Next Article in Journal
Advanced Treatment of Phosphorus Pesticide Wastewater Using an Integrated Process of Coagulation and Ozone Catalytic Oxidation
Next Article in Special Issue
Efficient Synthesis of 3-Sulfonyl-2-sulfonylmethyl-2H-chromenes via Tandem Knoevenagel Condensation/Oxa-Michael Addition Protocol
Previous Article in Journal
Thermal Stability of Potassium-Promoted Cobalt Molybdenum Nitride Catalysts for Ammonia Synthesis
Previous Article in Special Issue
Chiral Ionic Liquids Based on l-Cysteine Derivatives for Asymmetric Aldol Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Horizons in Asymmetric Organocatalysis: En Route to the Sustainability and New Applications

by
Sandra Ardevines
,
Eugenia Marqués-López
* and
Raquel P. Herrera
*
Laboratorio de Organocatálisis Asimétrica, Departamento de Química Orgánica, Instituto de Síntesis Química y Catálisis Homogénea (ISQCH), CSIC-Universidad de Zaragoza, C/ Pedro Cerbuna 12, 50009 Zaragoza, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(1), 101; https://doi.org/10.3390/catal12010101
Submission received: 24 December 2021 / Revised: 12 January 2022 / Accepted: 13 January 2022 / Published: 16 January 2022
(This article belongs to the Special Issue Organocatalysis: Advances, Opportunity, and Challenges)

Abstract

:
Nowadays, the development of new enantioselective processes is highly relevant in chemistry due to the relevance of chiral compounds in biomedicine (mainly drugs) and in other fields, such as agrochemistry, animal feed, and flavorings. Among them, organocatalytic methods have become an efficient and sustainable alternative since List and MacMillan pioneering contributions were published in 2000. These works established the term asymmetric organocatalysis to label this area of research, which has grown exponentially over the last two decades. Since then, the scientific community has attended to the discovery of a plethora of organic reactions and transformations carried out with excellent results in terms of both reactivity and enantioselectivity. Looking back to earlier times, we can find in the literature a few examples where small organic molecules and some natural products could act as effective catalysts. However, with the birth of this type of catalysis, new chemical architectures based on amines, thioureas, squaramides, cinchona alkaloids, quaternary ammonium salts, carbenes, guanidines and phosphoric acids, among many others, have been developed. These organocatalysts have provided a broad range of activation modes that allow privileged interactions between catalysts and substrates for the preparation of compounds with high added value in an enantioselective way. Here, we briefly cover the history of this chemistry, from our point of view, including our beginnings, how the field has evolved during these years of research, and the road ahead.

1. Introduction

More than twenty years ago, a new Big Bang occurred in organic chemistry [1]. In 2000, we witnessed the conceptualization of a new field, asymmetric organocatalysis, even though the first asymmetric organocatalytic reaction appeared in 1904 [2], when this discipline had not yet emerged. It was only at the beginning of the 21st century with the parallel publications of List [3] and MacMillan [4], when a stream of work focused on this amazing area of research was triggered (Scheme 1). In fact, these researchers have recently been awarded the 2021 Nobel Prize in Chemistry, as the maximum recognition for their contribution to the progress of this field.
However, we cannot forget pivotal achievements before those above-mentioned. Hence, regarding non-asymmetric organocatalysis, Kelly [5], Jorgensen [6,7], and Curran [8] are relevant authors, as well as Jacobsen [9] for the use of chiral thioureas as catalysts (Figure 1). They all became a source of inspiration for one of the greatest subareas within this field: activation by hydrogen bond interactions. Even earlier, other researchers, e.g., Bredig and Fiske [10,11], Pracejus [12,13], and Wynberg [14], employed Cinchona derivatives as catalysts, contributing to the subsequent launch of a parcel in asymmetric catalysis centered on the use of alkaloids as chiral catalytic bases. Moreover, the singular chemistry of N-heterocyclic carbenes [15], the activation using chiral phase transfer catalysts [16] or chiral guanidines [17], have been known for a long time. More recently, the use of phosphoric acids [18,19] to promote interesting processes has also efficiently contributed to the expansion of this field.
Nowadays, the term asymmetric organocatalysis covers a wide range of organic processes and methodologies, providing efficient and environmentally friendly access to enantiomerically pure compounds, including many drugs and bioactive natural products.
The origin of our research group can be traced to Professor Ricci’s group (Bologna, Italy), where we took our first steps in this burgeoning field. In 2003, when this discipline was still in its infancy and talk about asymmetric organocatalysis was barely common, as almost nobody knew this field, we explored the first approach to the Friedel–Crafts-type reaction (a well-known metal-catalyzed process) using simple thioureas as organic catalysts [20]. Looking back, we remember a funny anecdote. “Are you sure that you do not have any metal atom in the reaction medium?”. This was the question that a professor asked us in front of our poster at a congress in Ischia (Italy, 2004). Nowadays, no one would doubt this aspect.
Certainly, at the beginning, using organic molecules to promote and activate reactions was nearly a dogma of faith, maybe because the hegemony of metals was so great that it was hard to think that simpler structures could unseat this supremacy. Although there is still a long way to go, we are humbly proud to note that today many organocatalytic processes are at the level of those previously developed with metal catalysts, and in some cases, the achievements of metal catalysis have even been surpassed [21,22]. Nevertheless, the great success of this novel field does not imply the disappearance of the previously existing ones, i.e., metal and biocatalysis, which can be considered complementary and completely necessary for the progress and evolution of modern chemistry. In fact, the combination of organocatalysis and metal- or bio-catalysis for diversity-oriented synthesis is expected to broaden the scope of synthesis much more by allowing genuinely new drugs to be discovered. This approach is innovative and requires the integration of very different disciplines within common scientific efforts.
Many research groups have devoted their efforts to overcome these initial fears and during all these years, very interesting strategies have been developed. While at the beginning there was still a great deal to be done and maybe the first investigations were focused on more basic and known reactions, step-by-step, asymmetric organocatalysis has made its way into the field of catalysis, reaching a remarkable level of complexity in terms of both new reactions and more active catalysts. In this field, the authors’ imagination to design increasingly complex molecules has perhaps even outweighed the lower reactivity of the organocatalysts. In our case, we moved from structurally simple thioureas [23] to the discovery of the first demonstrated trifunctional squaramide catalyst supported with computational calculations [24,25]. This species exhibited an unusually high activity compared to other organocatalysts without impairing the enantioselectivity of the chemical process, since a catalyst loading of only 0.25 mol% is required versus 5–20 mol% usually employed (Scheme 2) [26].
The tremendous grade of complexity accomplished in the new molecular structures obtained by means of organocatalytic reactions during this time is also noteworthy. Without belittling Nature, many authors have been able to obtain numerous new stereogenic centers following a cascade of stereoselective reactions, as biomimetic of tandem reactions that take place during the biosynthesis of complex natural products [27], with incredible enantiocontrol (up to >99% ee) (Scheme 3) [28].
Taking advantage of the capacity of many organocatalysts, as a singular property of these molecules, to promote several different types of reactions through diverse activation modes, during these two decades, new modes of activation have also been discovered or potentially developed mainly in this field. Hence, bifunctional [29], trifunctional [30], dual [31,32], synergistic [33,34], SOMO (singly occupied molecular orbital) [35,36], and photoredox catalysis [37], among others, have been discovered and tuned. In addition, different sources of energy, such as light [38], microwave [39], ultrasounds [40,41], mechanochemical [42,43], and electrochemical [44,45], have been employed, in all cases looking for greener protocols motivated by the increasing concern over sustainable chemistry. Moreover, the use of different reaction media, such as ionic liquids [46,47], deep eutectic solvents [48,49], water [50,51], more sustainable solvents [52], solvent-free processes [53], and even immobilized catalysts [54,55], have been exciting topics of research, with the main objective of using step-by-step greener conditions [56,57,58]. We cannot forget challenging strategies that have acquired incredible notoriety in this field in order to achieve these common goals, preventing waste, energy, time, or even motivated by economic reasons, as multicomponent [59,60,61], one-pot [62,63,64], cascade, or tandem reactions [65,66,67,68,69]. Nowadays, the protocols developed 15–20 years ago can be revisited to support the ‘eco-design’ of molecules, providing products of high molecular complexity while ensuring sustainability as much as possible.

2. Discussion

If we should choose three different aspects mentioned above among all, we would like to highlight the growing importance of developing new multicomponent reactions, the continuous search for new applications in this field, and the possibility of reaching a remarkable level of complexity in the catalytic structures.

2.1. Organocatalytic Asymmetric Multicomponent Reactions

In the pursuit of developing more efficient and sustainable protocols for synthesis, multicomponent (MCRs) or one-pot reactions should be regarded as the suitable option. In such processes, the time-consuming isolation and purification of synthetic intermediates disappear and, in general, the energy, solvents, manipulations, and therefore costs are reduced, compared to classical stepwise methods. These processes also emulate Nature where metabolic pathways occurring in living cells involve an elegant orchestration of a series of biocatalytic steps into an exquisite multi-catalytic cascade, without the need for the separation of intermediates. Since the renascence of the organocatalytic field, many improvements have been made in the progress of organocatalyzed asymmetric MCRs and one-pot reactions using overall chiral secondary amines [70,71], and less explored with hydrogen bond-based catalysts [72]. These spectacular synthetic strategies serve as powerful tools for the efficient construction of complex molecular architectures, and generally these are based on biomimetic principles.
We would like to comment on a pivotal recent example, since sometimes it is difficult to separate a multicomponent reaction from a cascade process, and the complexity of both processes together can be of extraordinary elegance [73]. Recently, Jørgensen and co-workers reported an amazing example of an organocatalytic multicomponent cascade process via dienamine catalysis, involving at the same time a cycloaddition, a nucleophilic addition, and a ring-opening reaction between uncommon isobenzopyrylium precursors, α,β-unsaturated aldehydes, and H2O [74]. The process was successfully developed to prepare chiral tetrahydronaphthols containing four contiguous stereocenters in good to high yield (up to 95%), high diastereoselectivity (up to >20:1), and excellent enantioselectivity (93−98% ee) (Scheme 4).
On the basis of additional experimental results and computational calculations, the authors proposed a sophisticated multicomponent cascade mechanism supporting the final products obtained (Scheme 5).
First, the catalytic cycle would start with the formation of the dienamine intermedia A by condensation of the catalyst and the aldehyde. The isobenzopyrylium ion intermediate B would be generated in a concomitant process by the release of the ion isopropoxide from the starting reagent. Then, B would be attacked at the carbonyl position by the γ-carbon of dienamine A, giving rise to intermediate C. An ensuing ring-closure would generate intermediate D. The addition of a molecule of water at the carbonyl group would afford intermediate E. Finally, a C-O bond scission, hydrolysis, and proton transfer render final products and release the molecule of catalyst to start the cycle again.
The importance of the final products of this catalytic process was also demonstrated by generating subsequent interesting scaffolds (Scheme 6).

2.2. New Organocatalytic Applications

The interest of the pharmaceutical and agrochemical industries in chiral products has prompted the intensive search for more efficient methods of asymmetric synthesis [75]. The chiral drug industry has become a rapidly growing segment of the drug market and represents close to one-third of all drug sales worldwide [76]. This has been driven by the increased regulatory control of the enantiomeric composition of drug candidates and the potential of enantiomerically pure drugs to provide improvements over the previously available racemates. On the base of this continuous searching for interesting new applications in different areas of knowledge, it is no longer strange to find an organocatalytic protocol, for instance, as part of a total synthesis of a natural product [77,78], or in the synthesis of biologically active compounds [79,80,81] (Figure 2). In fact, during these two decades, asymmetric organocatalysis has evolved as a powerful synthetic tool for the preparation of chiral building blocks with a wide range of applications in medicinal chemistry and drug discovery [82].
Regarding this chapter, Peng, Li and co-workers have reported an incredible a crucial work comprehensively covering all the applications of asymmetric organocatalysis in medicinal chemistry since the early 2000s [90]. Returning to our chemistry, we have also developed some active compounds using organocatalysis as a key step in their synthesis, such as antimicrobial β-nitrohidrazydes [91,92], Warfarin and derivatives [93], iridodials [94,95], or the key intermediate to the analgesic Funapide [96] (Figure 3).
More recently, we have become interested in the preparation and evaluation of potential anticancer drugs, as cancer is one of the leading causes of death in the world. Hence, while working on the development of organocatalytic methodologies, we became aware of the features of the final products, as well as the interesting properties that the organocatalysts themselves could exhibit beyond their catalytic activity, since most of them are structural cores present in Nature, as (thio)ureas and cinchonas. For this reason, we have started to explore the antitumor properties of some organocatalysts such as squaramides [97,98], as well as the in vitro and in vivo activity of 1,4-dihydropyridine derivatives synthesized following an organocatalytic protocol as potential anticancer drugs [99] (Figure 4).

2.3. Increasingly Complex Catalytic Structures

Although organocatalysis has experienced great progress in the field of homogeneous catalysis, where rate and enantioselectivity together define the catalytic proficiency, the organocatalysts still require much more improvement to emulate the achievements with metal or enzyme catalysts. In this sense, the general catalyst loading is still in the range of 5–20 mol% and only in some cases this has been lower. Therefore, the catalytic activity and turnover number of metal catalysts or an enzyme are extremely high in comparison with an organocatalyst. These problems increase when at lower temperature the reaction rate decreases and a greater amount of organocatalyst is necessary to compare its activity with the same reaction at room temperature. Although organocatalysts have often been proposed as “artificial enzymes” or small enzyme mimics [100], there are still many differences and a long way to go to reach them. Among other factors, they are still inferior to natural enzymes in terms of catalytic activity and turnover [101]. One strategy to overcome this inconvenience has been to develop more acidic organocatalysts [102]. Another strategy has been to design plausible bifunctional organocatalysts [103,104,105] following the prevalence of multifunctional catalysis in enzyme catalysis. In this sense, the bifunctionality has served as the inspiration for many synthetic catalytic systems, such as chiral thioureas/ureas, cinchona alkaloids, chiral phosphoric acids, or chiral proline derivatives, which retain the simultaneous activation of nucleophile and electrophile. Interestingly, and more recently, the term trifunctional has acquired more importance [106,107]. Generally, this family of molecules is expected to activate the process by their three different functionalities, achieving optimal cooperativity and increasing the activity of the catalytic species. Some of these pioneering structures that are believed to act in a trifunctional manner are described in the Figure 5. In all these cases, the authors based their tentative proposal only on experimental results, but in any case, with additional computational calculations supporting a real trifunctional role of the catalysts or not.
It is important to remark that only in our developed example [24,25,26] did additional computational calculations support the trifunctional mode of action demonstrated by the catalysts. Therefore, we think that there is still room in this field to improve and to discover new real trifunctional catalysts, and even tretrafunctional ones.

3. Conclusions

We are convinced that even after the award of the Nobel Prize this year, there is still a long way to go in searching for new milestones in this field. Undoubtedly, new challenges with all the aforementioned characteristics and new sustainable aspects converging in the processes (sustainable methodology, solvent, energy source, structural complexity, and final application) will emerge in the near future. This opinion article is also a small tribute to the efforts and pivotal contributions made by all authors in this appealing field. Indeed, we apologize we cannot name every author involved, as we would desire, and whose seminal and crucial work has also contributed to the development of this broad area of research. We hope that researchers, encouraged by interesting applications [112], will invest efforts in order to expand the scope of this kind of catalysis, because these goals will also be crucial for the evolution of the pharmaceutical industry, among other sectors.

Author Contributions

Writing—original draft preparation, all authors; writing—review and editing, all authors; visualization, E.M.-L. and R.P.H.; supervision E.M.-L. and R.P.H.; project administration, E.M.-L. and R.P.H.; funding acquisition, E.M.-L. and R.P.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Agencia Estatal de Investigación (AEI), project PID2020-117455GB-I00/AEI/10.13039/501100011033; and Gobierno de Aragón-Fondo Social Europeo (Research Group E07_20R).

Acknowledgments

S.A. thanks the DGA (Aragon Government) for a predoctoral fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Carlone, A.; Bernardi, L. Enantioselective Organocatalytic Approaches to Active Pharmaceutical Ingredients—Selected Industrial Examples. In Organocatalysis: Stereoselective Reactions and Applications in Organic Synthesis; Benaglia, M., Ed.; De Gruyter: Berlin, Germany; Boston, MA, USA, 2021; pp. 401–434. [Google Scholar] [CrossRef]
  2. Marckwald, W. Ueber Asymmetrische Synthese. Ber. Dtsch. Chem. Ges. 1904, 37, 349–354. [Google Scholar] [CrossRef] [Green Version]
  3. List, B.; Lerner, R.A.; Barbas, C.F., III. Proline-Catalyzed Direct Asymmetric Aldol Reactions. J. Am. Chem. Soc. 2000, 122, 2395–2396. [Google Scholar] [CrossRef]
  4. Ahrendt, K.A.; Borths, C.J.; MacMillan, D.W.C. New Strategies for Organic Catalysis: The First Highly Enantioselective Organocatalytic Diels–Alder Reaction. J. Am. Chem. Soc. 2000, 122, 4243–4244. [Google Scholar] [CrossRef]
  5. Kelly, T.R.; Meghani, P.; Ekkundi, V.S. Diels–Alder reactions: Rate acceleration promoted by a biphenylenediol. Tetrahedron Lett. 1990, 31, 3381–3384. [Google Scholar] [CrossRef]
  6. Blake, J.F.; Jorgensen, W.L. Solvent Effects on a Diels–Alder Reaction from Computer Simulations. J. Am. Chem. Soc. 1991, 113, 7430–7432. [Google Scholar] [CrossRef]
  7. Severance, D.L.; Jorgensen, W.L. Effects of Hydration on the Claisen Rearrangement of Allyl Vinyl Ether from Computer Simulations. J. Am. Chem. Soc. 1992, 114, 10966–10968. [Google Scholar] [CrossRef]
  8. Curran, D.P.; Kuo, L.H. Altering the Stereochemistry of Allylation Reactions of Cyclic α-Sulfinyl Radicals with Diarylureas. J. Org. Chem. 1994, 59, 3259–3261. [Google Scholar] [CrossRef]
  9. Sigman, M.S.; Jacobsen, E.N. Schiff Base Catalysts for the Asymmetric Strecker Reaction Identified and Optimized from Parallel Synthetic Libraries. J. Am. Chem. Soc. 1998, 120, 4901–4902. [Google Scholar] [CrossRef]
  10. Bredig, G.; Fiske, P.S. Asymmetric Syntheses Caused by Catalyzers. Biochem. Z. 1913, 46, 7–23. [Google Scholar]
  11. Bredig, G.; Minaeff, M. Asymmetric synthesis by means of catalysts. II. Biochem. Z. 1932, 249, 241–244. [Google Scholar]
  12. Pracejus, H. Organische Katalysatoren, LXI. Asymmetrische Synthesen mit Ketenen, I. Alkaloid-katalysierte asymmetrische Synthesen von α-Phenyl-propionsäureestern. Justus Liebigs Ann. Chem. 1960, 634, 9–22. [Google Scholar] [CrossRef]
  13. Pracejus, H.; Mätje, H. Organische Katalysatoren. LXXI Asymmetrische Synthesen mit Ketenen. IV. Zusammenhänge zwischen dem räumlichen Bau einiger alkaloidartiger Katalysatoren und ihren stereospezifischen Wirkungen bei asymmetrischen Estersynthesen. J. Prakt. Chem. 1964, 24, 195–205. [Google Scholar] [CrossRef]
  14. Hiemstra, H.; Wynberg, H. Addition of Aromatic Thiols to Conjugated Cycloalkenones, Catalyzed by Chiral β-Hydroxy Amines. A Mechanistic Study of Homogeneous Catalytic Asymmetric Synthesis. J. Am. Chem. Soc. 1981, 103, 417–430. [Google Scholar] [CrossRef]
  15. Ukai, T.; Tanaka, R.; Dokawa, T. アチロイン縮合に於ける新觸媒群に就て (第1報) チアツオリウム鹽基或は共核開裂成績體の作用. J. Pharm. Soc. Jpn. 1943, 63, 296–300. [Google Scholar] [CrossRef]
  16. Makosza, M. Reactions of organic anions. XI. Catalytic alkylation of indene. Tetrahedron Lett. 1966, 7, 4621–4624. [Google Scholar] [CrossRef]
  17. Chinchilla, R.; Najera, C.; Sanchez-Agullo, P. Enantiomerically pure guanidine-catalysed asymmetric nitroaldol reaction. Tetrahedron Asymmetry 1994, 5, 1393–1402. [Google Scholar] [CrossRef]
  18. Akiyama, T.; Itoh, J.; Yokota, K.; Fuchibe, K. Enantioselective Mannich-type reaction catalyzed by a chiral Brønsted acid. Angew. Chem. Int. Ed. 2004, 43, 1566–1568. [Google Scholar] [CrossRef]
  19. Uraguchi, D.; Terada, M. Chiral Brønsted acid-catalyzed direct Mannich reactions via electrophilic activation. J. Am. Chem. Soc. 2004, 126, 5356–5357. [Google Scholar] [CrossRef]
  20. Herrera, R.P.; Sgarzani, V.; Bernardi, L.; Ricci, A. Catalytic Enantioselective Friedel–Crafts Alkylation of Indoles with Nitroalkenes by a Simple Thiourea Organocatalyst. Angew. Chem. Int. Ed. 2005, 44, 6576–6579. [Google Scholar] [CrossRef]
  21. Bertelsen, S.; Marigo, M.; Brandes, S.; Dinér, P.; Jørgensen, K.A. Dienamine Catalysis: Organocatalytic Asymmetric γ-Amination of α,β-Unsaturated Aldehydes. J. Am. Chem. Soc. 2006, 128, 12973–12980. [Google Scholar] [CrossRef]
  22. Jia, Z.-J.; Jiang, H.; Li, J.-L.; Gschwend, B.; Li, Q.-Z.; Yin, X.; Grouleff, J.; Chen, Y.-C.; Jørgensen, K.A. Trienamines in Asymmetric Organocatalysis: Diels–Alder and Tandem Reactions. J. Am. Chem. Soc. 2011, 133, 5053–5061. [Google Scholar] [CrossRef]
  23. Dessole, G.; Herrera, R.P.; Ricci, A. H-Bonding Organocatalysed Friedel–Crafts Alkylation of Aromatic and Heteroaromatic Systems with Nitroolefins. Synlett 2004, 2374–2378. [Google Scholar] [CrossRef]
  24. Alegre-Requena, J.V.; Marqués-López, E.; Herrera, R.P. Optimizing Accuracy and Computational Cost in Theoretical Squaramide Catalysis: The Henry Reaction. Chem. Eur. J. 2017, 23, 15336–15347. [Google Scholar] [CrossRef] [Green Version]
  25. Alegre-Requena, J.V.; Marqués-López, E.; Herrera, R.P. “Push-Pull π+/π-” (PPππ) Systems in Catalysis. ACS Catal. 2017, 7, 6430–6439. [Google Scholar] [CrossRef] [Green Version]
  26. Alegre-Requena, J.V.; Marqués-López, E.; Herrera, R.P. Trifunctional squaramide catalyst for efficient enantioselective Henry reaction activation. Adv. Synth. Catal. 2016, 358, 1801–1809. [Google Scholar] [CrossRef] [Green Version]
  27. Mann, J. (Ed.) Chemical Aspects of Biosynthesis. In Oxford Chemistry Primers; Oxford University Press: Oxford, UK, 1994. [Google Scholar]
  28. Enders, D.; Hüttl, M.R.M.; Grondal, C.; Raabe, G. Control of four stereocentres in a triple cascade organocatalytic reaction. Nature 2006, 441, 861–863. [Google Scholar] [CrossRef]
  29. Okino, T.; Hoashi, Y.; Takemoto, Y. Enantioselective Michael Reaction of Malonates to Nitroolefins Catalyzed by Bifunctional Organocatalysts. J. Am. Chem. Soc. 2003, 125, 12672–12673. [Google Scholar] [CrossRef]
  30. Chen, W.; Du, W.; Duan, Y.-Z.; Wu, Y.; Yang, S.-Y.; Chen, Y.-C. Enantioselective 1,3-Dipolar Cycloaddition of Cyclic Enones Catalyzed by Multifunctional Primary Amines: Beneficial Effects of Hydrogen Bonding. Angew. Chem. Int. Ed. 2007, 46, 7667–7670. [Google Scholar] [CrossRef]
  31. Nakoji, M.; Kanayama, T.; Okino, T.; Takemoto, Y. Chiral Phosphine-Free Pd-Mediated Asymmetric Allylation of Prochiral Enolate with a Chiral Phase-Transfer Catalyst. Org. Lett. 2001, 3, 3329–3331. [Google Scholar] [CrossRef]
  32. Nakoji, M.; Kanayama, T.; Okino, T.; Takemoto, Y. Pd-Catalyzed Asymmetric Allylic Alkylation of Glycine Imino Ester Using a Chiral Phase-Transfer Catalyst. J. Org. Chem. 2002, 67, 7418–7423. [Google Scholar] [CrossRef]
  33. Ibrahem, I.; Breistein, P.; Córdova, A. One-Pot Three-Component Catalytic Enantioselective Synthesis of Homoallylboronates. Angew. Chem. Int. Ed. 2011, 50, 12036–12041. [Google Scholar] [CrossRef]
  34. Jiang, G.; List, B. Direct Asymmetric α-Allylation of Aldehydes with Simple Allylic Alcohols Enabled by the Concerted Action of Three Different Catalysts. Angew. Chem. Int. Ed. 2011, 50, 9471–9474. [Google Scholar] [CrossRef]
  35. Beeson, T.D.; Mastracchio, A.; Hong, J.-B.; Ashton, K.; MacMillan, D.W.C. Enantioselective Organocatalysis Using SOMO Activation. Science 2007, 316, 582–585. [Google Scholar] [CrossRef]
  36. Sibi, M.P.; Hasegawa, M. Organocatalysis in Radical Chemistry. Enantioselective α-Oxyamination of Aldehydes. J. Am. Chem. Soc. 2007, 129, 4124–4125. [Google Scholar] [CrossRef] [Green Version]
  37. Nicewicz, D.A.; MacMillan, D.W.C. Merging photoredox catalysis with organocatalysis: The direct asymmetric alkylation of aldehydes. Science 2008, 322, 77–80. [Google Scholar] [CrossRef] [Green Version]
  38. Silvi, M.; Melchiorre, P. Enhancing the potential of enantioselective organocatalysis with light. Nature 2018, 554, 41–49. [Google Scholar] [CrossRef]
  39. Westermann, B.; Neuhaus, C. Dihydroxyacetone in Amino Acid Catalyzed Mannich-Type Reactions. Angew. Chem. Int. Ed. Engl. 2005, 44, 4077–4079. [Google Scholar] [CrossRef]
  40. Chatel, G.; Varma, R.S. Ultrasound and microwave irradiation: Contributions of alternative physicochemical activation methods to Green Chemistry. Green Chem. 2019, 21, 6043–6050. [Google Scholar] [CrossRef]
  41. Kantam, M.L.; Rajasekhar, C.V.; Gopikrishna, G.; Reddy, K.R.; Choudary, B.M. Proline catalyzed two-component, three-component and self-asymmetric Mannich reactions promoted by ultrasonic conditions. Tetrahedron Lett. 2006, 47, 5965–5967. [Google Scholar] [CrossRef]
  42. Rodríguez, B.; Rantanen, T.; Bolm, C. Solvent-Free Asymmetric Organocatalysis in a Ball Mill. Angew. Chem. Int. Ed. 2006, 45, 6924–6926. [Google Scholar] [CrossRef]
  43. Chauhan, P.; Chimni, S.S. Mechanochemistry assisted asymmetric organocatalysis: A sustainable approach. Beilstein J. Org. Chem. 2012, 8, 2132–2141. [Google Scholar] [CrossRef] [Green Version]
  44. Jensen, K.L.; Franke, P.T.; Nielsen, L.T.; Daasbjerg, K.; Jørgensen, K.A. Anodic Oxidation and Organocatalysis: Direct Regio- and Stereoselective Access to meta-Substituted Anilines by α-Arylation of Aldehydes. Angew. Chem. Int. Ed. 2010, 49, 129–133. [Google Scholar] [CrossRef]
  45. Yan, M.; Kawamata, Y.; Baran, P.S. Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance. Chem. Rev. 2017, 117, 13230–13319. [Google Scholar] [CrossRef]
  46. Kotrusz, P.; Kmentová, I.; Gotov, B.; Toma, Š.; Solčániová, E. Proline-catalysed asymmetric aldol reaction in the room temperature ionic liquid [bmim]PF6. Chem. Commun. 2002, 21, 2510–2511. [Google Scholar] [CrossRef]
  47. Zalewska, K.; Branco, L.C. Organocatalysis with Chiral Ionic Liquids. Mini-Rew. Org. Chem. 2014, 11, 141–153. [Google Scholar] [CrossRef]
  48. Alonso, D.A.; Burlingham, S.-J.; Chinchilla, R.; Guillena, G.; Ramón, D.J.; Tiecco, M. Asymmetric Organocatalysis in Deep Eutectic Solvents. Eur. J. Org. Chem. 2021, 2021, 4065–4071. [Google Scholar] [CrossRef]
  49. Fanjul-Mosteirín, N.; del Amo, V. Organocatalytic transformations in deep eutectic solvents: Green methodologies made greener. Tetrahedron 2021, 84, 131967. [Google Scholar] [CrossRef]
  50. Sakthivel, K.; Notz, W.; Bui, T.; Barbas, C.F., III. Amino Acid Catalyzed Direct Asymmetric Aldol Reactions: A Bioorganic Approach to Catalytic Asymmetric Carbon-Carbon Bond-Forming Reactions. J. Am. Chem. Soc. 2001, 123, 5260–5267. [Google Scholar] [CrossRef]
  51. Gruttadauria, M.; Giacalone, F.; Noto, R. Water in Stereoselective Organocatalytic Reactions. Adv. Synth. Catal. 2009, 351, 33–57. [Google Scholar] [CrossRef]
  52. Clarke, C.J.; Tu, W.-C.; Levers, O.; Bröhl, A.; Hallett, J.P. Green and Sustainable Solvents in Chemical Processes. Chem. Rev. 2018, 118, 747–800. [Google Scholar] [CrossRef]
  53. Avila-Ortiz, C.G.; Juaristi, E. Novel Methodologies for Chemical Activation in Organic Synthesis under Solvent-Free Reaction Conditions. Molecules 2020, 25, 3579. [Google Scholar] [CrossRef]
  54. Cozzi, F. Immobilization of Organic Catalysts: When, Why, and How. Adv. Synth. Catal. 2006, 348, 1367–1390. [Google Scholar] [CrossRef]
  55. Ferré, M.; Pleixats, R.; Man, M.W.C.; Cattoën, X. Recyclable organocatalysts based on hybrid silicas. Green Chem. 2016, 18, 881–922. [Google Scholar] [CrossRef] [Green Version]
  56. Bruckmann, A.; Krebs, A.; Bolm, C. Organocatalytic reactions: Effects of ball milling, microwave and ultrasound irradiation. Green Chem. 2008, 10, 1131–1141. [Google Scholar] [CrossRef]
  57. Krištofíková, D.; Modrocká, V.; Mečiarová, M.; Šebesta, R. Green Asymmetric Organocatalysis. ChemSusChem 2020, 13, 2828–2858. [Google Scholar] [CrossRef]
  58. Antenucci, A.; Dughera, S.; Renzi, P. Green Chemistry Meets Asymmetric Organocatalysis: A Critical Overview on Catalysts Synthesis. ChemSusChem 2021, 14, 2785–2853. [Google Scholar] [CrossRef]
  59. Zhu, J.; Bienaymé, H. (Eds.) Multicomponent Reactions; Wiley-VCH: Weinheim, Germany, 2005. [Google Scholar]
  60. Zhu, J.; Wang, Q.; Wang, M.-X. (Eds.) Multicomponent Reactions in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 2014. [Google Scholar]
  61. Herrera, R.P.; Marqués-López, E. (Eds.) Multicomponent Reactions. Concepts and Applications for Design and Synthesis; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015. [Google Scholar]
  62. Walji, A.M.; MacMillan, D.W.C. Strategies to Bypass the Taxol Problem. Enantioselective Cascade Catalysis, a New Approach for the Efficient Construction of Molecular Complexity. Synlett 2007, 2007, 1477–1489. [Google Scholar] [CrossRef]
  63. Hayashi, Y. Pot economy and one-pot synthesis. Chem. Sci. 2016, 7, 866–880. [Google Scholar] [CrossRef] [Green Version]
  64. Corma, A.; Navas, J.; Sabater, M.J. Advances in One-Pot Synthesis through Borrowing Hydrogen Catalysis. Chem. Rev. 2018, 118, 1410–1459. [Google Scholar] [CrossRef]
  65. Yu, X.; Wang, W. Organocatalysis: Asymmetric cascade reactions catalysed by chiral secondary amines. Org. Biomol. Chem. 2008, 6, 2037–2046. [Google Scholar] [CrossRef]
  66. Grondal, C.; Jeanty, M.; Enders, D. Organocatalytic cascade reactions as a new tool in total synthesis. Nat. Chem. 2010, 2, 167–178. [Google Scholar] [CrossRef]
  67. Vetica, F.; de Figueiredo, R.M.; Orsini, M.; Tofani, D.; Gasperi, T. Recent Advances in Organocatalytic Cascade Reactions toward the Formation of Quaternary Stereocenters. Synthesis 2015, 47, 2139–2184. [Google Scholar] [CrossRef]
  68. Evans, C.S.; Davis, L.O. Recent Advances in Organocatalyzed Domino C–C Bond-Forming Reactions. Molecules 2018, 23, 33. [Google Scholar] [CrossRef] [Green Version]
  69. Bortolami, M.; Leonelli, F.; Feroci, M.; Vetica, F. Step Economy in the Stereoselective Synthesis of Functionalized Oxindoles via Organocatalytic Domino/One-pot Reactions. Curr. Org. Chem. 2021, 25, 1321–1344. [Google Scholar] [CrossRef]
  70. Enders, D.; Grondal, C.; Hüttl, M.R.M. Asymmetric Organocatalytic Domino Reactions. Angew. Chem. Int. Ed. 2007, 46, 1570–1581. [Google Scholar] [CrossRef] [PubMed]
  71. Guillena, G.; Ramón, D.J.; Yus, M. Organocatalytic enantioselective multicomponent reactions (OEMCRs). Tetrahedron Asymmetry 2007, 18, 693–700. [Google Scholar] [CrossRef]
  72. Parvin, T.; Yadava, R.; Choudhury, L.H. Recent applications of thiourea-based organocatalysts in asymmetric multicomponent reactions (AMCRs). Org. Biomol. Chem. 2020, 18, 5513–5532. [Google Scholar] [CrossRef]
  73. Volla, C.M.R.; Atodiresei, I.; Rueping, M. Catalytic C−C Bond-Forming Multi-Component Cascade or Domino Reactions: Pushing the Boundaries of Complexity in Asymmetric Organocatalysis. Chem. Rev. 2014, 114, 2390–2431. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, Y.; Izzo, J.A.; McLeod, D.; Ričko, S.; Svenningsen, E.B.; Poulsen, T.B.; Jørgensen, K.A. Organocatalytic Asymmetric Multicomponent Cascade Reaction for the Synthesis of Contiguously Substituted Tetrahydronaphthols. J. Am. Chem. Soc. 2021, 143, 8208–8220. [Google Scholar] [CrossRef]
  75. Francotte, E.; Lindner, W. (Eds.) Chirality in Drug Research; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  76. Nicolaou, K.C.; Montagnon, T.; Snyder, S.A. Tandem reactions, cascade sequences, and biomimetic strategies in total synthesis. Chem. Commun. 2003, 5, 551–564. [Google Scholar] [CrossRef] [PubMed]
  77. de Figueiredo, R.M.; Christmann, M. Organocatalytic Synthesis of Drugs and Bioactive Natural Products. Eur. J. Org. Chem. 2007, 2007, 2575–2600. [Google Scholar] [CrossRef]
  78. Marqués-López, E.; Herrera, R.P.; Christmann, M. Asymmetric organocatalysis in total synthesis—A trial by fire. Nat. Prod. Rep. 2010, 27, 1138–1167. [Google Scholar] [CrossRef]
  79. Alemán, J.; Cabrera, S. Applications of asymmetric organocatalysis in medicinal chemistry. Chem. Soc. Rev. 2013, 42, 774–793. [Google Scholar] [CrossRef] [PubMed]
  80. Sun, B.-F. Total synthesis of natural and pharmaceutical products powered by organocatalytic reactions. Tetrahedron Lett. 2015, 56, 2133–2140. [Google Scholar] [CrossRef] [Green Version]
  81. Pellissier, H. Asymmetric Organocatalytic Tandem/Domino Reactions to Access Bioactive Products. Curr. Org. Chem. 2021, 25, 1457–1471. [Google Scholar] [CrossRef]
  82. Hughes, D.L. Asymmetric Organocatalysis in Drug Development–Highlights of Recent Patent Literature. Org. Process Res. Dev. 2018, 22, 574–584. [Google Scholar] [CrossRef]
  83. Ward, D.E.; Jheengut, V.; Beye, G.E. Thiopyran Route to Polypropionates: An Efficient Synthesis of Serricornin. J. Org. Chem. 2006, 71, 8989–8992. [Google Scholar] [CrossRef] [PubMed]
  84. Malkov, A.V.; Kabeshov, M.A.; Bella, M.; Kysilka, O.; Malyshev, D.A.; Pluháčková, K.; Kočovský, P. Vicinal Amino Alcohols as Organocatalysts in Asymmetric Cross-Aldol Reaction of Ketones: Application in the Synthesis of Convolutamydine A. Org. Lett. 2007, 9, 5473–5476. [Google Scholar] [CrossRef]
  85. Andrey, O.; Vidonne, A.; Alexakis, A. Organocatalytic Michael addition, a convenient tool in total synthesis. First asymmetric synthesis of (−)-botryodiplodin. Tetrahedron Lett. 2003, 44, 7901–7904. [Google Scholar] [CrossRef]
  86. Mukaiyama, T.; Ishikawa, H.; Koshino, H.; Hayashi, Y. One-Pot Synthesis of (−)-Oseltamivir and Mechanistic Insights into the Organocatalyzed Michael Reaction. Chem. Eur. J. 2013, 19, 17789–17800. [Google Scholar] [CrossRef]
  87. Zhang, F.-G.; Zhu, X.-Y.; Li, S.; Nie, J.; Ma, J.-A. Highly Enantioselective Organocatalytic Strecker Reaction of Cyclic N-Acyl Trifluoromethylketimines: Synthesis of anti-HIV Drug DPC 083. Chem. Commun. 2012, 48, 11552–11554. [Google Scholar] [CrossRef] [PubMed]
  88. Xu, F.; Zacuto, M.; Yoshikawa, N.; Desmond, R.; Hoerrner, S.; Itoh, T.; Journet, M.; Humphrey, G.R.; Cowden, C.; Strotman, N.; et al. Asymmetric Synthesis of Telcagepant, a CGRP Receptor Antagonist for the Treatment of Migraine. J. Org. Chem. 2010, 75, 7829–7841. [Google Scholar] [CrossRef] [PubMed]
  89. Gualtierotti, J.-B.; Pasche, D.; Wang, Q.; Zhu, J. Phosphoric acid catalyzed desymmetrization of bicyclic bislactones bearing an all-carbon stereogenic center: Total syntheses of (−)-Rhazinilam and (−)-Leucomidine B. Angew. Chem. Int. Ed. 2014, 53, 9926–9930. [Google Scholar] [CrossRef]
  90. Han, B.; He, X.-H.; Liu, Y.-Q.; He, G.; Peng, C.; Li, J.-L. Asymmetric organocatalysis: An enabling technology for medicinal chemistry. Chem. Soc. Rev. 2021, 50, 1522–1586. [Google Scholar] [CrossRef] [PubMed]
  91. Alcaine, A.; Marqués-López, E.; Herrera, R.P. Synthesis of Interesting β-Nitrohydrazides through Thiourea Organocatalyzed aza-Michael Addition. RSC Adv. 2014, 4, 9856–9865. [Google Scholar] [CrossRef] [Green Version]
  92. Sonsona, I.G.; Alegre-Requena, J.V.; Marqués-López, E.; Gimeno, M.C.; Herrera, R.P. Asymmetric organocatalyzed aza-Henry reaction of hydrazones: Experimental and computational studies. Chem. Eur. J. 2020, 26, 5469–5478. [Google Scholar] [CrossRef]
  93. Sonsona, I.G.; Marqués-López, E.; Gimeno, M.C.; Herrera, R.P. First aromatic amine organocatalyzed activation of ketones. New J. Chem. 2019, 43, 12233–12240. [Google Scholar] [CrossRef] [Green Version]
  94. Marqués-López, E.; Herrera, R.P.; Marks, T.; Jacobs, W.C.; Könning, D.; de Figueiredo, R.M.; Christmann, M. Crossed Intramolecular Rauhut-Currier-Type Reactions via Dienamine Activation. Org. Lett. 2009, 11, 4116–4119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Marqués-López, E.; Herrera, R.P.; Marks, T.; Jacobs, W.C.; Christmann, M. Enantioselective Rauhut-Currier-Type Cyclizations via Dienamine Activation: Scope and Mechanism. Synthesis 2013, 45, 1016–1028. [Google Scholar] [CrossRef] [Green Version]
  96. Sonsona, I.G.; Vicenzi, A.; Guidotti, M.; Bisag, G.D.; Fochi, M.; Herrera, R.P.; Bernardi, L. Investigation of Squaramide Catalysts in the Aldol Reaction en Route to Funapide. Eur. J. Org. Chem. 2022, 2022, e202101254. [Google Scholar] [CrossRef]
  97. Fernández-Moreira, V.; Alegre-Requena, J.V.; Herrera, R.P.; Marzo, I.; Gimeno, M.C. Synthesis of luminescent squaramide monoesters: Cytotoxicity and cell imaging studies in HeLa cells. RSC Adv. 2016, 6, 14171–14177. [Google Scholar] [CrossRef] [Green Version]
  98. Quintana, M.; Alegre-Requena, J.V.; Marqués-López, E.; Herrera, R.P.; Triola, G. Squaramides with cytotoxic activity against human gastric carcinoma cells HGC-27: Synthesis and mechanism of action. Med. Chem. Commun. 2016, 7, 550–561. [Google Scholar] [CrossRef] [Green Version]
  99. Auria-Luna, F.; Marqués-López, E.; Romanos, E.; Fernández-Moreira, V.; Gimeno, M.C.; Marzo, I.; Herrera, R.P. Novel ureido-dihydropyridine scaffolds as theranostic agents. Bioorg. Chem. 2020, 105, 104364. [Google Scholar] [CrossRef] [PubMed]
  100. Liu, F. The Upside of Downsizing: Asymmetric Trifunctional Organocatalysts as Small Enzyme Mimics for Cooperative Enhancement of Both Rate and Enantioselectivity with Regulation. Chirality 2013, 25, 675–683. [Google Scholar] [CrossRef]
  101. D’Souza, V.T.; Bender, M.L. Miniature Organic Models of Enzymes. Acc. Chem. Res. 1987, 20, 146–152. [Google Scholar] [CrossRef]
  102. Gimeno, M.C.; Herrera, R.P. Hydrogen Bonding and Internal or External Lewis or Brønsted Acid Assisted (Thio)urea Catalysts. Eur. J. Org. Chem. 2020, 2020, 1057–1068. [Google Scholar] [CrossRef]
  103. Miyabe, H.; Takemoto, Y. Discovery and Application of Asymmetric Reaction by Multi-Functional Thioureas. Bull. Chem. Soc. Jpn. 2008, 81, 785–795. [Google Scholar] [CrossRef] [Green Version]
  104. Connon, S.J. Asymmetric catalysis with bifunctional cinchona alkaloid-based urea and thiourea organocatalysts. Chem. Commun. 2008, 2499–2510. [Google Scholar] [CrossRef] [Green Version]
  105. Sonsona, I.G.; Marqués-López, E.; Herrera, R.P. The aminoindanol core as a key scaffold in bifunctional organocatalysts. Beilstein J. Org. Chem. 2016, 12, 505–523. [Google Scholar] [CrossRef] [Green Version]
  106. Kenny, R.; Liu, F. Trifunctional Organocatalysts: Catalytic Proficiency by Cooperative Activation. Eur. J. Org. Chem. 2015, 2015, 5304–5319. [Google Scholar] [CrossRef]
  107. Kenny, R.; Liu, F. Cooperative Trifunctional Organocatalysts for Proficient Proton Transfer Reactions. Chem. Rec. 2017, 17, 535–553. [Google Scholar] [CrossRef] [PubMed]
  108. Tan, B.; Candeias, N.R.; Barbas, C.F., III. Construction of bispirooxindoles containing three quaternary stereocentres in a cascade using a single multifunctional organocatalyst. Nat. Chem. 2011, 3, 473–477. [Google Scholar] [CrossRef]
  109. Zhong, N.-J.; Liu, L.; Wang, D.; Chen, Y.-J. Enantioselective tandem reaction of chromone-derived Morita–Baylis–Hillman carbonates with benzylamines catalyzed by a trifunctional organocatalyst: The synthesis of chiral 3-aminomethylene-flavanones. Chem. Commun. 2013, 49, 3697–3699. [Google Scholar] [CrossRef] [PubMed]
  110. Ding, X.-H.; Li, X.; Liu, D.; Cui, W.-C.; Ju, X.; Wang, S.; Yao, Z.-J. Target-oriented multifunctional organocatalysis for enantioselective synthesis of bicyclo[3.3.1]nona-2,6-dien-9-ones. A formal asymmetric synthesis of huperzine A. Tetrahedron 2012, 68, 6240–6248. [Google Scholar] [CrossRef]
  111. Garnier, J.-M.; Liu, F. Trifunctional organocatalyst-promoted counterion catalysis for fast and enantioselective aza-Morita–Baylis–Hillman reactions at ambient temperature. Org. Biomol. Chem. 2009, 7, 1272–1275. [Google Scholar] [CrossRef] [PubMed]
  112. Marqués-López, E.; Herrera, R.P. Asymmetric Organocatalysis in Total Synthesis. In Comprehensive Enantioselective Organocatalysis; Dalko, P., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; Chapter 13; pp. 1359–1383. [Google Scholar]
Scheme 1. Pioneering processes reported by (a) List [3] and (b) MacMillan [4].
Scheme 1. Pioneering processes reported by (a) List [3] and (b) MacMillan [4].
Catalysts 12 00101 sch001
Figure 1. Crucial organocatalytic structures reported previously to 2000.
Figure 1. Crucial organocatalytic structures reported previously to 2000.
Catalysts 12 00101 g001
Scheme 2. Henry reaction catalyzed by trifunctional organocatalysts VII.
Scheme 2. Henry reaction catalyzed by trifunctional organocatalysts VII.
Catalysts 12 00101 sch002
Scheme 3. Multicomponent triple cascade organocatalytic reaction giving rise to four stereocenters.
Scheme 3. Multicomponent triple cascade organocatalytic reaction giving rise to four stereocenters.
Catalysts 12 00101 sch003
Scheme 4. Asymmetric multicomponent cascade process for the synthesis of tetrahydronaphthols.
Scheme 4. Asymmetric multicomponent cascade process for the synthesis of tetrahydronaphthols.
Catalysts 12 00101 sch004
Scheme 5. Mechanistic proposal to explain the multicomponent cascade process.
Scheme 5. Mechanistic proposal to explain the multicomponent cascade process.
Catalysts 12 00101 sch005
Scheme 6. Synthesis of derivates of interest related with API (active pharmaceutical ingredients) and natural products.
Scheme 6. Synthesis of derivates of interest related with API (active pharmaceutical ingredients) and natural products.
Catalysts 12 00101 sch006
Figure 2. Selected biologically active compounds synthesized using organocatalytic key steps: Serriconin [83], Convolutamydine A [84], (–)-Botryodiplodin [85], (–)-Oseltamivir [86], DPC 083 [87], Telcagepant [88] and Leucomidine B [89]. The bond formed during the organocatalytic step is highlighted in green.
Figure 2. Selected biologically active compounds synthesized using organocatalytic key steps: Serriconin [83], Convolutamydine A [84], (–)-Botryodiplodin [85], (–)-Oseltamivir [86], DPC 083 [87], Telcagepant [88] and Leucomidine B [89]. The bond formed during the organocatalytic step is highlighted in green.
Catalysts 12 00101 g002
Figure 3. Biologically active compounds synthesized by our group using organocatalytic approaches. The bond formed during the organocatalytic step is highlighted in green.
Figure 3. Biologically active compounds synthesized by our group using organocatalytic approaches. The bond formed during the organocatalytic step is highlighted in green.
Catalysts 12 00101 g003
Figure 4. Antitumor squaramides and 1,4-dihydropyriridines synthesized by our group.
Figure 4. Antitumor squaramides and 1,4-dihydropyriridines synthesized by our group.
Catalysts 12 00101 g004
Figure 5. Catalytic structures proposed as trifunctional organocatalysts: X [30], XI [108], XII [109], XIII [110] and XIV [111].
Figure 5. Catalytic structures proposed as trifunctional organocatalysts: X [30], XI [108], XII [109], XIII [110] and XIV [111].
Catalysts 12 00101 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ardevines, S.; Marqués-López, E.; Herrera, R.P. Horizons in Asymmetric Organocatalysis: En Route to the Sustainability and New Applications. Catalysts 2022, 12, 101. https://doi.org/10.3390/catal12010101

AMA Style

Ardevines S, Marqués-López E, Herrera RP. Horizons in Asymmetric Organocatalysis: En Route to the Sustainability and New Applications. Catalysts. 2022; 12(1):101. https://doi.org/10.3390/catal12010101

Chicago/Turabian Style

Ardevines, Sandra, Eugenia Marqués-López, and Raquel P. Herrera. 2022. "Horizons in Asymmetric Organocatalysis: En Route to the Sustainability and New Applications" Catalysts 12, no. 1: 101. https://doi.org/10.3390/catal12010101

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