Next Article in Journal
New Programmable CMOS Fuzzifier and C2V Circuits Applicable in FLC Chip for Signal Processing of MEMS Glucose Sensors
Previous Article in Journal
An Investigation of the Thermal Expansion Coefficient for Resin Concrete with ZrW2O8
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthetic Applications of the Parkins Nitrile Hydration Catalyst [PtH{(PMe2O)2H}(PMe2OH)]: A Review

Laboratory of Organometallic Compounds and Catalysis (CSIC Associated Unit), Innovation Center in Advanced Chemistry (ORFEO-CINQA), Department of Organic and Inorganic Chemistry, Institute of Organometallic Chemistry "Enrique Moles", Faculty of Chemistry, University of Oviedo, Oviedo E-33006, Spain
Appl. Sci. 2015, 5(3), 380-401; https://doi.org/10.3390/app5030380
Submission received: 11 July 2015 / Revised: 6 August 2015 / Accepted: 13 August 2015 / Published: 18 August 2015

Abstract

:
The air-stable hydride-platinum(II) complex [PtH{(PMe2O)2H}(PMe2OH)], reported by Parkins and co-workers in 1995, is the most versatile catalyst currently available for the hydration of C≡N bonds. It features remarkable activity under relatively mild conditions and exceptionally high functional group compatibility, facts that have allowed the implementation of this complex in the synthesis of a large number of structurally complex, biologically active molecules and natural products. In this contribution, synthetic applications of the Parkins catalyst are reviewed.

Graphical Abstract

1. Introduction

The selective transformation of a functional group within a complex molecule to generate target compounds is a very important issue in organic synthesis. Although this is not a trivial task, the use of homogeneous metal catalysts allows, in many cases, to reach compounds with remarkable levels of selectivity. Indeed, the field of homogeneous catalysis is currently so widespread that, in the literature, it is difficult to find advanced syntheses not using metal catalysts at some stage [1,2,3]. One of the organic transformations in which metal catalysts have played a key role is the hydration of nitriles (Scheme 1), a process of great relevance in view of the broad industrial, synthetic, biological, and pharmacological applications of corresponding primary amide products [4,5,6].
Scheme 1. The catalytic hydration of nitriles.
Scheme 1. The catalytic hydration of nitriles.
Applsci 05 00380 g007
The conversion of nitriles into amides by conventional acid/base strategies usually requires harsh conditions and remains impractical for most purposes due to yield and selectivity issues, since these classical methods are frequently unable to control the over-hydrolysis of the primary amide product [7,8]. The use of transition metal complexes (mainly of Groups 8–12) as catalysts under neutral and mild conditions has allowed researchers to overcome these traditional barriers, and a large number of selective catalysts for the nitrile hydration process have seen the light during the last two decades [9,10,11,12,13,14,15]. Among them, the commercially available hydride-platinum(II) complex [PtH{(PMe2O)2H}(PMe2OH)] (1), first described by Ghaffar and Parkins in 1995 [16,17], has achieved great success among organic chemists because of its outstanding activity and selectivity and its exquisite compatibility with other functional groups (Figure 1). As a matter of fact, this platinum complex has found applications in the synthesis of a huge number of natural products and biologically active molecules of elaborate structure, making it the most widely used catalyst to date for the hydration of C≡N bonds.
Figure 1. Structure of the nitrile hydration catalyst [PtH{(PMe2O)2H}(PMe2OH)] (1).
Figure 1. Structure of the nitrile hydration catalyst [PtH{(PMe2O)2H}(PMe2OH)] (1).
Applsci 05 00380 g001
This article intends to provide a comprehensive account on the chemistry and synthetic applications of [PtH{(PMe2O)2H}(PMe2OH)] (1). The literature and patented works published up to July 2015 are covered.

2. Preparation of Complex [PtH{(PMe2O)2H}(PMe2OH)], First Catalytic Studies and Mechanism of Action

The synthesis of complex [PtH{(PMe2O)2H}(PMe2OH)] (1) involves the reaction of the platinum(0) precursor [Pt(PPh3)4] with an excess (five equiv.) of dimethylphosphine oxide Me2P(=O)H, which tautomerizes in situ into the phosphinite ligand Me2POH, in toluene under inert atmosphere. The reaction proceeds at room temperature, and the complex precipitates directly in the medium, thus making its isolation easy by simple filtration (isolated yields are usually in the range 75%–80%) [16,17,18,19]. In their first studies, Parkins and Ghaffar demonstrated the ability of 1 to selectively hydrate different model nitrile substrates, such as benzonitrile, acetonitrile, acrylonitrile, or 3-cyanopyridine, employing water, aqueous ethanol, or aqueous THF as the reaction medium at 70–90 °C [16,17,18]. The remarkable activity of [PtH{(PMe2O)2H}(PMe2OH)] (1) was readily evidenced in these seminal works, with reported turnover frequencies exceeding 1400 h−1 and turnover numbers of up to 77,000. Remarkably, contrary to other metal catalysts reported in the literature, complex 1 was able to operate under neutral conditions. The studies by Parkins and Ghaffar also showed that the size of the substituents on the phosphorus ligand is key to achieving good catalytic activity, the performances of the related complex [PtH{(PPh2O)2H}(PPh2OH)] containing a bulkier phosphinito ligand being significantly lower in comparison to those of 1 (an additional example has been recently described [20]). Further evidence of the synthetic potential of complex [PtH{(PMe2O)2H}(PMe2OH)] (1) was given by de Vries and co-workers, who reported excellent yields and chemoselectivities in the hydration of a variety of sterically hindered tertiary nitriles and nitriles containing acid- or base-sensitive functional groups (some representative examples are shown in Figure 2) [21].
Figure 2. Examples of nitriles that complex 1 is able to hydrate.
Figure 2. Examples of nitriles that complex 1 is able to hydrate.
Applsci 05 00380 g002
Concerning its mechanism of action, Ghaffar and Parkins proposed that in aqueous medium, protonation of the hydride ligand of 1 by water readily takes place, generating a vacant coordination site on the metal for substrate binding, with concomitant dihydrogen extrusion (Scheme 2) [16,17,18].
Scheme 2. Proposed mechanism for the catalytic hydration of nitriles with complex 1.
Scheme 2. Proposed mechanism for the catalytic hydration of nitriles with complex 1.
Applsci 05 00380 g008
Then, an intramolecular nucleophilic attack of the OH group of the PMe2OH ligand on the coordinated nitrile takes place, affording the five-membered metallacyclic intermediate A. Further hydrolysis of this metallacycle generates the amide product and regenerates the active platinum species [16,17,18].

3. Application of Complex [PtH{(PMe2O)2H}(PMe2OH)] in the Total Synthesis of Natural Products

Capuramycin (5), a nucleoside antibiotic isolated from the culture filtrate of Streptomyces griseus 446-S3 [22], and its congeners are considered to be important lead molecules for the development of new drugs for Mycobacterium tuberculosis infections [23]. Accordingly, remarkable efforts have been devoted to establish convergent synthetic routes to access 5 that allow the preparation of analogues for SAR (structure-activity relationship) studies [24]. In this context, in 2009, Kurosu and co-workers described a concise synthesis of capuramycin (5), starting from the partially protected uridine 2 (15 steps), in which the primary amide unit was conveniently generated by the hydration of the C≡N unit of intermediate 3 using [PtH{(PMe2O)2H}(PMe2OH)] (1) (Scheme 3) [25]. The hydration reaction, performed in aqueous ethanol at 70 °C, with 5 mol% of 1, allowed the quantitative conversion of 3 to the corresponding amide 4.
Scheme 3. The Parkins catalyst in the synthesis of capuramycin.
Scheme 3. The Parkins catalyst in the synthesis of capuramycin.
Applsci 05 00380 g009
(−)-Huperzine A (9) is a tricyclic alkaloid produced by the Chinese herb Hupercia serrata [26], with intense contemporary interest in clinical application for treating neurodegenerative diseases [27,28,29]. Several total syntheses of 9 have been reported [27,28,29]. In one of the most recent and scalable (eight steps starting from (R)-4-methyl-cyclohex-2-ene-1-one (6); 35%–45% overall yield), complex 1 was employed to promote the high-yield conversion of advanced nitrile intermediate 7 into amide 8 in aqueous ethanol (Scheme 4) [30]. From 8, Hoffmann's rearrangement of the amide group by means of [bis(trifluoroacetoxy)iodo]benzene (PIFA), followed by in situ deprotection of the pyridone unit with iodotrimethylsilane (TMSI), delivered the desired alkaloid.
Scheme 4. The Parkins catalyst in the synthesis of (−)-huperzine A.
Scheme 4. The Parkins catalyst in the synthesis of (−)-huperzine A.
Applsci 05 00380 g010
The selective hydration of nitrile 10 to generate primary amide 11, an advanced intermediate for the Nicolaou’s synthesis of the marine natural product diazonamide A (12) [31,32], using [PtH{(PMe2O)2H}(PMe2OH)] (1) was described by Sammakia and co-workers (Scheme 5) [33]. The reaction proceeded cleanly in aqueous ethanol at 120 °C with a catalyst loading of only 2 mol%, delivering 11 in 92% isolated yield.
Scheme 5. Synthesis of an amide intermediate for the preparation of diazonamide A.
Scheme 5. Synthesis of an amide intermediate for the preparation of diazonamide A.
Applsci 05 00380 g011
The welwitindolinones are a fascinating group of structurally unique metabolites isolated from blue-green algae that contain 3,4-bridged or spiro-cyclobutaneoxindole moieties [34,35]. Due to their challenging structures and their broad spectrum of biological activities, these compounds have attracted the attention of synthetic chemists for some time, although their complexity has prevented the completion of total syntheses for most of them thus far [36]. In an approach to welwitindolinone C isothiocyanate (welwistatin 17), Funk and co-workers have described the preparation of the related indole isocyanate derivative 16 starting from the vinylcyclohexanone 13 after 13 steps (Scheme 6) [37]. In this synthesis, the Parkins catalyst was employed to hydrate the encumbered nitrile 14. A high loading of complex 1 (42 mol%) was, in this case, required to generate the bridgehead primary amide 15 in high yield (73%). In the presence of Pb(OAc)4, amide 15 evolved into the final isocyanate 16 through the Baumgarten variant of the Hoffmann rearrangement.
Scheme 6. Funk’s approach to the synthesis of welwistatin.
Scheme 6. Funk’s approach to the synthesis of welwistatin.
Applsci 05 00380 g012
In 2009, Jones and Krische developed a total synthesis of (+)-geniposide (21), an iridoid β-glucoside of natural origin with anti-tumor and anti-inflammatory activity, starting from pivalate-protected lactol 18 (Scheme 7) [38]. During this synthesis the conversion of nitrile intermediate 19 into amide 20 was successfully accomplished in 87% yield using 20 mol% of the platinum complex [PtH{(PMe2O)2H}(PMe2OH)] (1).
Scheme 7. The hydration step in total synthesis of (+)-geniposide.
Scheme 7. The hydration step in total synthesis of (+)-geniposide.
Applsci 05 00380 g013
SB-203207 (25), an alkaloid isolated from a Streptomices species by researchers in the SmithKline Beecham group, was found to show marked isoleucyl tRNA synthetase inhibition activity [39,40]. Very recently, Kan and co-workers described a total synthesis of this compound starting from lactam 22, in which the primary amide unit was generated by catalytic hydration of nitrile intermediate 23 with complex 1 (Scheme 8) [41]. As in the two previous cases, a high loading of 1 was needed to obtain the corresponding amide product 24 in high yield. SB-203207 can also be accessed by acylation of the sulfonamide unit of naturally occurring altemicidin (26 in Scheme 8) [42]. In this context, the stereoselective synthesis of a key intermediate for altemicidin, i.e. amide 27 (see Scheme 8), was accomplished by Kan and Fukuyama employing [PtH{(PMe2O)2H}(PMe2OH)] (1) [43]. Compound 27 was obtained in quantitative yield by the hydration of the corresponding nitrile in aqueous ethanol at 75 °C for 12 h with 28 mol% of 1.
Scheme 8. The Parkins catalyst in the synthesis of alkaloid SB-203207.
Scheme 8. The Parkins catalyst in the synthesis of alkaloid SB-203207.
Applsci 05 00380 g014
In a synthetic approach toward the cytotoxic indole alkaloid avrainvillamide (30) and its dimeric form stephacidin B (31) [44], Myers and co-workers used [PtH{(PMe2O)2H}(PMe2OH)] (1) to synthesize primary amide intermediate 29 by the hydration of nitrile 28 (Scheme 9) [45,46]. The reaction proceeded cleanly in aqueous ethanol at 70 °C and with 20 mol% of 1, affording 29 in 85% isolated yield.
Scheme 9. The Parkins catalyst in the synthesis of avrainvillamide and stephacidin B.
Scheme 9. The Parkins catalyst in the synthesis of avrainvillamide and stephacidin B.
Applsci 05 00380 g015
The manzacidins are a family of marine natural products featuring a tetrahydropyrimidine carboxylic acid and an ester-linked bromopyrrole unit. This class of alkaloids has attracted significant attention due to their biological activities and unusual structure. Consequently, many groups have described the enantio- and diastereoselective synthesis of some of them [47]. In this context, Deng and co-workers reported synthetic routes to manzacidin A (34) [48,49] and manzacidin C (37) [50] involving a nitrile hydration step promoted by complex [PtH{(PMe2O)2H}(PMe2OH)] (1) (Scheme 10). In particular, 1 was used to transform the diasteromeric nitriles 32 and 35 into the corresponding primary amide intermediates 33 and 36, respectively. As shown in the scheme, almost quantitative yields were achieved in both cases, performing the reactions at 80 °C with low catalyst loading (5 mol%).
Scheme 10. Hydration steps in the syntheses of manzacidins developed by Deng.
Scheme 10. Hydration steps in the syntheses of manzacidins developed by Deng.
Applsci 05 00380 g016
In the context of their synthetic studies on the icetaxane family of diterpenoids, the hydration of nitrile 38 by complex 1 was successfully employed by Sarpong and co-workers to prepare primary amide 39, from which the formal synthesis of icetexone (40) and epi-icetexone (41) could be completed (Scheme 11) [51,52].
Scheme 11. Structure of icetexone and epi-icetexone, and synthesis of one intermediate.
Scheme 11. Structure of icetexone and epi-icetexone, and synthesis of one intermediate.
Applsci 05 00380 g017
The marine natural product psymberin (42 in Figure 3) has attracted a great deal of attention since its discovery in 2004 due to its complex architecture, biological properties, and paucity in nature. This interest has resulted in a large number of synthetic investigations and the development of analogous molecules [53]. The Parkins catalyst [PtH{(PMe2O)2H}(PMe2OH)] (1) has been involved in some of these studies [54,55,56,57,58,59]. In particular, in the first total synthesis of this compound, De Brabender and co-workers employed 1 to synthesize advanced amide intermediate 43 from the corresponding nitrile [54,55,57]. The same group also described the synthesis of different psymberin analogues from primary amide 44, which was also accessed though a nitrile hydration reaction promoted by 1 [56,57].
Figure 3. Structure of psymberin and amides 4344.
Figure 3. Structure of psymberin and amides 4344.
Applsci 05 00380 g003
The total syntheses of the prenylated indole alkaloids citrinalin B (47) and cyclopiamine (48) has been recently accomplished for the first time by Mercado-Marin, García-Reinada, and co-workers (Scheme 12) [60]. In their work, the key carbamate intermediate 46 was generated through a two-step sequence from tricyclic nitrile 45 via initial hydration of the C≡N catalyzed by complex [PtH{(PMe2O)2H}(PMe2OH)] (1), followed by Hofmann rearrangement of the resulting carboxamide (not isolated) promoted by PIFA in MeOH.
Scheme 12. The Parkins catalyst in the total synthesis of citrinalin B and cyclopiamine B.
Scheme 12. The Parkins catalyst in the total synthesis of citrinalin B and cyclopiamine B.
Applsci 05 00380 g018
In an independent work, Sarpong and co-workers reported the preparation of the related 17-hydroxy-citrinalin B (50), as well as those of the alkaloids stephacidin A (51) and notoamide I (52), using carbamate 49 as a common intermediate (Figure 4) [61]. As in the previous case, formation of 49 was accomplished through a hydration/Hofmann rearrangement sequence from the corresponding nitrile, the hydration step being again promoted by the Parkins platinum catalyst 1.
Figure 4. Structure of the carbamate 49 and the alkaloids 5052.
Figure 4. Structure of the carbamate 49 and the alkaloids 5052.
Applsci 05 00380 g004
Finally, Fleming and co-workers reported the preparation of a series of morinol-type lignans 55 starting from α,β-unsaturated nitriles 53 (Scheme 13) [62]. The selective transformation of the alkenenitrile unit of 53 into the allylic alcohol one involved an initial hydration step, catalyzed by complex [PtH{(PMe2O)2H}(PMe2OH)] (1), followed by an acylation/reduction of the resulting amides 54.
Scheme 13. Synthetic route to morinol-type lignans.
Scheme 13. Synthetic route to morinol-type lignans.
Applsci 05 00380 g019

4. Application of Complex [PtH{(PMe2O)2H}(PMe2OH)] in the Synthesis of Other Biologically Active Molecules and Pharmaceutical Compounds

One of the first synthetic applications of complex [PtH{(PMe2O)2H}(PMe2OH)] (1) was in the preparation of atenolol (57), a drug belonging to the group of β-blockers which contains a primary amide group and is commonly employed in the treatment of hypertension [63]. In particular, 1 had no difficulty in promoting the final hydration step of the nitrile intermediate 56. Thus, as shown in Scheme 14, performing the reaction in a refluxing ethanol/water (1:1 v/v) mixture, and using a very low catalyst loading of 0.0001 mol%, atenolol could be selectively obtained in 93% yield [64].
Scheme 14. Application of the Parkins catalyst in the synthesis of atenolol.
Scheme 14. Application of the Parkins catalyst in the synthesis of atenolol.
Applsci 05 00380 g020
The steroidal compound eplerenone (60) is an orally-active aldosterone antagonist featuring a canrenone-type structure that is also used for the treatment of hypertension and heart failure [65,66]. The synthesis of 60 has been the subject of several patents. In one of them, the Parkins catalyst was successfully employed to transform the advanced intermediate 7α-cyano-9(11)Δ-canrenone 58 into 7α-carbamoyl-9(11)Δ-canrenone 59 (Scheme 15) [67]. Conversion of the amide unit of 59 to the corresponding carboxylic acid, followed by esterification, or direct alcoholysis of the amide to the ester, completed the synthesis of 60.
Antagonists of the kappa opioid receptor have great therapeutic potential for the treatment of a range of central nervous system diseases and disorders, including, for example, schizophrenia, depression, and anxiety [68,69,70]. Quite recently, 8-azabicyclo[3.2.1]octan-3-yloxy-benzamides, as exemplified by 61 in Figure 5, were identified by Brugel and co-workers as potent and selective kappa opioid receptor antagonists [71]. In an additional SAR study, the same authors synthesized a large number of related compounds, such as 62 and 63, in which the benzamide unit was generated by hydration of the corresponding nitrile with the Parkins catalyst [72]. Reactions were performed in an EtOH/water (6:1 v/v) mixture at 80 °C, but no details on the yields, metal loadings, and reaction times were given in the document. The preparation of the structurally related azabicyclooctane derivatives 64 using 1 was also described, and their receptor binding affinity for kappa, mu, and delta opioid receptors was evaluated [73].
Scheme 15. The Parkins catalyst in the synthesis of the aldosterone antagonist eplerenone.
Scheme 15. The Parkins catalyst in the synthesis of the aldosterone antagonist eplerenone.
Applsci 05 00380 g021
Figure 5. Structure of the opioid receptor antagonists 61-64.
Figure 5. Structure of the opioid receptor antagonists 61-64.
Applsci 05 00380 g005
In a patented work directed to the identification of novel histamine H3 receptor ligands, Brugel and co-workers also employed complex [PtH{(PMe2O)2H}(PMe2OH)] (1) to promote the high yield conversion of the racemic cyclopropyl-substituted benzamide 65 into primary amide 66 (Scheme 16) [74]. It is of note that when the same reaction was carried out employing an excess of KOH as the promoter, 66 was obtained in a much lower yield (69% vs. 97%).
Scheme 16. Catalytic synthesis of a cyclopropyl-substituted benzamide using 1.
Scheme 16. Catalytic synthesis of a cyclopropyl-substituted benzamide using 1.
Applsci 05 00380 g022
In addition, hydration with the Parkins catalyst 1 was also successfully employed to transform the acetate- and pivaloate-protected aminonitrile glycoside 67 and the morphinan derivative 69 into the corresponding primary amides 68 and 70, respectively (Scheme 17) [75,76].
Scheme 17. Catalytic hydration of the biologically active nitriles 67 and 69.
Scheme 17. Catalytic hydration of the biologically active nitriles 67 and 69.
Applsci 05 00380 g023

5. Other Synthetic Applications of Complex [PtH{(PMe2O)2H}(PMe2OH)]

The synthetic utility of the Parkins catalyst [PtH{(PMe2O)2H}(PMe2OH)] (1) is not restricted to the catalytic hydration of nitriles. Using this complex, other amide-bond forming reactions could be developed. In this context, de Vries and co-workers found that 1 efficiently catalyzes the direct conversion of unactivated nitriles to N-substituted amides by reaction with both primary and secondary amines in the presence of water (Scheme 18) [77]. A mechanism involving the initial formation of an amidine intermediate, which subsequently hydrolyzes into the amide, was proposed. Indeed, when the reactions were performed in the absence of water, the corresponding amidines could be isolated.
Scheme 18. Hydrolytic amidation of nitriles catalyzed by [PtH{(PMe2O)2H}(PMe2OH)] (1).
Scheme 18. Hydrolytic amidation of nitriles catalyzed by [PtH{(PMe2O)2H}(PMe2OH)] (1).
Applsci 05 00380 g024
An interesting extension of this amidation process is the quantitative formation of 2-ethyl-2-oxazoline 72 from the reaction of propionitrile with 2-aminoethanol via ring-closure of the intermediate amidine 71 (Scheme 19) [77].
Scheme 19. Catalytic synthesis of 2-ethyl-2-oxazoline.
Scheme 19. Catalytic synthesis of 2-ethyl-2-oxazoline.
Applsci 05 00380 g025
Although scarcely explored, the ability of complex 1 to promote the hydration of thiocyanates was also demonstrated [78,79]. As a representative example, treatment of the optically active thiocyanate 73 with 25 mol% of 1 in a THF/H2O mixture led to the high-yield formation of thiocarbamate 74 (Scheme 20). Remarkably, in marked contrast with the use of classical acidic hydrolysis conditions, which induced the complete racemization at the stereogenic carbonyl α-carbon, the reaction proceeded without loss of chirality.
Scheme 20. Catalytic hydration of an optically active thiocyanate.
Scheme 20. Catalytic hydration of an optically active thiocyanate.
Applsci 05 00380 g026
On the other hand, in the presence of catalytic amounts of [PtH{(PMe2O)2H}(PMe2OH)] (1), separable mixtures of 1-ethoxyisoquinolines 76 and isoquinolones 77 were synthesized by intramolecular 6-endo-dig cyclization of ortho-alkynylbenzonitriles 75 in ethanol (Scheme 21) [80]. In these reactions the Parkins catalyst activates not only the nitrile unit, but also the alkyne moiety of the substrates. The use of other alcohols, such as methanol, iso-propanol, or cyclopentanol, also led to mixtures of the corresponding 1-alkoxyisoquinolines and isoquinolones 77.
Scheme 21. Platinum-catalyzed intramolecular cyclization of ortho-alkynylbenzonitriles.
Scheme 21. Platinum-catalyzed intramolecular cyclization of ortho-alkynylbenzonitriles.
Applsci 05 00380 g027
In an independent patented work, a more selective transformation of ortho-alkynylbenzonitrile 78 was observed [81]. Thus, as shown in Scheme 22, the treatment of this substrate with 2 mol% of platinum complex 1 in refluxing methanol resulted in the exclusive formation of isoquinolone 79.
Scheme 22. Selective synthesis of isoquinolone 79 using complex 1.
Scheme 22. Selective synthesis of isoquinolone 79 using complex 1.
Applsci 05 00380 g028

6. Limitations

Despite the broad scope shown by complex [PtH{(PMe2O)2H}(PMe2OH)] (1), a major limitation was encountered when studying its application for the catalytic hydration of cyanohydrins, i.e. α-hydroxynitriles. Unfortunately, the hydration rates for these substrates were very low in comparison to those observed with other nitriles, and only a few turnovers could be achieved [82,83]. The poor reactivity observed was ascribed to the low stability of the cyanohydrin substrates, which slowly equilibrate in solution with HCN and the corresponding aldehyde or ketone (Scheme 23), and the irreversible coordination of the cyanide anion to platinum leads to the deactivation of the catalyst. In fact, most of the nitrile hydration catalysts reported to date are poisoned by cyanide and, therefore, with some notable exceptions [84,85,86,87,88,89], are inoperative with this particular class of nitriles.
Scheme 23. Equilibrium between a cyanohydrin and the corresponding aldehyde/ketone and hydrogen cyanide.
Scheme 23. Equilibrium between a cyanohydrin and the corresponding aldehyde/ketone and hydrogen cyanide.
Applsci 05 00380 g029
Although with only one example, Parkins and co-workers showed that this limitation can be circumvented, protecting the OH group of the cyanohydrin prior to the hydration reaction [90]. Thus, as shown in Scheme 24, mandelamide (83) was accessible from mandelonitrile (80) by the hydration of the corresponding mixed acetal 81 and the subsequent removal of the α-butoxyethoxy protecting group in the resulting amide 82 with dilute acid. In this context, we must also stress that the ability of complex 1 to promote the selective conversion of cyanohydrin acetates into α-acetoxy amides has also been demonstrated [91].
Scheme 24. Access to mandelamide from mandelonitrile using the Parkins catalyst.
Scheme 24. Access to mandelamide from mandelonitrile using the Parkins catalyst.
Applsci 05 00380 g030
In addition to cyanohydrins, other nitriles mentioned in the literature for which complex [PtH{(PMe2O)2H}(PMe2OH)] (1) proved ineffective were 84 and 85 (Figure 6). In the case of 84, the treatment with 1 in an EtOH/H2O mixture at 70 °C resulted in the decyanation of the substrate and generation of the corresponding aldimine [92]. For the malonodinitrile 85, although its monohydration using 1 was possible, the reaction yield was very low in comparison to that achieved employing the more classical KF-Al2O3 system [93].
Figure 6. Structure of the nitriles 84 and 85.
Figure 6. Structure of the nitriles 84 and 85.
Applsci 05 00380 g006

7. Conclusions

In summary, in this paper we have tried to summarize, in a comprehensive manner, all the synthetic processes in which the commercially available phosphinito-platinum(II) complex [PtH{(PMe2O)2H}(PMe2OH)] (1) has been involved to date. Although tens of nitrile hydration catalysts are currently known, none has achieved nearly the same success of compex 1. From the examples shown along the manuscript, it is easy to realize the extraordinary synthetic usefulness of this complex, which is capable of promoting the selective hydration of C≡N units in densely functionalized molecules of very high structural complexity due to its exceptionally high functional group compatibility. The high synthetic versatility of the primary amide −C(=O)NH2 function, along with its presence in a multitude of natural products and biological active molecules already known, and certainly in many others waiting to be discovered, allow us to presume an increase in the use of 1 in the near future. In short, the Parkins catalyst [PtH{(PMe2O)2H}(PMe2OH)] (1) is called to occupy a preferred position within the toolbox of synthetic organic chemists.

Acknowledgments

The author thanks the Ministerio de Economía y Competitividad (MINECO) of Spain (projects CTQ2013-40591-P and CTQ2014-51912-REDC) and the Gobierno del Principado de Asturias (project GRUPIN14-006) for financial support.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Beller, M.; Bolm, C. Transition Metals for Organic Synthesis, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2004. [Google Scholar]
  2. Hegedus, L.S.; Söderberg, B.C.G. Transition Metals in the Synthesis of Complex Organic Molecules, 3rd ed.; University Science Books: Sausalito, CA, USA, 2009. [Google Scholar]
  3. Crawley, M.L.; Trost, B.M. Applications of Transition Metal Catalysts in Drug Discovery and Development; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  4. Zabicky, J. The Chemistry of Amides; John Wiley & Sons: New York, NY, USA, 1970. [Google Scholar]
  5. Greenberg, A.; Breneman, C.M.; Liebman, J.F. The Amide Linkage: Structural Significance in Chemistry, Biochemistry and Materials Science; John Wiley & Sons: New York, NY, USA, 2000. [Google Scholar]
  6. Lanigan, R.M.; Sheppard, T.D. Recent developments in amide synthesis: Direct amidation of carboxylic acids and transamidation reactions. Eur. J. Org. Chem. 2013, 2013, 7453–7465. [Google Scholar] [CrossRef]
  7. Bailey, P.D.; Mills, T.J.; Pettecrew, R.; Price, R.A. Amides. In Comprehensive Organic Functional Group Transformations, 2nd ed.; Katritzky, A.R., Taylor, R.J.K., Eds.; Elsevier: Oxford, UK, 2005; Volume 5, pp. 201–294. [Google Scholar]
  8. Dopp, D.; Dopp, H. Methoden Organic Chemistry (Houben Weyl); Thieme Verlag: Stuttgart, Germany, 1985; Volume E5(2), pp. 1024–1031. [Google Scholar]
  9. Parkins, A.W. Catalytic hydration of nitriles to amides: Platinum containing catalyst offers new opportunity. Platinum Metals Rev. 1996, 40, 169–174. [Google Scholar]
  10. Kukushkin, V.Y.; Pombeiro, A.J.L. Additions to metal-activated organonitriles. Chem. Rev. 2002, 102, 1771–1802. [Google Scholar] [CrossRef] [PubMed]
  11. Kukushkin, V.Y.; Pombeiro, A.J.L. Metal-mediated and metal-catalyzed hydrolysis of nitriles. Inorg. Chim. Acta 2005, 358, 1–21. [Google Scholar] [CrossRef]
  12. Ahmed, T.J.; Knapp, S.M.M.; Tyler, D.R. Frontiers in catalytic nitrile hydration: Nitrile and cyanohydrins hydration catalyzed by homogeneous organometallic complexes. Coord. Chem. Rev. 2011, 255, 949–974. [Google Scholar] [CrossRef]
  13. García-Álvarez, R.; Crochet, P.; Cadierno, V. Metal-catalyzed amide bond forming reactions in an environmentally friendly aqueous medium: Nitrile hydrations and beyond. Green Chem. 2013, 15, 46–66. [Google Scholar] [CrossRef]
  14. García-Álvarez, R.; Francos, J.; Tomás-Mendivil, E.; Crochet, P.; Cadierno, V. Metal-catalyzed nitrile hydration reactions: The specific contribution of ruthenium. J. Organomet. Chem. 2014, 771, 93–104. [Google Scholar] [CrossRef]
  15. Downs, E.L.; Tyler, D.R. Nanoparticle catalysts for nitrile hydration. Coord. Chem. Rev. 2014, 280, 28–37. [Google Scholar] [CrossRef]
  16. Ghaffar, T.; Parkins, A.W. A new homogeneous platinum containing catalyst for the hydrolysis of nitriles. Tetrahedron Lett. 1995, 36, 8657–8660. [Google Scholar] [CrossRef]
  17. Parkins, A.W.; Ghaffar, T. Catalyst and process for preparing amides. PCT Int. Appl. WO96/30379, 1996. [Google Scholar]
  18. Ghaffar, T.; Parkins, A.W. The catalytic hydration of nitriles to amides using a homogeneous platinum phosphinito catalyst. J. Mol. Catal. A 2000, 160, 249–261. [Google Scholar] [CrossRef]
  19. Tafesse, T. Terapeutic agents useful for treating pain. PCT Int. Appl. WO2008/133973, 2008. [Google Scholar]
  20. Gulyás, H.; Rivilla, I.; Curreli, S.; Freixa, Z.; van Leeuwen, P.W.N.M. Highly active, chemo- and enantioselective Pt-SPO catalytic systems for the synthesis of aromatic carboxamides. Catal. Sci. Technol. 2015, 5, 3822–3828. [Google Scholar] [CrossRef]
  21. Jiang, X.-B.; Minnaard, A.J.; Feringa, B.L.; de Vries, J.G. Platinum-catalyzed selective hydration of hindered nitriles and nitriles with acid- and base-sensitive groups. J. Org. Chem. 2004, 69, 2327–2331. [Google Scholar] [CrossRef] [PubMed]
  22. Yamaguchi, H.; Sato, S.; Yoshida, S.; Takada, K.; Itoh, M.; Seto, H.; Otake, N. Capuramycin, a new nucleoside antibiotic: Taxonomy, fermentation, isolation and characterization. J. Antibiot. 1986, 39, 1047–1053. [Google Scholar] [CrossRef] [PubMed]
  23. Koga, T.; Fukuoka, T.; Doi, N.; Harasaki, T.; Inoue, H.; Hotoda, H.; Kakuta, M.; Muramatsu, Y.; Yamamura, N.; Hoshi, M.; Hirota, T. Activity of capuramycin analogues against Mycobacterium tuberculosis, Mycobacterium avium and Mycobacterium intracellulare in vitro and in vivo. J. Antimicro. Chemother. 2004, 54, 755–760. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Siricilla, S.; Aleiwi, B.A.; Kurosu, M. Improved synthesis of capuramycin and its analogues. Chem. Eur. J. 2013, 19, 13847–13858. [Google Scholar] [CrossRef] [PubMed]
  25. Kurosu, M.; Li, K.; Crick, D.C. Concise synthesis of capuramycin. Org. Lett. 2009, 11, 2393–2396. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, J.S.; Zhu, Y.L.; Yu, C.M.; Zhou, Y.Z.; Han, Y.Y.; Wu, F.W.; Qi, B.F. The structures of huperzine A and B, two new alkaloids exhibiting marked anticholinesterase activity. Can. J. Chem. 1986, 64, 837–839. [Google Scholar] [CrossRef]
  27. Kozikowski, A.P.; Tückmantel, W. Chemistry, pharmacology, and clinical efficacy of the Chinese nootropic agent huperzine A. Acc. Chem. Res. 1999, 32, 641–650. [Google Scholar] [CrossRef]
  28. Tun, M.K.M.; Herzon, S.B. The pharmacology and therapeutic potential of (−)-huperzine A. J. Exp. Pharmacol. 2012, 4, 113–123. [Google Scholar]
  29. Sohel, S.M.A.; Opatz, T. Synthetic approaches towards huperzine A and B. Arkivoc 2014, 1, 92–108. [Google Scholar] [CrossRef]
  30. Tun, M.K.M.; Wüstmann, D.J.; Herzon, S.B. A robust and scalable synthesis of the potent neuroprotective agent (−)-huperzine A. Chem. Sci. 2011, 2, 2251–2253. [Google Scholar] [CrossRef]
  31. Nicolaou, K.C.; Bella, M.; Chen, D.Y.-K.; Huang, X.; Ling, T.; Snyder, S.A. Total synthesis of diazonamide A. Angew. Chem. Int. Ed. 2004, 41, 3495–3499. [Google Scholar] [CrossRef]
  32. Nicolaou, K.C.; Chen, D.Y.-K.; Huang, X.; Ling, T.; Bella, M.; Snyder, S.A. Chemistry and biology of diazonamide A: First total synthesis and confirmation of the true structure. J. Am. Chem. Soc. 2004, 126, 12888–12896. [Google Scholar] [CrossRef] [PubMed]
  33. Mai, C.-K.; Sammons, M.F.; Sammakia, T. A concise formal synthesis of diazonamide A by the stereoselective construction of the C10 quaternary center. Angew. Chem. Int. Ed. 2010, 49, 2397–2400. [Google Scholar] [CrossRef] [PubMed]
  34. Stratmann, K.; Moore, R.E.; Bonjouklian, R.; Deeter, J.B.; Patterson, G.M.L.; Shaffer, S.; Smith, C.D.; Smitka, T.A. Welwitindolinones, unusual alkaloids from the blue-green algae Hapalosiphon welwitschii and Westiella intricate. Relationship to fischerindoles and hapalindoles. J. Am. Chem. Soc. 1994, 116, 9935–9942. [Google Scholar] [CrossRef]
  35. Jimenez, J.I.; Huber, U.; Moore, R.E.; Patterson, G.M.L. Oxidized welwitindolinones from terrestrial Fischerella spp. J. Nat. Prod. 1999, 62, 569–572. [Google Scholar] [CrossRef] [PubMed]
  36. Menéndez, J.C. Chemistry of the welwitindolinones. Top. Heterocycl. Chem. 2007, 11, 63–101. [Google Scholar]
  37. Greshock, T-J.; Funk, R.L. An approach to the total synthesis of welwistatin. Org. Lett. 2006, 8, 2643–2645. [Google Scholar]
  38. Jones, R.A.; Krische, M.J. Asymmetric total synthesis of the iridoid β-glucoside (+)-geniposide vis phosphine organocatalysis. Org. Lett. 2009, 11, 1849–1851. [Google Scholar] [CrossRef] [PubMed]
  39. Stefanska, A.L.; Cassels, R.; Ready, S.J.; Warr, S.R. SB-203207 and SB-203208, two novel isoleucyl tRNA synthetase inhibitors from a Streptomyces sp. I. Fermentation, isolation and properties. J. Antibiot. 2000, 53, 357–363. [Google Scholar] [CrossRef] [PubMed]
  40. Houge-Frydrych, C.S.V.; Gilpin, M.L.; Skett, P.W.; Tyler, J.W.S.R. SB-203207 and SB-203208, two novel isoleucyl tRNA synthetase inhibitors from a Streptomyces sp. II. Structure determination. J. Antibiot. 2000, 53, 364–372. [Google Scholar] [CrossRef] [PubMed]
  41. Hirooka, Y.; Ikeuchi, K.; Kawamoto, Y.; Akao, Y.; Furuta, T.; Asakawa, T.; Inai, M.; Wakimoto, T.; Fukuyama, T.; Kan, T. Enantioselective synthesis of SB-203207. Org. Lett. 2014, 16, 1646–1649. [Google Scholar] [CrossRef] [PubMed]
  42. Banwell, M.G.; Castro, C.F.; Easton, C.J.; Forrest, A.K.; Karoli, T.; March, D.R.; Mensah, L.; Nairn, M.R.; O´Hanlon, P.J.; Oldham, M.D.; Yue, W. Analogues of SB-203207 as inhibitors of tRNA synthetases. Bioorg. Med. Chem. Lett. 2000, 10, 2263–2266. [Google Scholar] [CrossRef]
  43. Kan, T.; Kawamoto, Y.; Asakawa, T.; Furuta, T.; Fukuyama, T. Synthetic studies on altemicidin: Stereocontrolled construction of the core framework. Org. Lett. 2008, 10, 169–171. [Google Scholar] [CrossRef] [PubMed]
  44. Nising, C.F. Recent synthetic approaches towards the antiproliferative natural products avrainvillamide and stephacidin B. Chem. Soc. Rev. 2010, 39, 591–599. [Google Scholar] [CrossRef] [PubMed]
  45. Herzon, S.B.; Myers, A.G. Enantioselective synthesis of stephacidin B. J. Am. Chem. Soc. 2005, 127, 5342–5344. [Google Scholar] [CrossRef] [PubMed]
  46. Myers, A.G.; Herzon, S.B.; Wulff, J.E.; Siegrist, R.; Svenda, J.; Zajac, M.A. Synthesis of avrainvillamide, stephacidin B, and analogues thereof. PCT Int. Appl. WO2006/102097, 2006. [Google Scholar]
  47. Ohfune, Y.; Oe, K.; Namba, K.; Shinada, T. Total synthesis of manzacidins. An overview and perspective. Heterocycles 2012, 85, 2617–2649. [Google Scholar] [CrossRef]
  48. Wang, Y.; Liu, X.; Deng, L. Dual-function cinchona alkaloid catalysis: Catalytic Asymmetric tandem conjugate addition-protonation for the direct creation of nonadjacent sterocenters. J. Am. Chem. Soc. 2006, 128, 3928–3930. [Google Scholar] [CrossRef] [PubMed]
  49. Deng, L.; Li, H.; Wang, Y.; Wu, F. Asymmetric carbon-carbon-bond-forming reactions catalyzed by bifunctional cinchona alkaloids. US Pat. Appl. US2007/0083049, 2007. [Google Scholar]
  50. Wang, B.; Wu, F.; Wang, Y.; Liu, X.; Deng, L. Control of diasteroselectivity in tandem asymmetric reactions generating nonadjacent stereocenters with bifunctional catalysis by cinchona alkaloids. J. Am. Chem. Soc. 2007, 129, 768–769. [Google Scholar] [CrossRef] [PubMed]
  51. De Jesus Cortez, F.; Sarpong, R. Ga(III)-catalyzed cycloisomerization approach to (±)-icetexone and (±)-epi-icetexone. Org. Lett. 2010, 12, 1428–1431. [Google Scholar] [CrossRef] [PubMed]
  52. De Jesus Cortez, F.; Lapointe, D.; Hamlin, A.M.; Simmons, E.M.; Sarpong, R. Synthetic studies on the icetexones: Enantioselective formal synthesis of icetexone and epi-icetexone. Tetrahedron 2013, 69, 5665–5676. [Google Scholar] [CrossRef] [PubMed]
  53. Bielitza, M.; Pietruszka, J. The psymberin story―Biological properties and approaches towards total and analogue syntheses. Angew. Chem. Int. Ed. 2013, 52, 10960–10985. [Google Scholar] [CrossRef] [PubMed]
  54. Jiang, X.; García-Fortanet, J.; De Brabander, J.K. Synthesis and complete stereochemical assignment of psymberin/irciniastatin A. J. Am. Chem. Soc. 2005, 127, 11254–11255. [Google Scholar] [CrossRef] [PubMed]
  55. De Brabander, J.K.; Jiang, X. Synthesis and complete stereochemical assignment of psymberin/irciniastatin for use as antitumor compounds. US Pat. Appl. US2007/0015821, 2007. [Google Scholar]
  56. Jiang, X.; Williams, N.; De Brabander, J.K. Synthesis of psymberin analogues: Probing a functional correlation with the pederin/mycalamide family of natural products. Org. Lett. 2007, 9, 227–230. [Google Scholar] [CrossRef] [PubMed]
  57. Feng, Y.; Jiang, X.; De Brabander, J.K. Studies toward the unique pederin famility member psymberin: Full structure elucidation, two alternative total syntheses, and analogs. J. Am. Chem. Soc. 2012, 134, 17083–17093. [Google Scholar] [CrossRef] [PubMed]
  58. Rech, J.C.; Floreancig, P.E. Concise and stereoselective synthesis of the N7-C25 fragment of psymberin. Org. Lett. 2005, 7, 5175–5178. [Google Scholar] [CrossRef] [PubMed]
  59. Brown, L.E.; Landaverry, Y.R.; Davies, J.R.; Milinkevich, K.A.; Ast, S.; Carlson, J.S.; Oliver, A.G.; Konopelski, J.P. Progress toward the total synthesis of psymberin/irciniastatin A. J. Org. Chem. 2009, 74, 5405–5410. [Google Scholar] [CrossRef] [PubMed]
  60. Mercado-Marin, E.V.; García-Reynaga, P.; Romminger, S.; Pimenta, E.F.; Romeny, D.K.; Lodewyk, M.W.; Williams, D.E.; Andersen, R.J.; Miller, S.J.; Tantillo, D.J.; Berlinck, R.G.S.; Sarpong, R. Total synthesis and isolation of citrinalin and cyclopiamine congeners. Nature 2014, 509, 318–324. [Google Scholar] [CrossRef] [PubMed]
  61. Mercado-Marin, E.V.; Sarpong, R. Unified approach to prenylated indole alkaloids: Total synthesis of (−)-17-hydroxy-citrinalin B, (+)-stephacidin A, and (+)-notoamide I. Chem. Sci. 2015, 6, 5048–5052. [Google Scholar] [CrossRef]
  62. Yao, L.; Pitta, B.; Ravikumar, P.C.; Purzycki, M.; Fleming, F.F. Transmissive olefination route to putative “morinol I” lignans. J. Org. Chem. 2012, 77, 3651–3657. [Google Scholar] [CrossRef] [PubMed]
  63. Ripley, T.L.; Saseen, J.J. β-Blockers: A review of their pharmacological and physiological diversity in hypertension. Ann. Pharmacother. 2014, 48, 723–733. [Google Scholar] [CrossRef] [PubMed]
  64. Akisanya, J.; Parkins, A.W.; Steed, J.W. A synthesis of atenolol using a nitrile hydration catalyst. Org. Process Res. Dev. 1998, 2, 274–276. [Google Scholar] [CrossRef]
  65. Craft, J. Eplerenone (Inspra), a new aldosterone antagonist for the treatment of systemic hypertension and heart failure. BUMC Proceedings 2004, 17, 217–220. [Google Scholar] [PubMed]
  66. Danjuma, M.I.; Mukherjee, I.; Makaronidis, J.; Osula, S. Converging indications of aldosterone antagonists (spironolactone and eplerenone): A narrative review of safety profiles. Curr. Hypertens. Rep. 2014, 16, 1–10. [Google Scholar] [CrossRef] [PubMed]
  67. Gieling, R.G.; Groen, M.B. Process for making steroidal compounds. PCT Int. Appl. WO2006/097342, 2006. [Google Scholar]
  68. Metcalf, M.D.; Coop, A. Kappa opioid antagonists: Past, successes and future prospects. AAPS J. 2005, 7, E704–E722. [Google Scholar] [CrossRef] [PubMed]
  69. Carroll, F.I.; Carlezon, W.A., Jr. Development of κ opioid receptor antagonists. J. Med. Chem. 2013, 56, 2178–2195. [Google Scholar] [CrossRef] [PubMed]
  70. Urbano, M.; Guerrero, M.; Rosen, H.; Roberts, E. Antagonists of the kappa opioid receptor. Bioorg. Med. Chem. Lett. 2014, 24, 2021–2032. [Google Scholar] [CrossRef] [PubMed]
  71. Brugel, T.A.; Smith, R.W.; Balestra, M.; Becker, C.; Daniels, T.; Hoerter, T.N.; Koether, G.M.; Throner, S.R.; Panko, L.M.; Folmer, J.J.; et al. Discovery of 8-azabicyclo[3.2.1]octan-3-yloxy-benzamides as selective antagonists of the kappa opioid receptor. Bioorg. Med. Chem. Lett. 2010, 20, 5847–5862. [Google Scholar] [CrossRef] [PubMed]
  72. Brugel, T.A.; Smith, R.W.; Balestra, M.; Becker, C.; Daniels, T.; Koether, G.M.; Throner, S.R.; Panko, L.M.; Brown, D.G.; Liu, R.; et al. SAR development of a series of 8-azabicyclo[3.2.1]octan-3-yloxy-benzamides as kappa opioid receptor antagonists. Bioorg. Med. Chem. Lett. 2010, 20, 5405–5410. [Google Scholar] [CrossRef] [PubMed]
  73. Saito, D.R.; Long, D.D.; Van Dyke, P.; Church, T.J.; Jiang, L.; Frieman, B. 8-Azabicyclo[3.2.1]octyl-2-hydroxybenzamide compounds as mu opioid receptor antagonists. PCT Int. Appl. WO2009/029256, 2009. [Google Scholar]
  74. Arnold, J.; Brugel, T.A.; Edwards, P.; Griffin, A.; Groblewski, T.; Labrecque, D.; Throner, S.; Wesolowski, S. Cyclopropyl amide derivatives ´978. PCT Int. Appl. WO2009/024823, 2009. [Google Scholar]
  75. Andrews, R.S.; Becker, J.J.; Gagné, M.R. A photoflow reactor for the continuous photoredox-mediated synthesis of C-glycoaminoacids and C-glycolipids. Angew. Chem. Int. Ed. 2012, 51, 4140–4143. [Google Scholar] [CrossRef] [PubMed]
  76. Park, J.H.; Tafesse, L. Nitrogen containing morphinan derivatives and the use thereof. PCT Int. Appl. WO2014/091298, 2014. [Google Scholar]
  77. Cobley, C.J.; van den Heuvel, M.; Abbadi, A.; de Vries, J.G. Platinum catalysed hydrolytic amidation of unactivated nitriles. Tetrahedron Lett. 2000, 41, 2467–2470. [Google Scholar] [CrossRef]
  78. Falck, J.R.; Gao, S.; Prasad, R.N.; Koduru, S.R. Electrophilic α-thiocyanation of chiral and achiral N-acyl amides. A convenient route to 5-substituted and 5,5-disubstituted 2,4-thiazolidinediones. Bioorg. Med. Chem. Lett. 2008, 18, 1768–1771. [Google Scholar] [CrossRef] [PubMed]
  79. Finch, H.; Fox, C.; Sajad, M. Respiratory disease treatment. PCT Int. Appl. WO2010/015818, 2010. [Google Scholar]
  80. Li, J.; Chen, L.; Chin, E.; Lui, A.S.; Zecic, H. Platinum(II)-catalyzed intramolecular cyclization of alkynylbenzontriles: Synthesis of 1-alkoxyisoquinolines and isoquinolones. Tetrahedron Lett. 2010, 51, 6422–6425. [Google Scholar] [CrossRef]
  81. Brameld, K.A.; Carter, D.S.; Chin, E.; De Vicente Fidalgo, J.; Li, J.; Schoenfeld, R.C.; Sjogren, E.B.; Talamas, F.X. Heterocyclic antiviral compounds. PCT Int. Appl. WO2010/010017, 2010. [Google Scholar]
  82. Ahmed, T.J.; Fox, B.R.; Knapp, S.M.M.; Yelle, R.B.; Juliette, J.J.; Tyler, D.R. Investigation of the reactivity of Pt phosphinito and molybdocene nitrile hydration catalysts with cyanohydrins. Inorg. Chem. 2009, 48, 7828–7837. [Google Scholar] [CrossRef] [PubMed]
  83. Knapp, S.M.M.; Sherbow, T.J.; Ahmed, T.J.; Thiel, I.; Zakharov, L.N.; Juliette, J.J.; Tyler, D.R. Platinum phosphinito catalysts for nitrile hydration. J. Inorg. Organomet. Polym. 2014, 24, 145–156. [Google Scholar] [CrossRef]
  84. Knapp, S.M.M.; Sherbow, T.J.; Juliette, J.J.; Tyler, D.R. Cyanohydrin hydration with [Ru(η6-p-cymene)Cl2PR3] complexes. Organometallics 2012, 31, 2941–2944. [Google Scholar] [CrossRef]
  85. Knapp, S.M.M.; Sherbow, T.J.; Yelle, R.B.; Zakharov, L.N.; Juliette, J.J.; Tyler, D.R. Mechanistic investigations and secondary coordination sphere effects in the hydration of nitriles with [Ru(η6-arene)Cl2PR3] complexes. Organometallics 2013, 32, 824–834. [Google Scholar] [CrossRef]
  86. Knapp, S.M.M.; Sherbow, T.J.; Yelle, R.B.; Juliette, J.J.; Tyler, D.R. Catalytic nitrile hydration with [Ru(η6-p-cymene)Cl2(PR2R´)] complexes: Secondary coordination sphere effects with phosphine oxide and phosphinite ligands. Organometallics 2013, 32, 3744–3752. [Google Scholar] [CrossRef]
  87. Tomás-Mendivil, E.; Suárez, F.J.; Díez, J.; Cadierno, V. An efficient ruthenium(IV) catalyst for the selective hydration of nitriles to amides in water under mild conditions. Chem. Commun. 2014, 50, 9661–9664. [Google Scholar] [CrossRef] [PubMed]
  88. Sherbow, T.J.; Downs, E.L.; Sayler, R.I.; Razink, J.J.; Juliette, J.J.; Tyler, D.R. Investigation of 1,3,5-triaza-7-phosphaadamantane-stabilized silver nanoparticles as catalysts for the hydration of benzonitriles and acetone cyanohydrin. ACS Catal. 2014, 4, 3096–3104. [Google Scholar] [CrossRef]
  89. Downs, E.L.; Tyler, D.R. Nitrile and cyanohydrins hydration with nanoparticles formed in situ from a platinum dihydride complex. J. Inorg. Organomet. Polym. 2015, 25, 73–80. [Google Scholar] [CrossRef]
  90. Papakyprianou, A.; Parkins, A.W.; Prince, P.D.; Steed, J.W. Synthesis of O-protected mandelamide and its elaboration to 5-hydroxy-1,5-diphenylhydantoin. Org. Prep. Proced. Int. 2002, 34, 436–440. [Google Scholar] [CrossRef]
  91. North, M.; Parkins, A.W.; Shariff, A.N. Catalytic asymmetric synthesis of α-acetoxy amides. Tetrahedron Lett. 2004, 45, 7625–7627. [Google Scholar] [CrossRef]
  92. Trost, B.M.; Xie, J.; Sieber, J.D. The palladium catalyzed asymmetric addition of oxindoles and allenes: An atom-economical versatile method for the construction of chiral indole alkaloids. J. Am. Chem. Soc. 2011, 133, 20611–20622. [Google Scholar] [CrossRef] [PubMed]
  93. Grigg, R.; Hasakunpaisarn, A.; Kilner, C.; Kongkathip, B.; Kongkathip, N.; Pettman, A.; Sridharan, V. Catalytic processes for the functionalisation and desymmetrisation of malononitrile derivatives. Tetrahedron 2005, 61, 9356–9367. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Cadierno, V. Synthetic Applications of the Parkins Nitrile Hydration Catalyst [PtH{(PMe2O)2H}(PMe2OH)]: A Review. Appl. Sci. 2015, 5, 380-401. https://doi.org/10.3390/app5030380

AMA Style

Cadierno V. Synthetic Applications of the Parkins Nitrile Hydration Catalyst [PtH{(PMe2O)2H}(PMe2OH)]: A Review. Applied Sciences. 2015; 5(3):380-401. https://doi.org/10.3390/app5030380

Chicago/Turabian Style

Cadierno, Victorio. 2015. "Synthetic Applications of the Parkins Nitrile Hydration Catalyst [PtH{(PMe2O)2H}(PMe2OH)]: A Review" Applied Sciences 5, no. 3: 380-401. https://doi.org/10.3390/app5030380

Article Metrics

Back to TopTop