promote the asymmetric dl-diastereoselective pinacolization with high yields and enantioselectivity up to 91% (Scheme 15).
An analogous result was achieved by Joshi et al., who succeeded in performing a catalytic cycle in CH3CN with 10 mol % concentration of complex 2 and stoichiometric amounts of Me3SiCl (TMSCl), followed by the addition of tetrabutylammonium fluoride in THF (Scheme 16) [53].
In 2005, Knoop and Studer reported a new method for the in situ generation of Ti(III) complexes without requiring a co-reducing reagent [54]. Cyclohexadienyl-Ti(IV) compounds, prepared from the corresponding lithiated cyclohexadienes, led to the corresponding Ti(III) complexes upon thermal C-Ti bond homolysis (Scheme 17). Following this procedure, complex 3 was readily prepared and employed to perform pinacol coupling of aromatic aldehydes with diastereomeric and enantiomeric excess.

3.2. Allylation Reactions

Among radical coupling reactions, the Barbier-type allylation of carbonyl derivatives, mediated by different transition metals, has attracted particular attention due to its considerable synthetic relevance (Scheme 18). The generic strategy for this process consists in the transformation of allyl halides into the corresponding allyl radicals which, in turn, would add onto the carbonyl compounds present in the reaction medium.
Very recently, Justicia et al. have shown that an excess of titanocene(III) chloride, generated in situ by the Cp2TiCl2/Mn-dust system, allows to extend this protocol to the crotylation of carbonyl compounds [55].
However, most of the reported protocols usually require stoichiometric amounts of these metals, which is detrimental for the sustainability of the process. Stoichiometric quantities are generally necessary even when operating in the presence of Ti complexes, limiting the applicability of this approach to enatioselective additions [46]. In this context, in 2009 a comprehensive contribution by Oltra and co-workers has shown how catalytic proportions of titanocene(III) complexes, when combined with the regenerating agent 4 (obtained by mixing TMSCl and 2,4,6-collidine), are able to promote the Barbier-type allylation, in accordance with the mechanism reported in (Scheme 19) [56].
This approach has been successfully applied for the allyl radical addition onto aldehydes and ketones (including functionalized carbonyl substrates), for cyclization reactions and for prenylation of aldehdyes (Scheme 20).
The titanocene(III)-mediated Barbier-type prenylation to α,β-unsaturated aldehydes, which allows the introduction of hydroxyl groups at desiderable positions, and the subsequent radical cyclization promoted by the same titanium complex, open the straightforward access to polyhydroxylated terpenoids [57].
The main limitation of the Barbier-type approach is that it requires the use of allylic halides as reactive substrates, while allylic carbonates and carboxylates, which are easily prepared and handled, are inert against titanocene(III) complexes. To overcome this limit, Oltra and co-workers proposed the combination of Cp2TiCl with palladium and nickel complexes [58]. Indeed, it is well known that these metal derivatives are able to activate allylic carbonates and carboxylates by forming η3-allylmetal complexes. The authors have demonstrated how the choice of the metal is crucial in modulating the Cp2TiCl activity in the allylation reaction. In fact, while palladium salts promote the intermolecular coupling with electrophilic reagents (Scheme 21a), the intramolecular allylation of alkenes with allylic carbonates occurs in the presence of nichel complexes, leading to carbocylic derivatives (Scheme 21b).
The titanium/palladium-mediated system has been successfully applied by the same research group for the selective allylation, crotylation and prenylation of aldehydes and ketones [59] (Scheme 22), for the synthesis of homopropargyl alcohols using propargyl carbonates as pronucleophiles [60] (Scheme 23), and to promote the intramolecular Michael-type addition of allylic carboxylates to activated alkenes [61] (Scheme 24). More recently, the efficient allylation of carbonyl compounds starting with allyl carbonates has been performed by replacing Pd with Ni [62].

3.3. Other Coupling Reactions

Several other coupling reactions promoted by titanium complexes have been recently reported. Roy and co-workers have proposed the use of the Cp2TiCl2/Zn-dust system to promote the synthesis of furan derivatives such as (±)-evodone from α-bromo-β-keto enolethers [63]. The same approach has been successfully applied to the asymmetric synthesis of α-methylene bis-γ-butyrolactone [64].
The Cp2TiCl-catalyzed homolytic opening of ozonides, reported by Rosales et al. [65], provides carbon centered radicals suitable to form C-C bonds via both homocoupling and cross-coupling processes (Scheme 25).
Very recently, Streuff and co-workers have reported an efficient enatioselective reductive cyclization of ketonitriles promoted and catalyzed by chiral Ti(III) catalysts (Scheme 26) [66].
An example of the potential of radical chemistry is the use of the Ingold-Fisher “Persistent Radical Effect” (PRE) which has found ample employment in cross-coupling reactions for the selective synthesis of organic compounds. According to PRE, the cross-coupling between two radicals occurs when the two species are generated at similar rates and one is more persistent than the other, that is with a lifetime significantly greater than that for a methyl radical under the same conditions [67].
In 2008 we reported how α,β-dihydroxy ketones can be used as a source of stabilized radicals via selective C-C bond cleavage promoted by TiCl4 [68] the coordinating effect of the metal ion induces an increased stabilization onto the new radicals, driven by the captodative effect (Scheme 27).
This observation prompted us to develop a new one-pot multicomponent route leading to the synthesis of polyfunctional derivatives of high added value. The generation, in the same reaction medium, of transient radicals via thermal decomposition of 2,2′-azo-bis-isobutyronitrile (AIBN) yielded an innovative methodology for the synthesis of β-hydroxynitriles by cross-combination between Ti(IV)-stabilized radicals and the α-cyanoisopropyl radical (Scheme 28).

4. Epoxide Reactions Mediated by Titanocene

Bis (cyclopentadienyl) titanocene III chloride (Cp2TiCl) and substituted titanocenes are well known titanium complexes able to generate regioselective radical epoxide opening through single electron transfer (ET), both in stoichiometric and catalytic amounts [69,70]. The method of radical generation combines both the Lewis acid catalysis and the radical chemistry with interesting advantages: the initial ET by the metal, tuning the acidity of the metal center, avoids SN-type reactions, while the ensuing radical reaction can be controlled by the metal complex and his ligand. Cyclic voltammetry and kinetic measurements have disclosed the structure and mechanism of epoxide opening by Cp2TiCl reagents in solution: the activation of Cp2TiCl2 species by Mn or Zn reducing metals in THF solution leads to a mixture composed by the monomer Cp2TiCl and the chlorine bridged dimer (Cp2TiCl)2, but the more reactive species is the half-open dimer [71,72] (Scheme 29).
Typical radical reactions mediated by titanocene reagents regard selective reduction and deoxygenation of epoxides, C-C bond formation processes such as cyclizations, and intermolecular additions to activated olefins [73,74].
It is worth noting that, as the reaction mechanism generally proceeds via the most substituted (i.e., energetically most stable) radical, the regiochemistry is the opposite (and complementary) to that normally expected for conventional SN2-type epoxide openings with carbon nucleophiles or hydride reagents [73,74].

4.1. Epoxide Deoxygenation

Though epoxides are usually prepared from olefins, their deoxygenation for the generation of C=C double bonds finds ample application in highly oxidized substrates like carbohydrates and other natural derivatives [75,76]. The process proceeds in good to excellent yields, with exceptional chemo-selectivity.
It has been shown that the deoxygenation occurs when operating in the absence of radical acceptors and that it depends on the substitution pattern [77]. On the basis of these observations, Fernández-Mateos and co-workers have recently reported the selective synthesis of terpenoids starting from pinene-oxide derivatives [78].
Doris and co-workers reported in 2002 the Cp2TiCl-mediated deoxigenation of the alkaloid leurosine (from Catharantus roseus), leading to anhydrovinblastine (Scheme 30), the key intermediate in the symthesis of the anticancer drug navelbine [79].
Furthermore, short and straightforward access to analogues of the natural onoceranes mediated by Cp2TiCl2 has been proposed by Barrero and his group [80]. The dimerization of vinyl-epoxides to 1,5-dienes occurs via generation of very stable β-titanoxy allylradicals. The latter undergo dimerization to homo-onoceranes, followed by oxidation (Scheme 31).
In 2009, Aldegunge et al. proposed the synthesis of alk-2-ene-1,4-diols by a new cascade-opening of 1,3-diepoxides, suggesting an alternative route to dihydroxytaxoids [81].
In the same year, Justicia et coworkers [82] showed that β-titanoxy radicals may rearrange to new trisubstituted radicals which undergo a mixed disproportionation process (via hydrogen atom transfer, Scheme 32) leading to allylic alcohols.

4.2. Reductive Epoxide Opening

The catalytic reductive epoxide opening via electron transfer from titanocenes has emerged as an attractive alternative to the classical nucleophilic substitution strategies. The titanium-mediated mechanism combines the well-established advantages of radical chemistry with a regioselectivity of ring opening opposite to that of nucleophilic substitutions, leading to a number of interesting and unusual applications in the synthesis of complex molecules [83,84,85,86,87].

4.2.1. Synthesis of Alcohols

According to mechanistic studies, the depicted β-tytanoxyl radical epoxide complexes constitute the first intermediate of ring opening. These radicals are then reduced by hydrogen atom donor species, like for example 1,4-cyclohexadiene (Scheme 33).
If the use of catalytic amounts of titanocenes is attractive for reagent controlled radical reactions [88], on the other hand this process presents the disadvantage deriving from the generation of stoichiometric amounts of highly toxic reducing waste, such as benzene (Scheme 26). However, it has been demonstrated that titanocene reagents are able to activate water and methanol toward hydrogen atom transfer, by substantially lowering the bond dissociation energy (BDE) of the OH bond [89,90,91]. This combination is chemically more efficient and environmentally benign. Cp2TiCl-mediated reductive epoxide ring opening using water as a hydrogen source, allows access to alcohols with anti-Markovnikov regiochemistry from different epoxides. The use of D2O as a deuterium source, leads to an efficient synthesis of β-deuterated alcohols, which find successful application as internal standards in food analysis [92]. Another suitable hydrogen donor is H2 [93,94].
Titanocene-catalyzed procedure has been immediately extended by Gansäuer et al. to the enantioselective opening of meso-epoxides. Titanocene complexes with chiral ligands, such as those reported in Figure 2, were employed in combination with manganese dust and collidine hydrochloride, the former as a stoichiometric reductant and the latter having the role of regenerating the alcohol and titanocenes [95,96,97,98,99,100,101,102] (Scheme 34).
By operating in the presence of collidinium hydrochloride, the catalytic reaction could be also performed in water. This approach was exploited for the selective reduction of aromatic ketones [103,104] and the selective synthesis of 2-methyl-1,3-diol frameworks, using THF as hydrogen donor [105]. Also α,β-epoxy ketones could be opened with Cp2TiCl, affording β-hydroxy ketones in a better fashion than with SmI2 (Scheme 35) [106].
The ring-opening reaction of tri-substituted epoxides promoted by Cp2TiCl on carvone derivatives has led to exo-methylene allylic alcohols as major compounds [107]. The generation of exocyclic olefins has been reported in the course of radical cyclizations of epoxy alkenes leading to monocyclic terpens as achilleol A [108], oxobicylic drimanes [109], eudesmanolides [110], and different ring synthons of paclitaxel [111]. In these cases, termination of the radical cycle is not reductive, but rather it involves β-hydrogen elimination providing an alkene function.
Very recently Gansäuer et al. have reported a new protocol to perform the radical reduction of epoxides by hydrogen atom transfer catalyzed by the active specie [Cp2TiH] [112]. The active catalyst is generated in situ from [Cp2TiOC2H5)2] in the presence of (CH3)PhSiH2 (Scheme 36).

4.2.2. Intermolecular C-C Bond Formation

Epoxide-derived radicals generated in the presence of titanocenes are also intermediates for intermolecular C–C bond forming reactions [113]. The resultant products, like δ-hydroxyketones, esters, amides, nitriles and δ-lactones, constitute important classes of intermediates in organic synthesis (Scheme 37). It has been reported that the C–C bond formation employing α,β-unsaturated tungsten and chromium carbenes, as radical acceptors of epoxide-derived radicals, occurs with high chemoselectivity. This approach has been applied to both sugar-derived and α,β-unsaturated carbenes [114] (Scheme 38).

4.2.3. Intramolecular C-C Bond Formation

The use of Cp2TiCl (in stoichiometric or catalytic amounts) allowed to promote the synthesis of natural products by the construction of three- to six- and eight-membered carbocycles in good to excellent yields [115,116,117,118,119,120]. Small ring 3-exo and 4-exo cyclizations were achieved by nucleophilic intramolecular radical addition to aldehydes and ketones [117] and natural products, like (E)-endo-bergamoten-12-oic acids [115] or carbacephams, are significant examples of synthetic applications (Scheme 39) [121,122,123,124,125].
While gem-dialkyl or gem-dialkoxyl substitution is usually required to maintain an efficient propagation of the radical chain, very recently Gansäuer et al. have reported the first example of 4-exo cyclization on unsaturated epoxides without gem disubstitution [124]. This result was achieved by using cationic functionalized titanocenes as template catalysts (Scheme 40). According to the proposed mechanism [126], the two-point binding which occurs between the substituted radical and the template forces the radical center and the radical acceptor into close spatial proximity (Figure 3).
5-exo cyclizations are usually conducted by using alkenes and alkynes as radical traps and this approach has been widely applied for several synthetic routes [127,128,129,130,131,132,133,134,135,136,137]. Following an analogous procedure, a 6-endo cyclization employing Cp2TiCl was reported in 2001 by Takahaschi and co-workers for the synthesis of (±)-smenospondiol [138]. Titanocene(III) chloride mediated 5-exo and 6-exo cyclization has been recently applied for the regio- and diastereoselective synthesis of highly functionalized terpenic cyclopentanes [139], and for the synthesis of (−)-methylenolactocin and (−)-protolichesterinic acid [140]. Titanocene-promoted cyclization of unsaturated epoxylactones was also reported as a key step for diastereoselective synthesis of limonoid CDE fragments [141].
Epoxypolyene cyclizations were performed in the presence of Cp2TiCl2 using the collidine-chlorotrimethylsilane system, leading to exo-cyclic olefins by hydrogen atom abstraction (Scheme 41) [142,143,144]. These cyclization reactions proceed with high stereoselectivity giving only one of the many possible stereoisomers. Moreover, some of the cyclization products can be readily elaborated into natural products or key intermediates such as in the preparation of zonarol, a natural antifungal (Scheme 42).
Following the same procedure, 7-endo cyclizations can be performed in surprisingly high yields (Scheme 43) [145]. Finally, 8-endo cyclizations have been reported by Roy and his group for the preparation of aromatic ethers [146].

5. Living Polymerizations

Several metal complexes were designed and tested to effectively control the radical polymerization of a number of polar olefins, leading to the synthesis of complex macromolecular architectures [147,148].
The use of titanium complexes for this purpose is relatively rare. Nevertheless, in recent years it has been emphasized the Ti(III)/Ti(IV) capacity of modulating the equilibrium in atom transfer reactions. As a consequence, this redox pair was reported as a very active catalytic system for atom transfer radical polymerization [149].
Radical controlled polymerization reactions by epoxide radical ring opening (RRO) have been performed in the presence of Cp2TiCl2 in combination with Zn powder [150,151,152]. According to this approach, methyl methacrylate (MMA) polymerization has been performed with high-molecular weight in aqueous conditions (Mw = 51,000–73,000). An analogous route has been recently reported for the radical polymerization of styrene [151]. The Ti(III)Cp2Cl-catalyzed RRO of epoxides produces Ti-alkoxides, which initiate the ring opening polymerization of cyclic esters like ε-caprolactone (Scheme 44) [153], the radical polymerization of isoprene [154] and styrene/isoprene copolymers [155].
The ligand effect in Ti-mediated living radical styrene polymerizations initiated by epoxide RRO has been widely investigated by Asandei and Moran, focusing in particular onto alkoxide and bisketonate titanium complexes [156], scorpionate and half-sandwich LTiCl3 derivatives [157] and substituted sandwich metallocenes [158]. Moreover, the same research group disclosed in detail the role of solvents and additives for this polymerization reactions [159], the effect of the reducing agent and temperature, as well as of the ratio titanium/epoxide and titanium/zinc [160].
Excellent results in the living radical polymerization were also achieved by initiating the process via single electron transfer reduction of aldehydes promoted by Cp2TiCl [161]. This approach has been successfully applied also for the polymerization of ε-caprolactone [162].
Very recently, Cp2TiCl2 has been applied in the field of photopolymerization for applications requiring safe conditions (Scheme 45) [163]. Cp2TiCl2/silane (tris(trimethylsilyl)silane TTMSS)/ iodonium salt (Ph2I+) resulted to be a good photoinitiating system in the free radical promoted cationic photopolymerization process of an epoxide monomer (EPOX) upon long wavelength excitation (λ > 400 nm). Under visible LED bulbs, xenon lamp and laser diodes, the reactions exhibit an excellent performance and a remarkable final polymerization close to 100%. The basic idea is to produce a radical (from a photoinitiating system) which, in turn, should be oxidized by the iodonium salt: Cp2TiCl2 generates under argon cyclopentadienyl radicals (Cp) and/or metal-centered radicals CpTiCl2. In the presence of TTMSS, Cp can rapidly react by hydrogen abstraction with the silane Si-H generating a silyl radical. The latter can be easily oxidized by Ph2I+ with a rate constant of 2.6 × 106 M−1 s−1, leading to the formation of a silylium cation, which, in turn, can initiate the cationic polymerization by ring-opening of EPOX.

6. Conclusions

This study covers the literature over the past twelve years emphasizing the key role that titanium salts and complexes play in promoting synthetic free-radical processes. While several excellent reviews, focusing on specific Ti-mediated reactions and defined product targets or on the general role of transition metals in promoting radical chains, have been recently published, to the best of our knowledge this work represents the first comprehensive contribution which outlines the potential of titanium catalysis in radical reactions, analytically overviewing the recent progresses in selective free-radical organic synthesis.
Though multicomponent procedures are of great interest for the development of new synthetic routes, examples in the field of radical reactions are in general scant. Among the few protocols recently reported, the nucleophilic radical addition to imines mediated by Ti(III)/Ti(IV) systems represents an innovative one-pot approach leading to a wide range of intriguing functionalized derivatives under very mild experimental conditions.
Pinacolization of aldehydes is particularly attractive, not only for the importance of the final diols, but also because it represents a significant model for investigating enantioselectivity in radical chemistry. Other radical coupling reactions, when mediated by titanium complexes, lead to high added value products with chemo- and stereo-selectivity.
Finally, deoxygenation or reductive ring opening of epoxides by means of titanocenes find ample use in the synthesis of natural and biologically active molecules. Ti-mediated radical ring opening of epoxides is also successfully applied to promote living radical polymerization of olefins (such as styrene and isoprene) and lactones (for example ε-caprolactone).
The development and optimization of titanium ligands, including the introduction of chiral complexes, allows in many cases to perform stero- and/or enantio-selective syntheses of the desired products. The ongoing research for the design, synthesis, and application of these new ligands guarantees the development of new Ti-based catalytic systems for selective transformations under very mild conditions.

Acknowledgments

We thank MIUR for continual support of our free-radical (PRIN 2008, project 2008KRBLP5_003) and organic (FIRB-Future in Research 2008, project RBFR08XH0H_001) chemistry.

References

  1. Barton, D.H.; Beaton, J.M.; Geller, L.E.; Pechet, M.M. A new photochemical reaction. J. Am. Chem. Soc. 1960, 82, 2640–2641. [Google Scholar] [CrossRef]
  2. Renaud, P.; Sibi, M.P. (Eds.) Radicals in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 2001; Volumes 1 and 2. [Google Scholar]
  3. Chatgilialoglu, C.; Studer, A. (Eds.) Encyclopedia of Radicals in Chemistry, Biology and Materials; Wiley: Chichester, UK, 2012; Volumes 1 and 2. [Google Scholar]
  4. Jahn, U. Radicals in transition metal catalyzed reactions? Transition metal catalyzed radical reactions? A fruitful interplay anyway: Part 1. Radical catalysis by group 4 to group 7 elements. Top. Curr. Chem. 2012, 320, 121–189. [Google Scholar] [CrossRef] [PubMed]
  5. Jahn, U. Radicals in transition metal catalyzed reactions? Transition metal catalyzed radical reactions? A fruitful interplay anyway: Part 2. Radical catalysis by group 8 and 9 elements. Top. Curr. Chem. 2012, 320, 191–322. [Google Scholar] [CrossRef] [PubMed]
  6. Jahn, U. Radicals in transition metal catalyzed reactions? Transition metal catalyzed radical reactions? A fruitful interplay anyway: Part 3. Radical catalysis by group 10 and 11 elements and bimetallic catalysis. Top. Curr. Chem. 2012, 320, 323–451. [Google Scholar] [CrossRef] [PubMed]
  7. Yorimitsu, H. Transition metals and Radicals. In Encyclopedia of Radicals in Chemistry, Biology and Materials; Chatgilialoglu, C., Studer, A., Eds.; Wiley: Chichester, UK, 2012; Volume 2, pp. 1003–1018. [Google Scholar]
  8. Burton, J. Manganese (III) acetate, CAN, and Fe(III) salts in oxidative radical chemistry. In Encyclopedia of Radicals in Chemistry, Biology and Materials; Chatgilialoglu, C., Studer, A., Eds.; Wiley: Chichester, UK, 2012; Volume 2, pp. 901–941. [Google Scholar]
  9. Recupero, F.; Punta, C. Free Radical functionalization of organic compounds catalyzed by N-hydroxyphthalimide. Chem. Rev. 2007, 107, 3800–3842. [Google Scholar] [CrossRef] [PubMed]
  10. Minisci, F.; Recupero, F.; Cecchetto, A.; Gambarotti, C.; Punta, C.; Faletti, R.; Paganelli, R.; Pedulli, G.F. Mechanisms of the aerobic oxidation of alcohols to aldehydes and ketones, catalysed under mild conditions by persistent and non-persistent nitroxyl radicals and transition metal salts - Polar, enthalpic, and captodative effects. Eur. J. Org. Chem. 2004, 109–119. [Google Scholar] [CrossRef]
  11. Coote, S.C.; Flowers, R.A., II; Skrydstrup, T.; Procter, D.J. Organic synthesis using samarium diiodide. In Encyclopedia of Radicals in Chemistry, Biology and Materials; Chatgilialoglu, C., Studer, A., Eds.; Wiley: Chichester, UK, 2012; Volume 2, pp. 849–900. [Google Scholar]
  12. Friestad, G.K. Asymmetric radical addition to chiral hydrazones. In Topics in Current Chemistry: Radicals in Synthesis III; Gansauer, A., Heinrich, M., Eds.; Springer-Verlag: Berlin, Germany, 2012; Volume 320, pp. 61–92. [Google Scholar]
  13. Friestad, G.K. Organomanganese-mediated radical reactions. In The Chemistry of Organo-Manganese Compounds; Marek, I., Rappoport, Z., Eds.; Wiley: Chichester, UK, 2011; pp. 559–584. [Google Scholar]
  14. Yamada, K.; Tomioka, K. Copper-catalyzed asymmetric alkylation of imines with dialkylzinc and related reactions. Chem. Rev. 2008, 108, 2874–2886. [Google Scholar] [CrossRef] [PubMed]
  15. Akindele, T.; Yamada, K.; Tomioka, K. Dimethylzinc-initiated radical reactions. Acc. Chem. Res. 2009, 42, 345–355. [Google Scholar] [CrossRef] [PubMed]
  16. Pastori, N.; Gambarotti, C.; Punta, C. Recent developments in nucleophilic radical addition to imines: The key role of transition metals and the new Porta radical-type version of the Mannich and Strecker Reactions. Mini-Rev. Org. Chem. 2009, 6, 184–195. [Google Scholar] [CrossRef]
  17. Miyabe, H.; Yoshioka, E.; Kohtani, S. Progress in intermolecular carbon radical addition to imine derivatives. Curr. Org. Chem. 2010, 14, 1254–1264. [Google Scholar] [CrossRef]
  18. Rowlands, G.J. Radicals in organic synthesis. Part 1. Tetrahedron 2009, 65, 8603–8655. [Google Scholar] [CrossRef]
  19. Rowlands, G.J. Radicals in organic synthesis. Part 2. Tetrahedron 2010, 66, 1593–1636. [Google Scholar] [CrossRef]
  20. Kamigaito, M.; Ando, T.; Sawamoto, M. Metal-catalyzed living radical polymerization. Chem. Rev. 2001, 101, 3689–3745. [Google Scholar] [CrossRef] [PubMed]
  21. Ouchi, M.; Terashima, T.; Sawamoto, M. Transition metal-catalyzed living radical polymerization: toward perfection in catalysis and precision polymer synthesis. Chem. Rev. 2009, 109, 4963–5050. [Google Scholar] [CrossRef] [PubMed]
  22. Pintauer, T.; Matyjaszewski, K. Atom transfer radical addition and polymerization reactions catalyzed by ppm amounts of copper complexes. Chem. Soc. Rev. 2008, 37, 1087–1097. [Google Scholar] [CrossRef] [PubMed]
  23. Gaya, U.I.; Abdullah, A.H. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. J. Photochem. Photobiol. C. 2008, 9, 1–12. [Google Scholar] [CrossRef]
  24. Gambarotti, C.; Punta, C.; Recupero, F.; Caronna, T.; Palmisano, L. TiO2 in organic photosynthesis: Sunlight induced functionalization of heterocyclic bases. Curr. Org. Chem. 2010, 14, 1153–1169. [Google Scholar] [CrossRef]
  25. Fagnoni, M.; Dondi, D.; Ravelli, D.; Albini, A. Photocatalysis for the formation of the C-C bond. Chem. Rev. 2007, 107, 2725–2756. [Google Scholar] [CrossRef] [PubMed]
  26. Visai, L.; De Nardo, L.; Punta, C.; Melone, L.; Cigada, A.; Imbriani, M.; Arciola, C.R. Titanium oxide antibacterial surfaces in biomedical devices. Int. J. Artif. Org. 2011, 34, 929–946. [Google Scholar] [CrossRef] [PubMed]
  27. Friestad, G.K.; Ji, A.; Baltrusaitis, J.; Korapala, C.S.; Qin, J. Scope of stereoselective Mn-mediated radical addition to chiral hydrazones and application in a formal synthesis of quinine. J. Org. Chem. 2012, 77, 3159–3180. [Google Scholar] [CrossRef] [PubMed]
  28. Fujihara, H.; Nagai, K.; Tomioka, K. Copper-amidophosphine catalyst in asymmetric addition of organozinc to Imines. J. Am. Chem. Soc. 2000, 122, 12055–12056. [Google Scholar] [CrossRef]
  29. Nagai, K.; Fujihara, H.; Kuriyama, M.; Yamada, K.; Tomioka, K. Efficient chiral amidophosphine ligand for copper-catalyzed asymmetric addition of diethylzinc to N-sulfonylimines. Chem. Lett. 2002, 31, 8–9. [Google Scholar] [CrossRef]
  30. Soeta, T.; Nagai, K.; Fujihara, H.; Kuriyama, M.; Tomioka, K. Asymmetric alkylation of N-toluenesulfonylimines with dialkylzinc reagents catalyzed by copper-chiral amidophosphine. J. Org. Chem. 2003, 68, 9723–9727. [Google Scholar] [CrossRef] [PubMed]
  31. Porter, J.R.; Traverse, J.F.; Hoveyda, A.H.; Snapper, M.L. Enantioselective synthesis of arylamines through Zr-catalyzed addition of dialkylzincs to imines. Reaction development by screening of parallel libraries. J. Am. Chem. Soc. 2001, 123, 984–985. [Google Scholar] [CrossRef] [PubMed]
  32. Porter, J.R.; Traverse, J.F.; Hoveyda, A.H.; Snapper, M.L. Three-component catalytic asymmetric synthesis of aliphatic amines. J. Am. Chem. Soc. 2001, 123, 10409–10410. [Google Scholar] [CrossRef] [PubMed]
  33. Akullian, L.C.; Snapper, M.L.; Hoveyda, A.H. Three-component enantioselective synthesis of propargylamines through Zr-catalyzed additions of alkyl zinc reagents to alkynylimines. Angew. Chem. Int. Ed. 2003, 42, 4244–4247. [Google Scholar] [CrossRef] [PubMed]
  34. Clerici, A.; Porta, O. Arylative amination of aldehydes promoted by aqueous titanium trichloride. Tetrahedron Lett. 1990, 31, 2069–2072. [Google Scholar] [CrossRef]
  35. Cannella, R.; Clerici, A.; Pastori, N.; Regolini, E.; Porta, O. One-pot four-component reaction: Aqueous TiCl3/PhN2+-mediated alkyl radical addition to imines generated in situ. Org. Lett. 2005, 7, 645–648. [Google Scholar] [CrossRef] [PubMed]
  36. Clerici, A.; Cannella, R.; Panzeri, W.; Pastori, N.; Regolini, E.; Porta, O. TiCl3/PhN2+-mediated radical addition of ethers to aldimines generated in situ under aqueous conditions. Tetrahedron Lett. 2005, 46, 8351–8354. [Google Scholar] [CrossRef]
  37. Clerici, A.; Cannella, R.; Panzeri, W.; Porta, O. A Free radical Mannich type reaction: selective α-CH-aminomethylation of ethers by Ti(III)/t-BuOOH system under aqueous acidic conditions. Tetrahedron 2006, 62, 5986–5994. [Google Scholar] [CrossRef]
  38. Cannella, R.; Clerici, A.; Panzeri, W.; Pastori, N.; Punta, C.; Porta, O. Free-radical version of Strecker Synthesis of α-aminomides promoted by aqueous H2O2/TiCl3/HCONH2 System. J. Am. Chem. Soc. 2006, 128, 5358–5359. [Google Scholar] [CrossRef] [PubMed]
  39. Clerici, A.; Ghilardi, A.; Pastori, N.; Punta, C. A new one-pot, four-component synthesis of 1,2-amino alcohols: TiCl3/t-BuOOH-mediated radical hydroxymethylation of imines. Org. Lett. 2008, 10, 5063–5066. [Google Scholar] [CrossRef] [PubMed]
  40. Spaccini, R.; Ghilardi, A.; Pastori, N.; Clerici, A.; Porta, O.; Punta, C. Efficient radical domino approach to β-aminoalcohols from arylamines and alcohols triggered by Ti(III)/t-BuOOH. Tetrahedron 2010, 66, 2044–2052. [Google Scholar] [CrossRef]
  41. Clerici, A.; Greco, C.; Panzeri, W.; Pastori, N.; Punta, C.; Porta, O. Reductive coupling of aromatic aldehydes promoted by an aqueous TiCl3/t-BuOOH system in alcoholic cosolvents. Eur. J. Org. Chem. 2007, 4050–4055. [Google Scholar] [CrossRef]
  42. Pastori, N.; Greco, C.; Clerici, A.; Punta, C.; Porta, O. Free-Radical addition to ketimines generated in situ. One-pot synthesis of quaternary α-aminomides promoted by a H2O2/TiCl4-Zn/HCONH2 system. Org. Lett. 2010, 12, 3898–3901. [Google Scholar] [CrossRef] [PubMed]
  43. Rossi, B.; Pastori, N.; Clerici, A.; Punta, C. Free-radical hydroxymethylation of ketimines generated in situ: A one-pot multicomponent synthesis of β,β-disubstituted-β-aminoalcohols. Tetrahedron 2012, 68, 10151–10156. [Google Scholar] [CrossRef]
  44. Prosperini, S.; Pastori, N.; Ghilardi, A.; Clerici, A.; Punta, C. New domino radical synthesis of aminoalcohols promoted by TiCl4–Zn/t-BuOOH system: selective hydroxyalkylation of amines in alcohol or in cyclic ether cosolvents. Org. Biomol. Chem. 2011, 9, 3759–3767. [Google Scholar] [CrossRef] [PubMed]
  45. Saha, S.; Roy, S.C. Titanocene(III) chloride mediated radical induced allylation of aldimines: Formal synthesis of C-linked 40-deoxy aza-disaccharide. J. Org. Chem. 2011, 76, 7229–7234. [Google Scholar] [CrossRef] [PubMed]
  46. Ramόn, D.J.; Yus, M. In the arena of enantioselective synthesis, titanium complexes wear the laurel wreath. Chem. Rev. 2006, 106, 2126–2208. [Google Scholar] [CrossRef] [PubMed]
  47. Szymoniak, J.; Mose, C. Synthesis and reactivity of allyltitanium derivatives. In Titanium and Zirconium in Organic Synthesis; Marek, I., Ed.; Wiley-VCH: Weinheim, Germany, 2002; pp. 451–474. [Google Scholar]
  48. Chatterjee, A.; Joshi, N.N. Evolution of the stereoselective pinacol coupling reaction. Tetrahedron 2006, 62, 12137–12158. [Google Scholar] [CrossRef]
  49. Clerici, A.; Clerici, L.; Porta, O. A highly dl-stereoselective pinacolization of aromatic aldehydes mediated by titanium trichloride in dichloromethane. Tetrahedron Lett. 1996, 37, 3035–3038. [Google Scholar] [CrossRef]
  50. Clerici, A.; Pastori, N.; Porta, O. On the paradox of TiCl4 reducing power: pinacol coupling and two-carbon homologation of carbonyl compounds. Tetrahedron Lett. 2004, 45, 1825–1827. [Google Scholar] [CrossRef]
  51. Paradas, M.; Campaña, A.G.; Estévez, R.E.; de Cienfuegos, L.A.; Jiménez, T.; Robles, R.; Cuerva, J.M.; Oltra, J.E. Unexpected TiIII/Mn-promoted pinacol coupling of ketones. J. Org. Chem. 2009, 74, 3616–3619. [Google Scholar] [CrossRef] [PubMed]
  52. Bensari, A.; Renaud, J.-L.; Riant, O. Enantioselective pinacol coupling of aldehydes mediated and catalyzed by chiral titanium complexes. Org. Lett. 2001, 3, 3863–3865. [Google Scholar] [CrossRef] [PubMed]
  53. Chatterjee, A.; Bennur, T.H.; Joshi, N.N. Truly catalytic and enantioselective pinacol coupling of aryl aldehydes mediated by chiral Ti(III) complexes. J. Org. Chem. 2003, 68, 5668–5671. [Google Scholar] [CrossRef] [PubMed]
  54. Knoop, C.A.; Studer, A. A new method for the generation of Ti(III) complexes and its application in pinacol coupling reactions. Adv. Synth. Catal. 2005, 347, 1542–1546. [Google Scholar] [CrossRef]
  55. Sancho-Sanz, I.; Miguel, D.; Millán, A.; Estévez, R.E.; Oller-López, J.L.; Álvarez-Manzaneda, E.; Robles, R.; Cuerva, J.M.; Justicia, J. Titanocene(III)-promoted Barbier-type crotylation of carbonyl compounds. J. Org. Chem. 2011, 76, 732–735. [Google Scholar] [CrossRef] [PubMed]
  56. Estévez, R.E.; Justicia, J.; Bazdi, B.; Fuentes, N.; Paradas, M.; Choquesilo-Lazarte, D.; Garcia-Ruiz, J.M.; Robles, R.; Gansäuer, A.; Cuerva, J.M.; Oltra, J.E. Ti-catalyzed Barbier-type allylations and related reactions. Chem. Eur. J. 2009, 15, 2774–2791. [Google Scholar] [CrossRef] [PubMed]
  57. Jiménez, T.; Morcillo, S.P.; Martin-Lasanta, A.; Collado-Sanz, D.; Cárdenas, D.J.; Justicia, J.; Cuerva, J.M. Combining the power of TiIII-mediated processes for easy access to hydroxylated polycyclic terpenoids: Synthesis of sesterstatin 1 and C–D rings of aspergilloxide. Chem. Eur. J. 2012, 18, 12825–12833. [Google Scholar]
  58. Campaña, A.G.; Bazdi, B.; Fuentes, N.; Robles, R.; Cureva, J.M.; Oltra, J.E.; Porcel, S.; Echavarren, A.M. Divergent Titanium-Mediated Allylations with Modulation by Nickel or Palladium. Angew. Chem. Int. Ed. 2008, 47, 7625–7629. [Google Scholar] [CrossRef]
  59. Millán, A.; Campaña, A.G.; Bazdi, B.; Miguel, D.; de Cienfuegos, L.A.; Echavarren, A.M.; Cuerva, J.M. Ti/Pd bimetallic systems for the efficient allylation of carbonyl compounds and homocoupling reactions. Chem. Eur. J. 2011, 17, 3985–3994. [Google Scholar] [CrossRef] [PubMed]
  60. Millán, A.; de Cienfuegos, L.A.; Martin-Lasanta, A.; Campaña, A.G.; Cuerva, J.M. Titanium/Palladium-mediated regioselective propargylation of ketones using propargylic carbonates as pronucleophiles. Adv. Synth. Catal. 2011, 353, 73–78. [Google Scholar] [CrossRef]
  61. Millán, A.; Martin-Lasanta, A.; Miguel, D.; de Cienfuegos, L.A.; Cuerva, J.M. Ti/Pd-promoted intramolecular Michael-type addition of allylic carboxylates to activated alkenes. Chem. Commun. 2011, 47, 10470–10472. [Google Scholar] [CrossRef] [PubMed]
  62. Martínez-Peragón, Á.; Millán, A.; Campaña, A.G.; Rodríguez-Márquez, I.; Resa, S.; Miguel, D.; de Cienfuegos, L.A.; Cuerva, J.M. Ti/Ni-based multimetallic system for the efficient allylation of carbonyl compounds. Eur. J. Org. Chem. 2012, 1499–1503. [Google Scholar] [CrossRef]
  63. Mandal, S.K.; Paira, M.; Roy, S.C. Titanium(III) chloride mediated synthesis of furan derivatives: Synthesis of (±)-evodone. J. Chem. Sci. 2010, 122, 423–426. [Google Scholar] [CrossRef]
  64. Saha, S.; Mandal, S.K.; Roy, S.C. Titanocene(III) chloride mediated radical induced asymmetric synthesis of α-methylene bis-γ-butyrolactone. Tetrahedron Lett. 2011, 52, 3128–3130. [Google Scholar] [CrossRef]
  65. Rosales, A.; Muñoz-Bascón, J.; López-Sánchez, C.; Álvarez-Corral, M.; Muñoz-Dorado, M.; Rodríguez-García, I.; Oltra, J.E. Ti-catalyzed homolytic opening of ozonides: A sustainable C−C bond-forming reaction. J. Org. Chem. 2012, 77, 4171–4176. [Google Scholar] [CrossRef] [PubMed]
  66. Streuff, J.; Feurer, M.; Bichovski, P.; Frey, G.; Gellrich, U. Enantioselective Titanium(III)-catalyzed reductive cyclization of ketonitriles. Angew. Chem. Int. Ed. 2012, 51, 8661–8664. [Google Scholar] [CrossRef] [PubMed]
  67. Griller, D.; Ingold, K.U. Persistent carbon-centered radicals. Acc. Chem. Res. 1976, 9, 13–19. [Google Scholar] [CrossRef]
  68. Spaccini, R.; Pastori, N.; Clerici, A.; Punta, C.; Porta, O. Key role of Ti(IV) in the selective radical-radical cross-coupling mediated by the Ingold-Fischer effect. J. Am. Chem. Soc. 2008, 130, 18018–18024. [Google Scholar] [CrossRef] [PubMed]
  69. Gansäuer, A.; Justicia, J.; Fan, C.-A.; Worgull, D.; Piestert, F. Reductive C-C bond formation after opening via electron transfer. Top. Curr. Chem. 2007, 279, 25–52. [Google Scholar]
  70. Gansäuer, A.; Lauterbach, T.; Narayan, S. Strained heterocycles in radical chemistry. Angew. Chem. Int. Ed. 2003, 42, 5556–5573. [Google Scholar] [CrossRef] [PubMed]
  71. Daasbjerg, K.; Svith, H.; Grimme, S.; Gerenkamp, M.; Muck-Lichtenfeld, C.; Gansäuer, A.; Barchuk, A.; Keller, F. Elucidation of the mechanism of titanocene-mediated epoxide opening by a combined experimental and theoretical approach. Angew. Chem. Int. Ed. 2006, 45, 2041–2044. [Google Scholar] [CrossRef] [PubMed]
  72. Gansäuer, A.; Barchuk, A.; Keller, F.; Schmitt, M.; Grimme, S.; Gerenkamp, M.; Muck-Lichtenfeld, C.; Daasbjerg, K.; Svith, H. Mechanism of titanocene-mediated epoxide opening through homolytic substitution. J. Am. Chem. Soc. 2007, 129, 1359–1371. [Google Scholar] [CrossRef] [PubMed]
  73. RajanBabu, T.V.; Nugent, W.A. Intermolecular addition of epoxides to activated olefins: a new reaction. J. Am. Chem. Soc. 1989, 111, 4525–4527. [Google Scholar] [CrossRef]
  74. RajanBabu, T.V.; Nugent, W.A.; Beattie, M.S. Free radical-mediated reduction and deoxygenation of epoxides. J. Am. Chem. Soc. 1990, 112, 6408–6409. [Google Scholar] [CrossRef]
  75. Cuerva, J.M.; Juan, J.C.; Justicia, J.; Oller-Lòpez, J.L.; Oltra, J.E. Cp2TiCl in natural product synthesis. Top. Curr. Chem. 2006, 264, 63–91. [Google Scholar]
  76. Gansäuer, A.; Fleckhaus, A. Epoxides in titanocenes-mediated and -catalyzed radical reactions. In Encyclopedia of Radicals in Chemistry, Biology and Materials; Chatgilialoglu, C., Studer, A., Eds.; Wiley: Chichester, UK, 2012; Volume 2, pp. 989–1001. [Google Scholar]
  77. Fernández-Mateos, A.; Madrazo, S.E.; Teijón, P.H.; González, R.R. Titanocene-promoted eliminations on epoxy alcohols and epoxy esters. Eur. J. Org. Chem. 2010, 856–861. [Google Scholar] [CrossRef]
  78. Fernández-Mateos, A.; Teijón, P.H. Radical reactions on pinene-oxide derivatives induced by Ti(III). Tetrahedron 2011, 67, 9529–9534. [Google Scholar]
  79. Hardouin, C.; Doris, E.; Rousseau, B.; Mioskowski, C. Concise synthesis of anhydrovinblastine from leurosine. Org. Lett. 2002, 4, 1151–1153. [Google Scholar] [CrossRef] [PubMed]
  80. Barrero, A.F.; Quìlez del Moral, J.F.; Sànchez, E.M.; Arteaga, J.F. Regio- and diastereo-selective reductive coupling of vinylepoxides catalyzed by titanocene chloride. Org Lett. 2006, 8, 669–672. [Google Scholar] [CrossRef] [PubMed]
  81. Aldegunge, M.J.; Astedo, L.; Granja, J.R. Synthesis of 2-ene-1,4-diols by a new cascade-opening of 1,3-diepoxides: towards an efficient synthesis of dihydroxytaxoid derivatives. Chem. Eur. J. 2009, 15, 4785–4787. [Google Scholar] [CrossRef] [PubMed]
  82. Justicia, J.; Jimenez, T.; Morcillo, S.P.; Cuerva, J.M.; Oltra, J.E. Mixed disproportionation versus radical trapping in titanocene (III)-promoted epoxide openings. Tetrahedron 2009, 65, 10837–10841. [Google Scholar] [CrossRef]
  83. Barrero, A.F.; Rosales, A.; Cuerva, J.M.; Oltra, J.E. Unified synthesis of eudesmanolides, combining biomimetic strategies with homogeneous catalysis and free-radical chemistry. Org. Lett. 2003, 5, 1935–1938. [Google Scholar] [CrossRef] [PubMed]
  84. Justicia, J.; Rosales, B.E.; Oller-Lopez, J.L.; Valdivia, N.; Haidour, A.; Oltra, J.E.; Barrero, A.F.; Cardenas, D.J.; Cuerva, J.M. Titanocene-catalyzed cascade cyclization of epoxypolyprenes: Straightforward synthesis of terpenoids by free-radical chemistry. Chem. Eur. 2004, 10, 1778–1788. [Google Scholar] [CrossRef] [PubMed]
  85. Justicia, J.; Oltra, J.E.; Cuerva, J.M. Exploiting Pd-II and Ti-III chemistry to obtain gamma-dioxygenated terpenoids: Synthesis of rostratone and novel approaches to aphidicolin and pyripyropene A. J. Org. Chem. 2005, 70, 8265–8272. [Google Scholar] [CrossRef] [PubMed]
  86. Gansäuer, A.; Worgull, D.; Justicia, J. Catalytic epoxypolyene cyclization via radicals: A simple total synthesis of sclareol oxide and its 8-epimer. Synthesis-Stuttgart 2006, 2151–2154. [Google Scholar]
  87. Justicia, J.; Campanã, G.A.; Bazdi, B.; Robles, R.; Cuerva, J.M.; Oltra, J.E. Titanium-catalyzed enantioselective synthesis of alpha-ambrinol. Adv. Synth. Catal. 2008, 350, 571–576. [Google Scholar] [CrossRef]
  88. Gansäuer, A.; Bluhm, H. Reagent-controlled transition-metal-catalyzed radical reactions. Chem. Rev. 2000, 100, 2771–2788. [Google Scholar] [CrossRef] [PubMed]
  89. Cuerva, J.M.; Campanã, A.G.; Justicia, J.; Rosales, A.; Oller-Lopez, J.L.; Robles, R.; Cardenas, D.J.; Bunuel, E.; Oltra, J.E. Water: The ideal hydrogen-atom source in free-radical chemistry mediated by Ti-III and other single-electron-transfer metals? Angew. Chem. Int. Ed. 2006, 45, 5522–5526. [Google Scholar] [CrossRef] [PubMed]
  90. Paradas, M.; Campaña, A.G.; Marcos, M.L.; Justicia, J.; Haidour, A.; Robles, R.; Cárdenas, D.J.; Oltra, E.J.; Cuerva, J.M. Unprecedented H-atom transfer from water to ketyl radicals mediated by Cp2TiCl. Dalton Trans. 2010, 39, 8796–8800. [Google Scholar] [CrossRef] [PubMed]
  91. Paradas, M.; Campaña, A.G.; Jiménez, T.; Robles, R.; Oltra, J.E.; Buñuel, E.; Justicia, J.; Cárdenas, D.J.; Cuerva, J.M. Understanding the exceptional hydrogen-atom donor characteristics of water in TiIII- mediated free-radical chemistry. J. Am. Chem. Soc. 2010, 132, 12748–12756. [Google Scholar] [CrossRef] [PubMed]
  92. Jiménez, T.; Campaña, A.G.; Bazdi, B.; Paradas, M.; Arráez-Román, D.; Segura-Carretero, A.; Fernández-Gutiérrez, A.; Oltra, J.E.; Robles, R.; Justicia, J.; Cuerva, J.M. Radical reduction of epoxides using a titanocene (III) water system: synthesis of β-deuterated alcohols and their use as internal standards in food analysis. Eur. J. Org. Chem. 2010, 4288–4295. [Google Scholar] [CrossRef]
  93. Gansäuer, A.; Otte, M.; Shi, L. Radical cyclizations terminated by Ir-catalyzed hydrogen atom transfer. J. Am. Chem. Soc. 2011, 133, 416–417. [Google Scholar] [CrossRef] [PubMed]
  94. Gansäuer, A.; Fan, C.A.; Piester, F. Sustainable radical reduction through catalytic hydrogen atom transfer. J. Am. Chem. Soc. 2008, 130, 6916–6917. [Google Scholar] [CrossRef] [PubMed]
  95. Gansäuer, A.; Narayan, S. Titanocene-catalysed electron transfer-mediated opening of epoxides. Adv. Synth. Cat. 2002, 344, 465–475. [Google Scholar] [CrossRef]
  96. Gansäuer, A.; Bluhm, H.; Lauterbach, T. Titanocene-catalysed enantioselective opening of meso-epoxides. Adv. Synth. Catal. 2001, 343, 785–787. [Google Scholar] [CrossRef]
  97. Gansäuer, A.; Bluhm, H.; Rinker, B.; Narayan, S.; Schick, M.; Lauterbach, T.; Pierobon, M. Reagent-controlled stereoselectivity in titanocene-catalyzed epoxide openings: Reductions and intermolecular additions to alpha,beta-unsaturated carbonyl compounds. Chem. Eur. J. 2003, 9, 531–542. [Google Scholar] [CrossRef] [PubMed]
  98. Gansäuer, A.; Narayan, S.; Schiffer-Ndene, N.; Bluhm, H.; Oltra, J.E.; Cuerva, J.M.; Rosales, A.; Nieger, M. An improved synthesis of Kagan’s menthyl substituted titanocene and zirconocene dichloride, comparison of their crystal structures, and preliminary catalyst evaluation. J. Organomet. Chem. 2006, 691, 2327–2331. [Google Scholar] [CrossRef]
  99. Gansäuer, A.; Bluhm, H.; Pierobon, M.; Keller, M. Conformational preferences of titanocene dichlorides with ligands derived from menthol: Comparison of structures in solution and in the crystal. Organometallics 2001, 20, 914–919. [Google Scholar] [CrossRef]
  100. Gansäuer, A.; Fan, C.-A.; Keller, F.; Karbaum, P. Regiodivergent epoxide openining: A concept in stereoselective catalysis beyond classical kinetic resolution and desymmetrizations. Chem. Eur. J. 2007, 13, 8084–8090. [Google Scholar] [CrossRef] [PubMed]
  101. Gansäuer, A.; Shi, L.; Otte, M. Catalytic enantioselective radical cyclization via regiodivergent epoxide opening. J. Am. Chem. Soc. 2010, 132, 11858–11859. [Google Scholar] [CrossRef] [PubMed]
  102. Gansauer, A.; Fan, C.-A.; Keller, F.; Keil, J. Titanocene-catalyzed regiodivergent epoxide openings. J. Am. Chem. Soc. 2007, 129, 3484–3485. [Google Scholar] [CrossRef] [PubMed]
  103. Barrero, A.F.; Rosales, A.; Cuerva, J.M.; Gansäuer, A.; Oltra, J.E. Titanocene-catalyzed, selective reduction of ketones in aqueous media. A safe, mild, inexpensive procedure for the synthesis of secondary alcohols via radical chemistry. Tetrahedron Lett. 2003, 44, 1079–1082. [Google Scholar] [CrossRef]
  104. Oller-López, J.L.; Campaña, A.G.; Cuerva, J.M.; Oltra, J.E. Aromatic carbonyl compound reduction and pinacol coupling processes mediated by titanocene(III)/Zn in water. Synthesis 2005, 2619–2622. [Google Scholar]
  105. Chakraborty, T.K.; Das, S. Synthesis of chiral 1,3-diols by radical-mediated regioselective opening of 2,3-epoxy alcohols using Cp(2)TiCl. Tetrahedron Lett. 2002, 43, 2313–2315. [Google Scholar] [CrossRef]
  106. Hardouin, C.; Chevallier, F.; Rousseau, B.; Doris, E. Cp2TiCl-mediated selective reduction of alpha,beta-epoxy ketones. J. Org. Chem. 2001, 66, 1046–1048. [Google Scholar] [CrossRef] [PubMed]
  107. Bermejo, F.; Sandoval, C. Cp2TiCl-promoted isomerization of trisubstituted epoxides to exo-methylene allylic alcohols on carvone derivatives. J. Org. Chem. 2004, 69, 5275–5280. [Google Scholar] [CrossRef] [PubMed]
  108. Barrero, A.F.; Cuerva, J.M.; Alvarez-Manzaneda, E.J.; Oltra, J.E.; Chahboun, R. First synthesis of achilleol A using titanium(III) chemistry. Tetrahedron Lett. 2002, 43, 2793–2796. [Google Scholar] [CrossRef]
  109. Barrero, A.F.; Cuerva, J.M.; Herrador, M.M.; Valdivia, M.V. A new strategy for the synthesis of cyclic terpenoids based on the radical opening of acyclic epoxypolyenes. J. Org. Chem. 2001, 66, 4074–4078. [Google Scholar] [CrossRef] [PubMed]
  110. Barrero, A.F.; Oltra, J.E.; Cuerva, J.M.; Rosales, A. Effects of solvents and water in Ti(III)-mediated radical cyclizations of epoxygermacrolides. Straightforward synthesis and absolute stereochemistry of (+)-3α-hydroxyreynosin and related eudesmanolides. J. Org. Chem. 2002, 67, 2566–2571. [Google Scholar] [CrossRef] [PubMed]
  111. Nakai, K.; Kamoshita, M.; Doi, T.; Yamada, H.; Takahashi, T. Stereo- and regio-selective Ti-mediated radical cyclization of epoxy-alkenes: Synthesis of the A and C ring synthons of paclitaxel. Tetrahedron Lett. 2001, 42, 7855–7857. [Google Scholar] [CrossRef]
  112. Gansäuer, A.; Klatte, M.; Brändle, G.M.; Friedrich, J. Catalytic hydrogen atom transfer (HAT) for sustainable and diastereoselective radical reduction. Angew. Chem. Int. Ed. 2012, 51, 8891–8894. [Google Scholar] [CrossRef] [PubMed]
  113. Chakraborty, T.K.; Samanta, R.; Das, S. Radical-mediated opening of 2,3-epoxy alcohols using Cp2TiCl: Stereoselective construction of quaternary chiral centers. J. Org. Chem. 2006, 71, 3321–3324. [Google Scholar] [CrossRef] [PubMed]
  114. Dötz, K.H.; Gomes da Silva, E. Carbohydrate-modified fused pyranosylidene complexes via radical addition of epoxides to unsaturated metal carbenes. Tetrahedron 2000, 56, 8291–8299. [Google Scholar] [CrossRef]
  115. Fernández-Mateos, A.; Martín de la Nava, E.; Pascual Coca, G.; Ramos Silvo, A.; Rubio González, R. Radicals from epoxides. Intramolecular addition to aldehyde and ketone carbonyls. Org. Lett. 1999, 1, 607–609. [Google Scholar] [CrossRef]
  116. Fernández-Mateos, A.; Mateos Burón, I.; Rabanedo Clemente, R.; Ramos Silvo, A.I.; Rubio González, R. Radical cyclization of epoxynitriles mediated by titanocene chloride. Synlett 2004, 1011–1014. [Google Scholar] [CrossRef]
  117. Gansäuer, A.; Lauterbach, T.; Geich-Gimbel, D. Polarity matching of radical trapping: High yielding 3-exo and 4-exo cyclizations. Chem. Eur. J. 2004, 10, 4983–4990. [Google Scholar] [CrossRef] [PubMed]
  118. Fernández-Mateos, A.; Mateos Burón, L.; Martín de la Nava, E.M.; Rabanedo Clemente, R.; Rubio González, R.; Sanz González, F. Effect of tether length on Ti(III)-mediated cyclization of epoxyalkenes and unsaturated epoxyketones. Synlett 2004, 2553–2557. [Google Scholar] [CrossRef]
  119. Friedrich, J.; Dolg, M.; Gansäuer, A.; Geich-Gimbel, D.; Lauterbach, T. A combined theoretical and experimental study of efficient and fast titanocene-catalyzed 3-exo cyclizations. J. Am. Chem. Soc. 2005, 127, 7071–7077. [Google Scholar] [CrossRef] [PubMed]
  120. Justicia, J.; de Cienfuegos, L.A.; Campaña, A.G.; Miguel, D.; Jakoby, V.; Gansäuer, A.; Cuerva, J.M. Bioinspired terpene synthesis: a radical approach. Chem. Soc. Rev. 2011, 40, 3525–3537. [Google Scholar] [CrossRef] [PubMed]
  121. Ruano, G.; Grande, M.; Anaya, J.J. Stereospecific synthesis of highly functionalized tricyclic beta-lactams by radical cyclizations using titanocene monochloride. J. Org. Chem. 2002, 67, 8243–8246. [Google Scholar] [CrossRef] [PubMed]
  122. Ruano, G.; Matiánez, J.; Grande, M. Stereospecific synthesis of polyfunctionalized carbacephams induced by titanocene(III) chloride. J. Org. Chem. 2003, 68, 2024–2027. [Google Scholar] [CrossRef] [PubMed]
  123. Friedrich, J.; Walczak, K.; Dolg, M.; Piestert, F.; Lauterbach, T.; Worgull, D.; Gansäuer, A. Titanocene catalyzed 4-exo cyclizations: Mechanism, experiment, catalyst design. J. Am. Chem. Soc. 2008, 130, 1788–1796. [Google Scholar] [CrossRef] [PubMed]
  124. Gansäuer, A.; Worgull, D.; Knebel, K.; Huth, I.; Schnakenburg, G. 4-Exo cyclizations by template catalysis. Angew. Chem. Int. Ed. 2009, 48, 8882–8885. [Google Scholar] [CrossRef] [PubMed]
  125. Gansäuer, A.; Greb, A.; Huth, I.; Worgull, D.; Knebel, K. Formal total synthesis of (+/-)-fragranol via template catalyzed 4-exo cyclization. Tetrahedron 2009, 65, 10791–10796. [Google Scholar] [CrossRef]
  126. Gansäuer, A.; Knebel, K.; Kube, C.; van Gastel, M.; Cangonul, A.; Daasbjerg, K.; Hangele, T.; Hulsen, M.; Dolg, M.; Friedrich, J. Radical 4-exo cyclization via template catalysis. Chem. Eur. J. 2012, 18, 2591–2599. [Google Scholar] [CrossRef] [PubMed]
  127. Ziegler, F.E.; Sarpong, M.A. Radical cyclization studies directed toward the synthesis of BMS-200475 ‘entecavir’: The carbocyclic core. Tetrahedron 2003, 59, 9013–9018. [Google Scholar] [CrossRef]
  128. Fernández-Mateos, A.; Herrero Teijón, P.; Rabanes Clemente, R.; Rubio Gonzalez, R. Radical reactions of epoxy esters induced by titanocene chloride. Tetrahedron Lett. 2006, 47, 7755–7758. [Google Scholar] [CrossRef]
  129. Leca, D.; Fensterbank, L.; Lacôte, E.; Malacria, M. Titanium-mediated domino radical cyclization/beta elimination of phosphine oxides. Angew Chem. Int. Ed. 2004, 43, 4220–4222. [Google Scholar] [CrossRef] [PubMed]
  130. Leca, D.; Song, K.; Albert, M.; Goançalves, M.G.; Fensterbank, L.; Lacôte, E.; Malacria, M. Titanocene-mediated intramolecular radical vinylations. Synthesis 2005, 1405–1420. [Google Scholar] [CrossRef]
  131. Gansäuer, A.; Rinker, B.; Pierobon, M.; Grimme, S.; Gerenkamp, M.; Muck-Lichtenfeld, C. A radical tandem reaction with homolytic cleavage of a Ti-O bond. Angew. Chem. Int. Ed. 2003, 42, 3687–3690. [Google Scholar] [CrossRef] [PubMed]
  132. Gansäuer, A.; Rinker, B.; Ndene-Schiffer, N.; Pierobon, M.; Grimme, S.; Gerenkamp, M.; Muck-Lichtenfeld, C. A radical roundabout for an unprecedented tandem reaction including a homolytic substitution with a titanium-oxygen bond. Eur. J. Org. Chem. 2004, 2337–2351. [Google Scholar] [CrossRef]
  133. Gansäuer, A.; Fleckhaus, A.; Lafont, M.A.; Okkel, A.; Kotsis, K.; Anoop, A.; Neese, F. Catalysis via homolytic substitutions with C-O and Ti-O bonds: Oxidative additions and reductive eliminations in single electron steps. J. Am. Chem. Soc. 2009, 131, 16989–16999. [Google Scholar] [CrossRef] [PubMed]
  134. Apte, S.; Radetich, B.; Shin, S.; RajanBabu, T.V. Silylstannylation of highly functionalized acetylenes. Synthesis of precursors for annulations via radical or Heck reactions. Org. Lett. 2004, 6, 4053–4056. [Google Scholar] [CrossRef] [PubMed]
  135. Gansäuer, A.; Pierobon, M.; Bluhm, H. Stereoselective synthesis of tri- and tetrasubstituted olefins by tandem cyclization addition reactions featuring vinyl radicals. Angew. Chem. Int. Ed. 2002, 41, 3206–3208. [Google Scholar] [CrossRef]
  136. Trost, B.M.; Shen, H.C.; Surivet, J.P. An enantioselective biomimetic total synthesis of (−)-siccanin. Angew. Chem. Int. Ed. 2003, 42, 3943–3947. [Google Scholar] [CrossRef] [PubMed]
  137. Trost, B.M.; Shen, H.C.; Surivet, J.P. Biomimetic enantioselective total synthesis of (-)-siccanin via the Pd-catalyzed asymmetric allylic alkylation (AAA) and sequential radical cyclizations. J. Am. Chem. Soc. 2004, 126, 12565–12579. [Google Scholar] [CrossRef] [PubMed]
  138. Haruo, Y.; Hasegawa, T.; Tanaka, H.; Takahashi, T. Total synthesis of (+/-)-smenospondiol by titanium(III)-mediated tandem radical cyclization. Synlett 2001, 1935–1937. [Google Scholar] [CrossRef]
  139. Arteaga, J.F.; Diéguez, H.R.; González-Delgado, J.A.; Quílez del Moral, J.F.; Barrero, A.F. Control of the regio- and diastereoselectivity for the preparation of highly functionalized terpenic cyclopentanes through radical cyclization. Eur. J. Org. Chem. 2011, 5002–5011. [Google Scholar] [CrossRef]
  140. Saha, S.; Roy, S.C. Titanocene (III) chloride mediated radical induced synthesis of (-)-methylenolactocin and (-)-protolichesterinic acid. Tetrahedron 2010, 66, 4278–4283. [Google Scholar] [CrossRef]
  141. Fernández-Mateos, A.; Ramos Silvo, A.I.; González, R.R.; Simmonds, M.S.J. Synthesis of CDE molecular fragments related to sendanin mediated by titanocene(III). Org. Biomol. Chem. 2012, 10, 5620–5628. [Google Scholar] [CrossRef] [PubMed]
  142. Cuerva, J.M.; Justicia, J.; Oller-López, J.L.; Bazdi, B.; Oltra, J.E. The growing impact of titanocene(III)-mediated radical epoxide opening on the synthesis of natural products. Mini-Rev. Org. Chem. 2006, 3, 23–35. [Google Scholar] [CrossRef]
  143. Gansäuer, A.; Justicia, J.; Rosales, A.; Worgull, D.; Rinker, B.; Cuerva, J.M.; Oltra, J.E. Transition-metal-catalyzed allylic substitution and titanocene-catalyzed epoxypolyene cyclization as a powerful tool for the preparation of terpenoids. Eur. J. Org. Chem. 2006, 4115–4127. [Google Scholar] [CrossRef]
  144. Yamaoka, M.; Fukatsu, Y.; Nakazaki, A.; Kobayashi, S. Synthetic study of fomitellic acids: Construction of the AB ring moiety. Tetrahedron Lett. 2009, 50, 3849–3852. [Google Scholar] [CrossRef]
  145. Justicia, J.; Oller-López, J.L.; Campaña, A.G.; Oltra, J.E.; Cuerva, J.M.; Buñuel, E.; Cárdenas, D.J. 7-Endo radical cyclizations catalyzed by titanocene (III). Straightforward synthesis of terpenoids with seven-membered carbocycles. J. Am. Chem. Soc. 2005, 127, 14911–14921. [Google Scholar] [CrossRef] [PubMed]
  146. Mandal, S.K.; Roy, S.C. Titanocene(III) mediated 8-endo radical cyclizations for the synthesis of eight-membered cyclic ethers. Tetrahedron Lett. 2006, 47, 1599–1601. [Google Scholar] [CrossRef]
  147. di Lena, F.; Matyjaszewski, K. Transition metal catalysts for controlled radical polymerization. Prog. Polym. Sci. 2010, 35, 959–1021. [Google Scholar] [CrossRef]
  148. Nomura, K. New approaches in precise synthesis of polyolefins containing polar functionalities by olefin copolymerizations using transition metal catalysis. J. Synth. Org. Chem. Jpn. 2012, 68, 1150–1158. [Google Scholar] [CrossRef]
  149. Kwark, Y.-J.; Kim, J.; Novak, B.M. Titanium complexes: A possible catalyst for controlled radical polymerization. Macromol. Res. 2007, 15, 31–38. [Google Scholar] [CrossRef]
  150. Bhattacharjee, M.; Patra, B.N. [CP2TiCl2] catalyzed polymerization in water: polymerization of methylmethacrylate to a high molecular weight polymer. Polymer 2004, 45, 3111–3114. [Google Scholar] [CrossRef]
  151. Asandei, A.D.; Moran, I.W. TiCp2Cl-catalyzed living radical polymerization of styrene initiated by oxirane radical ring opening. J. Am. Chem. Soc. 2004, 126, 15932–15933. [Google Scholar] [CrossRef] [PubMed]
  152. Ma, L.; Liu, W.; Sheng, Y.; Huang, Q.; Yang, W. Synthesis of well-defined star-shaped organosiloxane-functionalized polymethylmethacrylate promoted by epoxide-derived titanocene alkoxides via radical polymerization. J. Appl. Polym. Sci. 2011, 12, 1652–1658. [Google Scholar] [CrossRef]
  153. Asandei, A.D.; Saha, G. Living ring-opening polymerization of cyclic esters with epoxide-derived titanium alkoxides. Macromol. Rapid. Commun. 2005, 26, 626–631. [Google Scholar] [CrossRef]
  154. Asandei, A.D.; Simpson, C.P.; Yu, H.S. Cp2TiCl-catalyzed controlled radical polymerization of isoprene initiated from epoxides, aldehydes and halides. Polym. Prepr. Am. Chem. Soc. 2008, 49, 73–74. [Google Scholar]
  155. Asandei, A.D.; Simpson, C.P. Cp2TiCl-catalyzed synthesis of styrene/isoprene copolymers by controlled radical polymerization. Polym. Prepr. Am. Chem. Soc. 2008, 49, 75–76. [Google Scholar]
  156. Asandei, A.D.; Moran, I.W. The ligand effect in Ti-mediated living radical styrene polymerizations initiated by epoxide radical ring opening. I. Alkoxide and bisketonate Ti complexes. J. Polym. Sci. Part A Polym. Chem. 2005, 43, 6028–6038. [Google Scholar] [CrossRef]
  157. Asandei, A.D.; Moran, I.W. The ligand effect in Ti-mediated living radical styrene polymerizations initiated by epoxide radical ring opening. II. Scorpionate and half-sandwich LTiCl3 complexes. J. Polym. Sci. Part A Polym. Chem. 2005, 43, 6039–6047. [Google Scholar] [CrossRef]
  158. Asandei, A.D.; Moran, I.W. The ligand effect in Ti-mediated living radical styrene polymerizations initiated by epoxide radical ring opening. III. Substituted sandwich metallocenes. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 1060–1070. [Google Scholar] [CrossRef]
  159. Asandei, A.D.; Moran, I.W.; Saha, G.; Chen, Y. Titanium-mediated living radical styrene polymerizations. V. Cp2TiCl-catalyzed initiation by epoxide radical ring opening: Effect of solvents and additives. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 2015–2026. [Google Scholar] [CrossRef]
  160. Asandei, A.D.; Moran, I.W.; Saha, G.; Chen, Y. Titanium-mediated living radical styrene polymerizations. VI. Cp2TiCl-catalyzed initiation by epoxide radical ring opening: effect of the reducing agents, temperature, and titanium/epoxide and titanium/zinc ratios. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 2156–2165. [Google Scholar] [CrossRef]
  161. Asandei, A.D.; Chen, Y. Cp2TiCl-catalyzed set reduction of aldehydes: A new initiating protocol for living radical polymerization. Macromolecules 2006, 39, 7549–7554. [Google Scholar] [CrossRef]
  162. Asandei, A.D.; Chen, Y.; Adebolu, O.I.; Simpson, C.P. Living ring-opening polymerization of ɛ-caprolactone with Ti alkoxides derived from the Cp2TiCl-catalyzed SET reduction of aldehydes. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 2869–2877. [Google Scholar] [CrossRef]
  163. Tehfe, M.A.; Lalevèe, J.; Morlet-Savary, F.; Graff, B.; Fouassier, J.P. On the use of bis(cyclopentadienyl)titanium(IV) dichloride in visible-light-induced ring-opening photopolymerization. Macromolecules 2012, 45, 356–361. [Google Scholar] [CrossRef]
Scheme 1. Arylative amination of aldehydes promoted by the TiCl3/ArN2+ system.
Scheme 1. Arylative amination of aldehydes promoted by the TiCl3/ArN2+ system.
Molecules 17 14700 sch001
Scheme 2. Nucleophilic radical addition to aldimines promoted by the TiCl3/PhN2+ system.
Scheme 2. Nucleophilic radical addition to aldimines promoted by the TiCl3/PhN2+ system.
Molecules 17 14700 sch002
Scheme 3. Radical addition to imines triggered by TiCl3 or TiCl4-Zn/ROOH systems.
Scheme 3. Radical addition to imines triggered by TiCl3 or TiCl4-Zn/ROOH systems.
Molecules 17 14700 sch003
Scheme 4. Aminoalkylation of ethers promoted by the TiCl3/t-BuOOH system.
Scheme 4. Aminoalkylation of ethers promoted by the TiCl3/t-BuOOH system.
Molecules 17 14700 sch004
Scheme 5. Nucleophilic radical generation promoted by the TiCl3/t-BuOOH system.
Scheme 5. Nucleophilic radical generation promoted by the TiCl3/t-BuOOH system.
Molecules 17 14700 sch005
Scheme 6. (a) Carbamoylation of aldimines. (b) Hydroxymethylation of aldimines.
Scheme 6. (a) Carbamoylation of aldimines. (b) Hydroxymethylation of aldimines.
Molecules 17 14700 sch006
Scheme 7. Domino reaction in alcoholic solvent promoted by the TiCl3/t-BuOOH system.
Scheme 7. Domino reaction in alcoholic solvent promoted by the TiCl3/t-BuOOH system.
Molecules 17 14700 sch007
Scheme 8. (a) Carbamoylation of ketimines. (b) Hydroxymethylation of ketimines.
Scheme 8. (a) Carbamoylation of ketimines. (b) Hydroxymethylation of ketimines.
Molecules 17 14700 sch008
Scheme 9. Nucleophilic addition to ketimines promoted by the TiCl4-Zn/ROOH system.
Scheme 9. Nucleophilic addition to ketimines promoted by the TiCl4-Zn/ROOH system.
Molecules 17 14700 sch009
Scheme 10. Domino reaction in alcohols promoted by the TiCl4-Zn/ROOH system.
Scheme 10. Domino reaction in alcohols promoted by the TiCl4-Zn/ROOH system.
Molecules 17 14700 sch010
Scheme 11. Domino reaction in cyclic ethers promoted by the TiCl4-Zn/ROOH system.
Scheme 11. Domino reaction in cyclic ethers promoted by the TiCl4-Zn/ROOH system.
Molecules 17 14700 sch011
Scheme 12. Cp2TiCl-mediated allylation of aromatic aldimines.
Scheme 12. Cp2TiCl-mediated allylation of aromatic aldimines.
Molecules 17 14700 sch012
Scheme 13. dl-Stereodirecting pinacolization of aromatic aldehydes promoted by TiCl3.
Scheme 13. dl-Stereodirecting pinacolization of aromatic aldehydes promoted by TiCl3.
Molecules 17 14700 sch013
Scheme 14. Proposed mechanism for the TiCl3/t-BuOOH mediated pinacolization.
Scheme 14. Proposed mechanism for the TiCl3/t-BuOOH mediated pinacolization.
Molecules 17 14700 sch014
Figure 1. Chiral Ti(IV) complexes.
Figure 1. Chiral Ti(IV) complexes.
Molecules 17 14700 g001
Scheme 15. Riant’s pinacolization mediated by stoichiometric amounts of complex 1.
Scheme 15. Riant’s pinacolization mediated by stoichiometric amounts of complex 1.
Molecules 17 14700 sch015
Scheme 16. Pinacolization mediated by catalytic amounts of complex 2.
Scheme 16. Pinacolization mediated by catalytic amounts of complex 2.
Molecules 17 14700 sch016
Scheme 17. Pinacolization of aromatic aldehydes promoted by 3.
Scheme 17. Pinacolization of aromatic aldehydes promoted by 3.
Molecules 17 14700 sch017
Scheme 18. Barbier-type allylation of carbonyl derivatives.
Scheme 18. Barbier-type allylation of carbonyl derivatives.
Molecules 17 14700 sch018
Scheme 19. Barbier-type allylation mechanism promoted by catalytic titanocene.
Scheme 19. Barbier-type allylation mechanism promoted by catalytic titanocene.
Molecules 17 14700 sch019
Scheme 20. Examples of Barbier-type allylations promoted by catalytic titanocene.
Scheme 20. Examples of Barbier-type allylations promoted by catalytic titanocene.
Molecules 17 14700 sch020
Scheme 21. Divergent Ti-mediated allylations with modulation by Palladium or Nickel.
Scheme 21. Divergent Ti-mediated allylations with modulation by Palladium or Nickel.
Molecules 17 14700 sch021
Scheme 22. Titanium/Palladium mediated reactions.
Scheme 22. Titanium/Palladium mediated reactions.
Molecules 17 14700 sch022
Scheme 23. Titanium/palladium-mediated propargylation of ketones.
Scheme 23. Titanium/palladium-mediated propargylation of ketones.
Molecules 17 14700 sch023
Scheme 24. Titanium/palladium-mediated intramolecular Michael-type addition.
Scheme 24. Titanium/palladium-mediated intramolecular Michael-type addition.
Molecules 17 14700 sch024
Scheme 25. Homocoupling versus cross-coupling in the Ti-catalyzed opening of ozonides.
Scheme 25. Homocoupling versus cross-coupling in the Ti-catalyzed opening of ozonides.
Molecules 17 14700 sch025
Scheme 26. Reductive cyclization of ketonitriles.
Scheme 26. Reductive cyclization of ketonitriles.
Molecules 17 14700 sch026
Scheme 27. C-C bond cleavage of α,β-dihydroxy ketones promoted by TiCl4.
Scheme 27. C-C bond cleavage of α,β-dihydroxy ketones promoted by TiCl4.
Molecules 17 14700 sch027
Scheme 28. Synthesis of β-hydroxynitriles from α,β-dihydroxy ketones and AIBN.
Scheme 28. Synthesis of β-hydroxynitriles from α,β-dihydroxy ketones and AIBN.
Molecules 17 14700 sch028
Scheme 29. Structures of Cp2TiCl in solution: dimer (a), half-open dimer (b), reactive species binding the solvent (c).
Scheme 29. Structures of Cp2TiCl in solution: dimer (a), half-open dimer (b), reactive species binding the solvent (c).
Molecules 17 14700 sch029
Scheme 30. Deoxygenation of leurosine.
Scheme 30. Deoxygenation of leurosine.
Molecules 17 14700 sch030
Scheme 31. Dimerization of vinylepoxides for the preparation of homo-onecerans.
Scheme 31. Dimerization of vinylepoxides for the preparation of homo-onecerans.
Molecules 17 14700 sch031
Scheme 32. Disproportionation of epoxide through hydrogen atom transfer.
Scheme 32. Disproportionation of epoxide through hydrogen atom transfer.
Molecules 17 14700 sch032
Scheme 33. Titanocene-mediated reductive epoxide opening by 1,4-cyclohexadiene.
Scheme 33. Titanocene-mediated reductive epoxide opening by 1,4-cyclohexadiene.
Molecules 17 14700 sch033
Figure 2. Different substitution patterns at the cyclopentadienyl ligand (Cp).
Figure 2. Different substitution patterns at the cyclopentadienyl ligand (Cp).
Molecules 17 14700 g002
Scheme 34. Mechanism of the titanocene-catalyzed reductive epoxide opening.
Scheme 34. Mechanism of the titanocene-catalyzed reductive epoxide opening.
Molecules 17 14700 sch034
Scheme 35. Example of elective reduction of α,β-epoxyketones.
Scheme 35. Example of elective reduction of α,β-epoxyketones.
Molecules 17 14700 sch035
Scheme 36. Reduction of epoxides mediated by [Cp2TiOC2H5)2]/(CH3)PhSiH2 system.
Scheme 36. Reduction of epoxides mediated by [Cp2TiOC2H5)2]/(CH3)PhSiH2 system.
Molecules 17 14700 sch036
Scheme 37. Intermolecular C-C addition reaction.
Scheme 37. Intermolecular C-C addition reaction.
Molecules 17 14700 sch037
Scheme 38. Addition reactions onto carbenes.
Scheme 38. Addition reactions onto carbenes.
Molecules 17 14700 sch038
Scheme 39. 3-exo and 4-exo cyclization with aldehydes as radical traps.
Scheme 39. 3-exo and 4-exo cyclization with aldehydes as radical traps.
Molecules 17 14700 sch039
Scheme 40. 4-exo-cyclization of unsaturated epoxides.
Scheme 40. 4-exo-cyclization of unsaturated epoxides.
Molecules 17 14700 sch040
Figure 3. Possible radical intermediate in 4-exo-cyclization of unsaturated epoxides.
Figure 3. Possible radical intermediate in 4-exo-cyclization of unsaturated epoxides.
Molecules 17 14700 g003
Scheme 41. Catalytic epoxypolyene cyclization.
Scheme 41. Catalytic epoxypolyene cyclization.
Molecules 17 14700 sch041
Scheme 42. Tandem cyclization approach for meroterpenoid (+)-zonarol synthesis.
Scheme 42. Tandem cyclization approach for meroterpenoid (+)-zonarol synthesis.
Molecules 17 14700 sch042
Scheme 43. 7-endo cyclization.
Scheme 43. 7-endo cyclization.
Molecules 17 14700 sch043
Scheme 44. RRO of epoxides for the initiation of the living polymerization of e-caprolactone.
Scheme 44. RRO of epoxides for the initiation of the living polymerization of e-caprolactone.
Molecules 17 14700 sch044
Scheme 45. Polymerization of EPOX mediated by titanocene.
Scheme 45. Polymerization of EPOX mediated by titanocene.
Molecules 17 14700 sch045

Share and Cite

MDPI and ACS Style

Rossi, B.; Prosperini, S.; Pastori, N.; Clerici, A.; Punta, C. New Advances in Titanium-Mediated Free Radical Reactions. Molecules 2012, 17, 14700-14732. https://doi.org/10.3390/molecules171214700

AMA Style

Rossi B, Prosperini S, Pastori N, Clerici A, Punta C. New Advances in Titanium-Mediated Free Radical Reactions. Molecules. 2012; 17(12):14700-14732. https://doi.org/10.3390/molecules171214700

Chicago/Turabian Style

Rossi, Bianca, Simona Prosperini, Nadia Pastori, Angelo Clerici, and Carlo Punta. 2012. "New Advances in Titanium-Mediated Free Radical Reactions" Molecules 17, no. 12: 14700-14732. https://doi.org/10.3390/molecules171214700

Article Metrics

Back to TopTop