Next Article in Journal
Hawthorn Drying: An Exploration of Ultrasound Treatment and Microwave–Hot Air Drying
Previous Article in Journal
The Beneficial Effect of Salicornia herbacea Extract and Isorhamnetin-3-O-glucoside on Obesity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Cp2TiCl/H2O as a Sustainable System for the Reduction of Organic Functional Groups: Potential Application of Cp2TiCl/D2O to the Analysis of Bioactive Phenols in Olive Oil

by
Antonio Rosales Martínez
1,*,
Juan F. García-Martín
2 and
Ignacio Rodríguez-García
3
1
Department of Chemical Engineering, Polytechnic School, University of Seville, E-41011 Sevilla, Spain
2
Departamento de Ingeniería Química, Facultad de Química, Universidad de Sevilla, E-41012 Sevilla, Spain
3
Department of Chemistry and Physics, Research Institute CIAIMBITAL, University of Almería, E-04120 Almería, Spain
*
Author to whom correspondence should be addressed.
Processes 2023, 11(4), 979; https://doi.org/10.3390/pr11040979
Submission received: 13 February 2023 / Revised: 13 March 2023 / Accepted: 15 March 2023 / Published: 23 March 2023
(This article belongs to the Section Food Process Engineering)

Abstract

:
Significant efforts have been made toward developing sustainable reduction reactions of organic and bioorganic compounds. In these studies, the selection of reagents and solvents has played a very important role in the development of environment-friendly methodologies. In this context, the reducing agent Cp2TiCl/H2O has been introduced as a safe, efficient, selective, and low-cost reagent, and thus as a sustainable alternative for the reduction of organic compounds. To facilitate understanding of the reductions mediated by this system, in this study we focus on describing the intermediates, mechanisms, and representative examples. Finally, a reflection is made on the future perspectives of this reducing agent, including its analog Cp2TiCl/D2O as a powerful tool for the preparation of deuterated phenols, which can be successfully used as an internal standard for analyzing bioactive phenols in olive oil.

1. Introduction

The reduction of functional groups in organic compounds is of central importance in biochemistry and organic chemistry. This reaction can be achieved through several methods, including hydride transfer reactions, catalytic hydrogenations, dissolving metal reductions, and organometallic complexes [1]. Although the reduction reaction has been known about for a long time, this reaction is a central focus of modern organic synthesis, with significant efforts directed at developing more sustainable reductions [2,3] in accordance with the principles of green chemistry [4]. In this context, this manuscript summarizes the use of Cp2TiCl/water as a green system for the reduction of organic functional groups and its analog Cp2TiCl/D2O as an efficient system for the synthesis of deuterated compounds, although it is not intended to be a comprehensive review. This system is constituted by Cp2TiCl, a single-electron transfer (SET) reagent, which has been proposed as a new green reagent that can carry out an important variety of chemical reactions such as C–C and C–O bond-forming reactions, as well as reduction, isomerization, deoxygenation, and polymerization reactions, under mild reaction conditions and with high diastereo- and regioselectivities [5,6,7,8,9,10,11]. This SET reagent fulfills some of the twelve principles of green chemistry, such as catalysis, safer solvents, waste minimization, atom economy, toxicity, and energy efficiency [5,6]. The other reagent is water or heavy water, an environmentally benign solvent because it is nonflammable, abundant, and nontoxic. Furthermore, the chemical nature of water leads to remarkable new reactions that cannot be achieved otherwise [12]. The most relevant mechanisms of the reduction reactions mediated by the Cp2TiCl/H2O system will be discussed. Of note is that the analog of this system, the one using heavy water as a deuterium atom source, is an ideal reagent for the efficient synthesis of deuterated compounds, which present chemical and physical properties virtually identical to those of their non-deuterated analogs. These deuterated compounds have many potential applications [13], especially in the food and pharmaceutical industries. For example, in the food industry, deuterated compounds can be used as an internal standard [14,15] in the analysis of bioactive compounds. In this context, compared to other deuteration methodologies [16,17], the Cp2TiCl/D2O system [18] is a sustainable and inexpensive reagent, very useful for the preparation of deuterated phenols that can be used as an internal standard in the analysis of some of the bioactive phenolic compounds present in olive oil [15]. In fact, the antioxidant properties of olive oil have been attributed to some of these bioactive phenols, such as tyrosol, and hydroxytyrosol and its derivatives (e.g., oleuropein) [19]. For this reason, the development of analytical methods using deuterated systems as internal standards which allow the quantification of bioactive phenols is an important goal in food and analytical chemistry. Finally, the future perspectives of this system are discussed.

2. Discussion

The use of Cp2TiCl in reduction reactions of organic compounds using water as a hydrogen atom donor has been known for more than a decade [15,16]. Cp2TiCl is a single electron transfer reagent obtained by reducing Cp2TiCl2 with a nonhazardous metal in dust form, such as Zn or Mn [20]. Alternatively, photoredox catalysis [21], electrochemical reduction [22], and organosilicon reducing agents [23] can be used to obtain Cp2TiCl from Cp2TiCl2. In tetrahydrofuran (THF) or toluene, the Cp2TiCl solution is lime green. However, it turns dark blue when water is added to the solution [24,25].
The presence of a vacant site and an unpaired d-electron in the inner sphere of this complex allows its coordination to functional groups with heteroatoms with free valence electrons, obtaining different intermediates if the process is energetically favorable. From these intermediates, important chemical synthesis reactions take place, such as C–C and C–O bond-forming reactions, as well as reduction, isomerization, deoxygenation, protection, epoxidation, and decyanation reactions (Scheme 1) [5,6,7,8,9,10,11].
Considering the study of reduction reactions mediated by the Cp2TiCl/H2O system, three types of primary intermediates are formed from epoxides, ozonides, carbonyl compounds, activated halides, and transition metals. Initially, the formation of carbon-centered radicals 1 was observed when C–Y is coordinated to Cp2TiCl after a single-electron transfer (pathway A, Scheme 2). If a carbon radical 1 is trapped by a second molecule of titanocene(III), an alkyl-TiIV intermediate 2 is formed (pathway A1, Scheme 2). However, if the carbon radical 1 is hindered, it can be reduced by hydrogen atom transfer from the aqueous complex of titanocene(III) (pathway A2, Scheme 2). The titanaoxirane intermediates 3 are generated from carbonyl groups and Cp2TiCl/Mn (pathway B, Scheme 2). Hydrolysis of these intermediates 3 generates the reduced compounds. Finally, a dihydro metal complex 4 is obtained when the Cp2TiCl/water system reacts with a transition metal such as Pd or Rh, allowing the reduction of alkenes or alkynes (pathway C, Scheme 2). All these reaction species are susceptible to producing reductive products through the mechanisms described below.
A breakthrough in the use of the Cp2TiCl/H2O system as a new reducing reagent was the development of a catalytic cycle under reductive conditions reported by Gansäuer et al. [26]. In this catalytic cycle, hydrochloride-substituted pyridines (collidine hydrochloride) were employed as versatile acids to protonate the Ti–C or the Ti–O bonds to generate Cp2TiCl2 (Scheme 3) [26]. This catalytic cycle has been used successfully for different target functional groups and uses water as a hydrogen source.
The mechanism involved in the reduction of these intermediates and some examples are discussed below.

2.1. Carbon-Centered Radical Intermediates

Homolysis of one epoxide C–O bond, one ozonide O–O bond, or C–halogen bond occurs by inner-sphere electron transfer and results in the generation of a carbon-centered radical intermediate (pathway A, Scheme 2). Under aqueous conditions, the carbon-centered radical intermediate 1 can be reduced through hydrogen atom transfer (HAT) (pathway A, Scheme 4) from an aqua complex of titanocene(III) or by hydrolysis of an organometallic alkyl-TiIV intermediate 2 generated when the radical is trapped by a second species of titanocene(III) (pathway B, Scheme 4).
The HAT pathway is only possible because the activation of water by Cp2TiCl takes place. This activation of water for HAT was initially reported by us in the radical cyclization of epoxygermacrolides [27]. In that paper, we published the following comment: “Contrary to general belief, water can act as a reactive hydrogen atom donor in radical chemistry mediated by Ti(III) species”. The propensity of water to act as a HAT reagent could be explained by a lowering of the dissociation energy of the O–H bond of almost 60 kcal/mol in the presence of Cp2TiCl. Initially, aqua complex 5 was proposed as a HAT reagent (pathway A, Scheme 5) [17]. However, after extending and refining the theoretical calculations, studies of cyclic voltammetry, mass spectrometric analysis, and electro-paramagnetic resonance techniques suggested that aqua complex 6 is the active HAT reagent (pathway B, Scheme 5) [25,28].
Water as a HAT reagent was rigorously demonstrated in the transannular cyclization of caryophyllene oxide 7 (Scheme 6) [24]. The radical opening of epoxide 7 mediated by Cp2TiCl generates a hindered primary radical 8. Several experiments were carried out to determine the reduction mechanism of the radical generated. In this way, treating 7 with titanocene(III) in the presence of 1,4-CHD (a hydrogen atom donor) and subsequent addition of D2O (after consumption of 7 (1 h)) gave a mixture of 9 and 10 (pathway A, Scheme 6). The lack of deuterium labeling in 9 was a result that allowed us to indicate that the reduction of the hindered radical as 8 took place by the transfer of hydrogen atoms from 1,4-CHD and not by the hydrolysis of the organometallic intermediate which may be generated when the radical is trapped by a second molecule of Cp2TiCl. In another experiment, epoxide 7 was treated with Cp2TiCl in the presence of D2O and 1,4-CHD. Under these conditions, the main isolated product was 11 (pathway B, Scheme 6). The presence of deuterium labeling in 11 indicated that deuterium atom transfer from D2O was faster than hydrogen atom transfer from 1,4-CHD, which in the previous experiment (pathway A, Scheme 6) was observed to be much faster than the formation of an organometallic alkyl-TiIV intermediate.
This mechanism of reduction through HAT was also observed with propargyl radicals [29]. Based on deuteration studies, propargyl radicals were reported to be effectively reduced by HAT from H2O in a process mediated by Cp2TiCl.
As can be observed, due to their mild experimental conditions and high chemoselectivity, HAT reactions using the Cp2TiCl/H2O system represent an excellent methodology for reducing carbon-centered radicals of diverse nature. A deep theoretical and experimental study [30] indicated that the success of this reaction is based on two key features: (a) an excellent binding capacity of H2O to Cp2TiCl and (b) a low activation energy for the HAT step. If the activation energy for the HAT step is high, the dimerization of the radicals usually prevails over their reduction.
Alternatively, the carbon-centered radical intermediate 1 can be reduced by a more conventional mechanism that involves the formation of an organometallic alkyl-TiIV intermediate 2 and subsequent hydrolysis by water (pathway B, Scheme 4). This other mechanism was demonstrated based on the experimental and theoretical observations we obtained in the radical opening of ozonides [31]. When ozonide 12 is treated with Cp2TiCl, an unhindered primary radical 13 is formed. Several reactions were carried out to determine the mechanism of reduction of this benzylic radical 13. When ozonide 12 was treated with titanocene(III) and water, a mixture of homocoupling product 14 (62%) and reductive product 15 (38%) was obtained (pathway A, Scheme 7) [32]. This result is reported in Supplementary Materials. In another experiment, ozonide 12 was treated with Cp2TiCl and, after consumption of 12 (1 h), D2O was added, observing a mixture of 14 and deuterium-labeled product 16 (pathway B, Scheme 7) [31]. The 45% deuterium incorporation in 16 together with the same relative ratio of homocoupling and reductive products in both experiments indicated that the reduction of the unhindered primary radical could be due to the hydrolysis of an organometallic alkyl-TiIV intermediate formed by trapping the radical with a second molecule of titanocene(III).
To decrease the dimerization of radical 13, Cp2TiCl was added dropwise to ozonide 12 in dry THF, and finally hydrolysis with HCl (2N) resulted in 15 as the main product [32] (see the example in Scheme 8).
Although more computational and experimental studies are being conducted [32] to verify the proposed mechanism of reduction of carbon radicals, as a general rule it can be observed that hindered and tertiary carbon radicals are normally reduced by HAT from water in a process promoted by Cp2TiCl, while unhindered and primary radicals are generally reduced via hydrolysis of an organometallic alkyl-TiIV intermediate. Several examples of the reduction of epoxides and ozonides by the Cp2TiCl/H2O system are illustrated in Scheme 8. As can be observed in Scheme 8, the regioselective opening of epoxides mediated by Cp2TiCl, where water is used as the donor of hydrogen atoms, is a highly green procedure for obtaining anti-Markovnikov alcohols [26,33].

2.2. Titanaoxirane Intermediates

The Cp2TiCl/H2O system can also be used to reduce titanaoxirane species 3 (pathway B, Scheme 2) obtained when carbonyl groups are treated with Cp2TiCl/Mn. Rosales Martínez et al. [34] explored the reduction mechanism of the carbonyl group with titanocene(III) in an aqueous medium. After treating acetophenone 17 with Cp2TiCl in the absence of manganese, the starting material 17 was recovered unchanged (pathway A, Scheme 9). However, the same reaction in the presence of manganese gave alcohol 18 with a yield of 94% (pathway B, Scheme 9). These results indicated that manganese is required not only for reducing Cp2TiCl2 to Cp2TiCl, but also for generating the intermediate titanaoxirane 20 as the key intermediate (Scheme 10a).
The observations mentioned were explained by the mechanism proposed in Scheme 10a. Coordination between acetophenone 17 and titanocene(III) would provide the reaction intermediate 19 in an equilibrium reaction shifted toward titanocene(III) and acetophenone 17. This equilibrium was proposed because, in the absence of manganese, the starting material 17 was recovered unchanged. Therefore, the manganese present in the reaction medium could reduce the Cp2TiCl coordinated to acetophenone 17 through an irreversible process leading to the formation of the titanaoxirane intermediate 20. When water is present in the reaction medium, the aqua complex 6 is formed, which would probably be more acidic than the non-coordinated water and, therefore, could promote the hydrolysis of the intermediate 20 to an alkyl-TiIV intermediate such as 21, a precursor of alcohol 18. Several examples of the selective reduction of ketones by Cp2TiCl/H2O are presented in Scheme 10b.

2.3. Dihydro Metal Complex Intermediates

In the presence of transition metals such as Rh or Pd, the Cp2TiCl/H2O system was also used to reduce alkynes and alkenes [36]. The mechanism of this reduction was explained by the formation of the aqua complex 6, which is able to facilitate the transfer of hydrogen atoms from water to late transition metals (Pd or Rh) to give metal hydride intermediates 4. These intermediates could subsequently bring about alkyne (and alkene) hydrogenation as depicted in Scheme 11a. This mechanism was supported by theoretical and experimental evidence. In this way, theoretical calculations suggested an activation energy of only 17.3 kcal/mol for the hydrogen atom transfer. Several examples of the reduction of alkynes/alkenes by the Cp2TiCl/H2O system are presented in Scheme 11b.
All these results show that the Cp2TiCl/H2O system is an excellent sustainable reagent for reduction reactions. This system avoids the use of traditional hydrogen atom donors, such as trialkyl stannane hydrides, which are toxic and sometimes produce side products. Moreover, the moderate efficiency of HAT from this system compared with the trialkylstannanes results in chemical advantages in the involved radical chain reactions, since the radical species will not be prematurely reduced. Moreover, this system avoids the use of cyclohexadiene, silanes, and related compounds as HAT reagents, which are expensive, toxic, and/or foul-smelling. Another attractive feature of this system is that it can be used with a catalytic amount of Cp2TiCl, unlike other novel HAT reagents such as borane complex with N-heterocyclic carbenes since these reagents are employed in stoichiometric amounts.
Finally, the analog of this system that uses heavy water as a deuterium source is used to generate complex deuterium-labeled compounds [16], which can be applied as an internal standard in food analysis. The first synthesis of a deuterated sample of tyrosol, using the Cp2TiCl/D2O system as a reducing reagent, was used as an internal standard in determining the concentration of bioactive tyrosol in an olive leaf extract [15]. The result given in mg tyrosol/g of dry weight olive leaves was 0.2 ± 0.05 mg/g.
Although the overall yield of deuterated tyrosol (22) was not high (10.5%), this compound was synthesized from commercially available acetoxystyrene (23) in only three steps: (a) epoxidation of 23 with m-chloroperoxybenzoic acid (mCPBA) to give 24; (b) radical opening of epoxide 24 with the Ti(III)/D2O system to produce deuterated alcohol 25; and (c) deacetylation of 25 to form 22 (Scheme 12). The moderate deuterium incorporation obtained in alcohol 25 may be due to the fact that part of the alkyl-TiIV intermediate generated after the opening of the epoxide 24 with titanocene(III) is not hydrolyzed with the heavy water present in the reaction medium due to a high isotope effect. Part of this intermediate could be hydrolyzed after working up the reaction with H3O+.

3. Conclusions and Perspectives

In summary, we have presented an article showing that Cp2TiCl/H2O is a novel hydrogen atom donor system. It is characterized by being a sustainable, efficient, selective, and economic system that allows the reduction of ozonides, epoxides, carbonyl compounds, alkenes, and alkynes to be carried out. These reductions occur under mild and environmentally safe reaction conditions avoiding the use of other toxic and relatively expensive hydrogen atom donors, such as 1,4-cyclohexadiene or Bu3SnH. Moreover, its analog using heavy water has been postulated as a much cheaper and safer alternative compared to the traditional reducing reagents used to obtain deuterated alcohols such as NaBD4 and LiAlH4. Despite recent advances in the use of this system as a reducing reagent in radical and organometallic chemistry, many challenges remain: (1) A wide range of other functional groups needs to be further explored with Cp2TiCl. (2) Asymmetric reduction is another issue due to two factors: the need to use chiral titanocene(III) reagents [37] and that the key intermediate in the reduction process must be organometallic and non-radical in nature. (3) Developing and optimizing the solvent-free reduction process. (4) Extrapolating this system to industrial chemical reductions. (5) Using the Cp2TiCl/D2O system as a deuterium source to generate more deuterated compounds of interest in the food industry, such as deuterated isomers of hydroxytyrosol, oleacein, oleocanthal, and oleuropein, some of them of major importance in the olive oil industry. These deuterated compounds can be used as internal standards to quantify their non-deuterated analogs present in olive oil. For this purpose, high-performance liquid chromatography (HPLC) allied with mass spectrometry detection, using electrospray ionization as an interface can be used [15].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr11040979/s1, Figure S1: Experimental procedures for reference [25]; Figure S2: 1H NMR (CDCl3, 300 MHz) of the reaction of the crude ozonide 12 with Cp2TiCl2 and Mn in THF, showing the relative ratio of products 14 and 15; Figure S3: 1H NMR (CDCl3, 300 MHz) of 14; Figure S4: DEPT 135 and 13C NMR (CDCl3, 75 MHz) of 14; Figure S5: 1H NMR (CDCl3, 300 MHz) of 15; Figure S6: DEPT 135 and 13C NMR (CDCl3, 75 MHz) of 15.

Author Contributions

A.R.M.: design and coordination of the project, writing—original draft, and writing—review and editing. J.F.G.-M.: review. I.R.-G.: writing—review. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the University of Seville, through the Vicerrectorado de Investigación (Projects 2020/00001014 and 2021/00000422: Ayudas a Consolidación de Grupos de la Junta de Andalucía and Project Politec-Biomat: Red de Biomateriales en la Universidad de Sevilla), the University of Almería, Junta de Andalucía (Conserjería de Transformación Económica, Industria, Conocimiento y Universidades), Fondo Europeo de Desarrollo Regional (FEDER) for the projects UALFEDER 2020-FQM-B1989, PY20_01027, and CEIA3 PYC20 RE 060 UAL, and the Horizon 2020—Research and Innovation Framework Program of the European Commission for the project 101022507 LAURELIN.

Acknowledgments

Antonio Rosales Martínez and Juan F. García Martín acknowledge the University of Seville for their positions as professors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Smith, M.B.; March, J. March’s Advanced Organic Chemistry: Reactions, Mechanisms, and Structure, 6th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2007; pp. 1786–1869. [Google Scholar]
  2. Sheldon, R.A. Green solvents for sustainable organic synthesis: State of the art. Green Chem. 2005, 7, 267–278. [Google Scholar] [CrossRef]
  3. Álvarez de Cienfuegos, L.; Robles, R.; Miguel, D.; Justicia, J.; Cuerva, J.M. Reduction reactions in green solvents: Water, supercritical carbon dioxide, and ionic liquids. ChemSusChem 2011, 4, 1035–1048. [Google Scholar] [CrossRef] [PubMed]
  4. Anastas, P.T.; Warner, J.C. Green Chemistry: Theory and Practice; Oxford University Press: Oxford, UK, 2000; p. 152. [Google Scholar]
  5. Castro Rodríguez, M.; Rodríguez García, I.; Rodríguez Maecker, R.N.; Pozo Morales, L.; Oltra, J.E.; Rosales Martínez, A. Cp2TiCl: An Ideal Reagent for Green Chemistry? Org. Process Res. Dev. 2017, 21, 911–923. [Google Scholar] [CrossRef]
  6. Rosales Martínez, A.; Castro Rodriguez, M.; Rodríguez-García, I.; Pozo Morales, L.; Rodriguez Maecker, R.N. Titanocene dichloride: A new green reagent in organic chemistry. Chin. J. Catal. 2017, 38, 1659–1663. [Google Scholar] [CrossRef]
  7. Rosales Martínez, A.; Pozo Morales, L.; Díaz Ojeda, E.; Castro Rodríguez, M.; Rodríguez-García, I. The Proven Versatility of Cp2TiCl. J. Org. Chem. 2021, 86, 1311–1329. [Google Scholar] [CrossRef]
  8. Gansäuer, A. From enantioselective to regiodivergent epoxide opening and radical arylation—useful or just interesting? Synlett 2021, 32, 447–456. [Google Scholar] [CrossRef]
  9. Barrero, A.F.; Quílez del Moral, J.F.; Sánchez, E.M.; Arteaga, J.F. Titanocene-mediated radical cyclization: An emergent method towards the synthesis of natural products. Eur. J. Org. Chem. 2006, 1627–1641. [Google Scholar] [CrossRef]
  10. Gansäuer, A.; Justicia, J.; Fan, C.-A.; Worgull, D.; Piestert, F. Reductive C–C Bond Formation after Epoxide Opening via Electron Transfer. Top. Curr. Chem. 2007, 279, 25–52. [Google Scholar]
  11. Botubol-Ares, J.M.; Durán-Peña, M.J.; Hanson, J.R.; Hernández-Galán, R.; Collado, I.G. Cp2Ti(III)Cl and analogues as sustainable templates in organic synthesis. Synthesis 2018, 50, 2163–2180. [Google Scholar]
  12. Breslow, R. The principles of and reasons for using water as a solvent for green chemistry. In Handbook of Green Chemistry; Chao-Jun, L., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2010; Volume 5, pp. 1–29. [Google Scholar]
  13. Liu, J.; Liu, X. Deuteride Materials; Springer Nature Singapore Pte Ltd.: Singapore, 2019; pp. 231–284. [Google Scholar]
  14. Stokvis, E.; Rosing, H.; Beijnen Jos, H. Stable isotopically labeled internal standards in quantitative bioanalysis using liquid chromatography/mass spectrometry: Necessity or not? Rapid Commun. Mass Spectrom. 2005, 19, 401–407. [Google Scholar] [CrossRef]
  15. 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.; et al. 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, 2010, 4288–4295. [Google Scholar] [CrossRef]
  16. Prakash, G.; Paul, N.; Oliver, G.A.; Werz, D.B.; Maiti, D. C–H deuteration of organic compounds and potential drug candidates. Chem. Soc. Rev. 2022, 51, 3123–3163. [Google Scholar] [CrossRef]
  17. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M. C-H Functionalisation for hydrogen isotope exchange. Angew. Chem. Int. Ed. 2018, 57, 3022–3047. [Google Scholar] [CrossRef]
  18. Rosales, A.; Rodríguez-García, I. Cp2TiCl/D2O/Mn, a formidable reagent for the deuteration of organic compounds. Beilstein J. Org. Chem. 2016, 12, 1585–1589. [Google Scholar] [CrossRef] [Green Version]
  19. Marković, A.K.; Tori´c, J.; Barbarić, M.; Brala, C.J. Hydroxytyrosol, tyrosol and derivatives and their potential effects on human health. Molecules 2019, 24, 2001. [Google Scholar] [CrossRef] [Green Version]
  20. Green, M.L.H.; Lucas, C.R. Some d1 Bis-π-cyclopentadienyl titanium complexes with nitrogen or phosphorus ligands. J. Chem. Soc. Dalton Trans. 1972, 1000–1003. [Google Scholar] [CrossRef]
  21. Zhang, Z.; Richrath, R.B.; Gansäeuer, A. Merging catalysis in single electron steps with photoredox catalysis-efficient and sustainable radical chemistry. ACS Catal. 2019, 9, 3208–3212. [Google Scholar] [CrossRef]
  22. Liedtke, T.; Hilche, T.; Klare, S.; Gansäeuer, A. Condition screening for sustainable catalysis in single-electron steps by cyclic voltammetry: Additives and solvents. ChemSusChem 2019, 12, 3166–3171. [Google Scholar] [CrossRef]
  23. Saito, T.; Nishiyama, H.; Tanahashi, H.; Kawakita, K.; Tsurugi, H.; Mashima, K. 1,4-Bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadienes as Strong Salt-Free Reductants for Generating Low-Valent Early Transition Metals with Electron-Donating Ligands. J. Am. Chem. Soc. 2014, 136, 5161–5170. [Google Scholar] [CrossRef]
  24. Cuerva, J.M.; Campaña, A.G.; Justicia, J.; Rosales, A.; Oller-Lopez, J.L.; Robles, R.; Cárdenas, D.J.; Buñuel, E.; Oltra, J.E. Water: The ideal hydrogen-atom source in free-radical chemistry mediated by TiIII and other single-electron-transfer metals? Angew. Chem. Int. Ed. 2006, 45, 5522–5526. [Google Scholar] [CrossRef]
  25. Rosales Martínez, A.; Enríquez, L.; Jaraíz, M.; Thiessen, T.; Sidhu, J.; McIndoe, J.S.; Rodríguez-García, I. Understanding the color change of the solutions of Cp2TiCl upon addition of water. Appl. Organomet. Chem. 2023, 37, e6979. [Google Scholar]
  26. Gansäeuer, A.; Pierobon, M.; Bluhm, H. Catalytic, highly regio- and chemoselective generation of radicals from epoxides: Titanocene dichloride as an electron transfer catalyst in transition metal catalyzed radical reactions. Angew. Chem. Int. Ed. 1998, 37, 101–103. [Google Scholar] [CrossRef]
  27. 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]
  28. Gansäuer, A.; Behlendorf, M.; Cangoenuel, A.; Kube, C.; Cuerva, J.M.; Friedrich, J.; van Gastel, M. H2O Activation for Hydrogen-Atom Transfer: Correct Structures and Revised Mechanisms. Angew. Chem. Int. Ed. 2012, 51, 3266–3270. [Google Scholar] [CrossRef] [PubMed]
  29. Muñoz-Bascón, J.; Sancho-Sanz, I.; Álvarez-Manzaneda, E.; Rosales, A.; Oltra, J.E. Highly selective Barbier-type propargylations and allenylations catalyzed by titanocene(III). Chem. Eur. J. 2012, 18, 14479–14486. [Google Scholar] [CrossRef]
  30. 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]
  31. Rosales, A.; Muñoz-Bascón, J.; Lopez-Sanchez, C.; Alvarez-Corral, M.; Muñoz-Dorado, M.; Rodriguez-Garcia, 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]
  32. Previously unpublished results. Experimental Procedures and NMR Spectra are Reported in Supporting Information.
  33. Rosales Martínez, A.; Rodríguez-García, I.; López-Martínez, J.L. Green reductive regioselective opening of epoxides: A green chemistry laboratory experiment. J. Chem. Educ. 2022, 99, 2710–2714. [Google Scholar] [CrossRef]
  34. Rosales, A.; Muñoz-Bascón, J.; Roldan-Molina, E.; Castaneda, M.A.; Padial, N.M.; Gansauer, A.; Rodríguez-García, I.; Oltra, J.E. Selective reduction of aromatic ketones in aqueous medium mediated by Ti(III)/Mn: A revised mechanism. J. Org. Chem. 2014, 79, 7672–7676. [Google Scholar] [CrossRef]
  35. Barrero, A.F.; Rosales, A.; Cuerva, J.M.; Gansaeuer, 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]
  36. Campaña, A.G.; Estevez, R.E.; Fuentes, N.; Robles, R.; Cuerva, J.M.; Buñuel, E.; Cardenas, D.; Oltra, J.E. Unprecedented hydrogen transfer from water to alkenes and alkynes mediated by TiIII and late transition metals. Org. Lett. 2007, 9, 2195–2198. [Google Scholar] [CrossRef]
  37. 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]
Scheme 1. Target functional groups of titanocene(III).
Scheme 1. Target functional groups of titanocene(III).
Processes 11 00979 sch001
Scheme 2. Intermediates generated in the reduction of functional groups with the Cp2TiCl/H2O system.
Scheme 2. Intermediates generated in the reduction of functional groups with the Cp2TiCl/H2O system.
Processes 11 00979 sch002
Scheme 3. Catalytic Cp2TiCl epoxide opening using 2,4,6-collidine as a regenerator and 1,4-cyclohexadiene (1,4-CHD) as a hydrogen source.
Scheme 3. Catalytic Cp2TiCl epoxide opening using 2,4,6-collidine as a regenerator and 1,4-cyclohexadiene (1,4-CHD) as a hydrogen source.
Processes 11 00979 sch003
Scheme 4. Reduction of carbon-centered radical intermediates. Pathway A: HAT from water to free radicals mediated by an aqua complex of titanocene(III). Pathway B: Hydrolysis of an alkyl-TiIV intermediate.
Scheme 4. Reduction of carbon-centered radical intermediates. Pathway A: HAT from water to free radicals mediated by an aqua complex of titanocene(III). Pathway B: Hydrolysis of an alkyl-TiIV intermediate.
Processes 11 00979 sch004
Scheme 5. Proposed Cp2TiCl aqua complexes.
Scheme 5. Proposed Cp2TiCl aqua complexes.
Processes 11 00979 sch005
Scheme 6. Deuteration experiments in the cyclization of 7. (a) Addition of D2O after 1h of reaction; (b) addition of D2O from the beginning.
Scheme 6. Deuteration experiments in the cyclization of 7. (a) Addition of D2O after 1h of reaction; (b) addition of D2O from the beginning.
Processes 11 00979 sch006
Scheme 7. Experiments in the radical opening of ozonide 12. (a) Reaction in the absence of deuterium source [32]; (b) Reaction quenched with D2O [31].
Scheme 7. Experiments in the radical opening of ozonide 12. (a) Reaction in the absence of deuterium source [32]; (b) Reaction quenched with D2O [31].
Processes 11 00979 sch007
Scheme 8. Examples of reductions of epoxides and ozonides with the Cp2TiCl/H2O system [24,27,30,32,33]. a Substoichiometric amount of Cp2TiCl. b Stoichiometric amount of Cp2TiCl.
Scheme 8. Examples of reductions of epoxides and ozonides with the Cp2TiCl/H2O system [24,27,30,32,33]. a Substoichiometric amount of Cp2TiCl. b Stoichiometric amount of Cp2TiCl.
Processes 11 00979 sch008
Scheme 9. Experiments on the reduction of acetophenone 17 with the Cp2TiCl/H2O system.
Scheme 9. Experiments on the reduction of acetophenone 17 with the Cp2TiCl/H2O system.
Processes 11 00979 sch009
Scheme 10. (a) Proposed mechanism for the titanocene(III)/Mn-promoted reduction of acetophenone 17 in an aqueous medium [34]. (b) Examples of selective reduction of ketones by Cp2TiCl/H2O [19,35]. a Substoichiometric amount of Cp2TiCl. b Stoichiometric amount of Cp2TiCl.
Scheme 10. (a) Proposed mechanism for the titanocene(III)/Mn-promoted reduction of acetophenone 17 in an aqueous medium [34]. (b) Examples of selective reduction of ketones by Cp2TiCl/H2O [19,35]. a Substoichiometric amount of Cp2TiCl. b Stoichiometric amount of Cp2TiCl.
Processes 11 00979 sch010
Scheme 11. (a) Proposed mechanism for the hydrogenation of alkenes or alkynes with Cp2TiCl/H2O in the presence of Pd or Rh catalysts [36]. (b) Examples of the reduction of alkenes and alkynes. a Stoichiometric amount of Cp2TiCl.
Scheme 11. (a) Proposed mechanism for the hydrogenation of alkenes or alkynes with Cp2TiCl/H2O in the presence of Pd or Rh catalysts [36]. (b) Examples of the reduction of alkenes and alkynes. a Stoichiometric amount of Cp2TiCl.
Processes 11 00979 sch011
Scheme 12. Synthesis of deuterated tyrosol (22).
Scheme 12. Synthesis of deuterated tyrosol (22).
Processes 11 00979 sch012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rosales Martínez, A.; García-Martín, J.F.; Rodríguez-García, I. Cp2TiCl/H2O as a Sustainable System for the Reduction of Organic Functional Groups: Potential Application of Cp2TiCl/D2O to the Analysis of Bioactive Phenols in Olive Oil. Processes 2023, 11, 979. https://doi.org/10.3390/pr11040979

AMA Style

Rosales Martínez A, García-Martín JF, Rodríguez-García I. Cp2TiCl/H2O as a Sustainable System for the Reduction of Organic Functional Groups: Potential Application of Cp2TiCl/D2O to the Analysis of Bioactive Phenols in Olive Oil. Processes. 2023; 11(4):979. https://doi.org/10.3390/pr11040979

Chicago/Turabian Style

Rosales Martínez, Antonio, Juan F. García-Martín, and Ignacio Rodríguez-García. 2023. "Cp2TiCl/H2O as a Sustainable System for the Reduction of Organic Functional Groups: Potential Application of Cp2TiCl/D2O to the Analysis of Bioactive Phenols in Olive Oil" Processes 11, no. 4: 979. https://doi.org/10.3390/pr11040979

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