Shelby V.
McCowen
,
Nicolle A.
Doering
and
Richmond
Sarpong
*
Department of Chemistry, University of California, Berkeley, California 94720, USA. E-mail: rsarpong@berkeley.edu
First published on 21st April 2020
Retrosynthetic analysis is a cornerstone of modern natural product synthesis, providing an array of tools for disconnecting structures. However, discussion of retrosynthesis is often limited to the reactions used to form selected bonds in the forward synthesis. This review details three strategies for retrosynthesis, focusing on how they can be combined to plan the synthesis of polycyclic natural products, such as atropurpuran and the related arcutane alkaloids. Recent syntheses of natural products containing the arcutane framework showcase how these strategies for retrosynthesis can be combined to plan the total synthesis of highly caged scaffolds. Comparison of multiple syntheses of the same target provides a unique opportunity for detailed analysis of the impact of retrosynthetic disconnections on synthesis outcomes.
The broader class of diterpenoid alkaloids are known to possess a range of biological activities, largely due to their interactions with ion channels.1,2,7 The biological activity of the arcutane alkaloids, however, has not been evaluated. This is also true of atropurpuran. However, there have been a number of patents filed regarding the use of atropurpuran derivatives to treat myriad conditions, including—but not limited to—pancreatic fibrosis,8 hypoxia,9 and inflammation.10 Synthetic access to a larger quantity of arcutane-type natural products could allow for detailed study into their biological activity as well as evaluate their potential as therapeutics.
The remainder of this review is divided into four sections. The first is focused on retrosynthetic strategies, including discussion of the use of biosynthetic pathways as inspiration, application of bond-network analysis, as well as of complementary functional group- and transform- oriented analyses. The second compares total syntheses of atropurpuran, while the third discusses total syntheses of the arcutane alkaloids. The final section includes other studies of bridged polycycles en route to these alkaloids.
Currently, there are two proposed biosynthetic pathways to the arcutane natural products. These pathways provide many opportunities from which to draw inspiration. In their report detailing the isolation of atropurpuran,6 Wang and coworkers proposed that an oxygenated hetidine-type scaffold, such as 6, could fragment along the C13–C14 bond to produce dialdehyde 7 (Scheme 1A).6 Further fragmentation would release ethylene to generate hemi-quinone 8, into which hydroxide could add in a conjugate fashion to give enolate 9. The dienyl moiety in 9 would then recombine with ethylene to forge the central BE bicycle (10). Dehydration to form 11, followed by an aldol cyclization, would then close the D ring, completing the arcutane framework. Dicarbonyl compound 12 would then undergo deoxygenation at C13 followed by allylic oxidation to afford atropurpuran (5). The Wang group did not discuss the biosynthesis of the arcutane alkaloids.
Scheme 1 Postulated biosyntheses of atropurpuran by Wang (A) and of arcutane natural products by Sarpong (B). |
An alternative biosynthesis was proposed by Sarpong and coworkers,14 who believed aspects of the Wang proposal were unlikely (see 7 → 8, 9 → 10). While the authors agreed that the arcutane framework likely arises from an oxygenated hetidine scaffold, they postulated that conversion from the hetidine to arcutane framework occurs through a deep-seated rearrangement (Scheme 1B). Beginning with alcohol 13, protonation and loss of water could produce non-classical carbocation 14. A [1,2]-acyl shift to forge the C20–C5 bond with concomitant cleavage of the C20–C10 bond followed by elimination, would give dioxygenated arcutane 15. Hydration of the double bond and introduction of a nitrogen atom “N” through a reductive process would then give the arcutane alkaloids, represented by 16. In a complementary fashion, nitrogen could be introduced directly to diketone 13 to produce 17. In an analogous manner as before, loss of water could produce cation 18 that could, in turn, undergo a related [1,2]-alkyl shift (Wagner–Meerwein rearrangement) followed by trapping with water to give 16. The Sarpong group supported their hypothesis with DFT calculations, which suggested that the transition state energies for the diterpenoid and diterpenoid alkaloid [1,2] shifts are +4.8 and +7.7 kcal mol−1 in the gas phase, starting from 14 and 18, respectively.14 This is consistent with previous studies, which showed greater migratory aptitude for acyl groups than for alkyl in Wagner–Meerwein rearrangements.15
In the case of other diterpenoid alkaloids, the nitrogen atom is believed to come from either serine or ethanolamine.1,16 As such, aconicarmicharcutinium (4) could be an intermediate in the biosynthetic pathway from terpenoid 15 to alkaloid 16, where ethanolamine has been incorporated but the ethanol moiety is yet to be cleaved. Feeding studies, similar to those used to determine the source of the nitrogen atom in other alkaloids,16 could be used to interrogate arcutane alkaloid biosynthesis. In addition, isotopic labeling of the C20 carbon16 (marked with a blue dot in Scheme 1) on either proposed precursor 6 or 13 could provide insight into the veracity of both proposed biosynthetic pathways.
Both proposed biosyntheses illustrated in Scheme 1 could provide inspiration to synthetic chemists targeting arcutane natural products to develop non-obvious and efficient routes for their syntheses. Given that both biosynthetic proposals for the arcutane natural products begin with an oxygenated hetidine scaffold, it may also be possible to leverage existing hetidine syntheses17–21 to examine the feasibility of the proposed biosynthetic pathways in a laboratory setting.
Such an approach to retrosynthesis focuses primarily on the topology of the target, rendering consideration of functional group manipulation secondary. In a recent perspective,24 Sarpong and Hoffmann analyzed the strategic disconnections utilized in other recent diterpenoid alkaloid syntheses and argued that the rules Corey laid out for network analysis, while well-suited to their original goal of automating retrosynthesis, should be applied more flexibly when analyzing such highly bridged natural products. For example, consideration should be given not to only single-bond disconnections, but also to two-bond disconnections as well (e.g., bicyclization transforms). The following discussion will include consideration of multi-bond disconnections and will refer to this modified version of Corey's protocol as ‘bond-network analysis’.
This bond-network analysis can aid significantly in the development of a retrosynthetic strategy for the caged, architecturally complex, arcutane scaffold. In addition to containing a caged, highly bridged framework, the functional groups in the arcutane natural products are mostly peripheral, allowing for an analysis that almost completely disregards functional groups, making it an ideal candidate for bond-network analysis. The maximally bridged ring of the arcutane framework (see emphasis in Fig. 2A) was identified as a central six-membered primary ring containing 4 bridging atoms in the diterpenoid scaffold and 6 bridging atoms in the arcutane alkaloids, which contain an additional pyrrolidine ring. Analysis of the strategic bonds reveals three potential options for the first retrosynthetic disconnection. First is a two-bond disconnection across either the C (purple bonds), D (red bonds), or the maximally bridged E (in the case of atropurpuran) rings, which would lead directly back to an all fused precursor containing no bridging elements (option 1, Fig. 2B). A second option is disconnection of a strategic single-bond (highlighted in green, option 2, Fig. 2C),† which would lead back to a bridged precursor. Alternatively, two-bond disconnection involving ring A (red bonds) and/or F (purple bonds, option 3, Fig. 2C) would also lead back to a bridged precursor. In either of the latter two cases, the remaining CD bicycle could then be subjected to bond-network analysis, which furnishes options for single-bond (highlighted in green, option 2.2/3.2, Fig. 2D) or two-bond (highlighted in blue and red, option 2.1/3.1) disconnection, either of which would produce a fused precursor. Considering the options outlined in this thought exercise, there are nearly 50 sets of possible bond-network analysis derived disconnections, however only 3 disconnections (those delineated in option 1) lead directly back to a fused precursor.
In the case of the arcutane scaffold, the heteroatom-based functional groups are mainly peripheral. However, the framework still provides ample possibilities for functional group- and transform-based disconnections (Fig. 3). For example, one could imagine that a [2.2.2]bicycle (highlighted in blue in Fig. 3) could arise from a Diels–Alder reaction between a dienophile and a cyclic diene; however, this transform requires a double bond within the bicyclic structure (highlighted in red), which would necessitate an extra step to adjust the oxidation state of the ring system.
If one were to map all the potential functional group-oriented transforms for a given molecule, a large array of retrosynthetic plans would emerge, many of which would be unproductive. However, when considered together with bond-network analysis, these strategies can provide the means for comprehensive retrosynthetic planning, as well as provide flexibility for the development of novel transforms.
The ideas of bond-network analysis, functional group-oriented retrosynthesis, and transform-oriented retrosynthesis will guide the discussion throughout this review as we examine how closely syntheses of arcutane natural products and related structures follow principles from each type of retrosynthesis and how the choice of strategy contributes to efficiency‡ in the forward synthesis. This review aims to not only familiarize the reader with the current literature on the arcutane framework, but to also provide insight from each synthesis, specifically through the lens of retrosynthetic analysis.
Overall, Qin's racemic synthesis of atropurpuran was completed in 27 steps from known materials (28 from commercial). While the oxidative dearomatization/Diels–Alder sequence is well-precedented for construction of the CD bicycle of other diterpenoid alkaloids, the decision to do so early in this synthesis is unique and proved effective.17,20,21 The Qin group also took advantage of the symmetry of functionality required in the A ring precursor to introduce a relatively simple A-ring structural moiety, exemplifying another principle of efficient synthesis: exploitation of hidden symmetry in synthetic intermediates.28,29 This strategic decision provided an advantage over their initial approach,30 which had involved the synthesis of a more elaborate A ring. As the first completed synthesis of the arcutane skeleton, the Qin atropurpuran synthesis serves as a benchmark against which subsequent syntheses may be measured.
Xu and coworkers' synthesis of atropurpuran benefited greatly from leveraging the most efficient disconnections, as identified by bond-network analysis, to access the tetracyclo[5.3.3.04,9.04,12]tridecane arcutane core in just four steps from commercial starting materials, showcasing the power of bond-network analysis as a tool for retrosynthesis of highly caged polycycles. Still, elements from the earlier synthesis of atropurpuran by Qin are also evident in Xu's strategy. For example, Xu's introduction of C18 and reduction of the penultimate intermediate (36), both proceeded analogously to transformations in the Qin synthesis. In addition, both groups utilized an oxidative dearomatization/intramolecular Diels–Alder sequence, a powerful, well-established way to generate [2.2.2]bicycles in diterpenoid alkaloid synthesis.17–21
Perhaps more intriguing than the similarities between the Qin and Xu syntheses are the differences—arising from small variations in the functional groups employed—that underpin the two routes to atropurpuran. The means by which the C19 aldehyde was installed, for example, differed significantly despite both routes utilizing C4 ketones as precursors with the same carbon backbone (see 37 and 39, Scheme 4A). In the case of the Xu synthesis, the β-ketone moiety at C20 in 39 (highlighted in purple) complicated the introduction of a carbon nucleophile into C4 since direct addition led to retro-aldol ring opening or non-specific decomposition. Xu therefore introduced the C19 aldehyde in a roundabout manner (see 31 → 35, Scheme 3). In contrast, these issues were circumvented in the Qin synthesis through the use of 37, which contained a silyl ether at C20. The other major difference is seen in the site of hydroxylation on the arene used in the shared oxidative dearomatization/Diels–Alder sequence. In the Qin synthesis, the hydroxy-bearing carbon in the arene maps onto the C16 carbonyl (see abbreviated structure 41, Scheme 4B). After olefination at this site, an allylic oxidation was required to introduce the C15 hydroxy group (41 → 42). Unfortunately, allylic oxidation favored the undesired diastereomer, necessitating two additional redox manipulation steps. Xu was able to avoid this issue by mapping the arene hydroxy group onto C15, facilitating a more direct conversion of the Diels–Alder adduct to the desired allylic alcohol (43 → 44), highlighting the dependence of synthetic efficiency not only on the sites of bond formation, but also on the functional handles chosen to build those bonds.
Scheme 5 Bond-network analysis (A) and key bond forming steps (B) in Qin's synthesis of (±)-arcutinine. (C) Synthesis of enantioenriched 50. |
Although the initial 21-step (25 from commercial materials) synthesis of 2 by the Qin group was racemic, they rendered it enantioselective through an alternative route to access enantioenriched 50 (Scheme 5B). The modified synthesis began with an enantioselective conjugate addition of a methyl group into cyclohexenone 55,36,37 mediated by a copper complex generated from phosphoramidite 56, CuTC, and AlMe3. The resultant enol ether was then treated with aldehyde 48 to give aldol product 57 in 92% ee. Alcohol 57 was then converted to enantioenriched 50 in 8 steps. The enantioselective synthesis of (−)-arcutinine was achieved in 26 steps from known starting materials (30 from commercial).
It is valuable to compare the two Qin syntheses (of atropurpuran and arcutinine) and examine how they differ in their strategies despite both deriving from the same bond-network analysis derived option set (2.2, Fig. 2) and using the same key ring forming steps (oxidative dearomatization/intramolecular Diels–Alder and SmI2-mediated ketyl–alkene cyclization). One of the larger differences is the point at which the CD bicycle is built. In the atropurpuran synthesis, the CD bicycle was constructed relatively early (step 7) whereas in the arcutinine synthesis, it was built toward the end of the synthesis (step 15). By constructing the [2.2.2]bicycle early in the atropurpuran synthesis, the synthetic intermediates benefited from the greater conformational rigidity, facilitating high levels of diastereoselectivity. In contrast, amine 52 had larger conformational freedom, requiring the addition of a bulky TMS group to achieve the desired diastereoselectivity of the aza-Wacker cyclization (2.5:1 d.r., see Scheme 5, 52 → 53). Once the cis-fused AF ring system was installed with the desired stereochemistry at C20, the remaining synthesis proceeded smoothly and with diastereocontrol. In addition, by having the pyrrolidine ring in place, the ketyl–alkene cyclization occurred without the necessity of introducing a bulky group to force C9 and C10 close together, unlike in their atropurpuran synthesis (see Scheme 2, 24 → 25). In addition, the phenolic isomer that Qin used to build the arcutinine CD ring system is different from that used to build atropurpuran. Pursued contemporaneously with Xu's atropurpuran synthesis, Qin's arcutinine synthesis again highlights the improved efficiency imparted by mapping the phenolic hydroxy group onto C15 instead of C16 (see Scheme 5, 46 → 45), as had been done in the original atropurpuran synthesis (see Scheme 2, 20 → 19).
Scheme 6 Bond-network analysis (A) and key bond forming steps (B) in Sarpong's synthesis of (±)-arcutinidine. (C) Unsuccessful transformations in Sarpong's synthesis. |
It is informative to compare the Qin and Sarpong syntheses as both take advantage of the bond-network derived disconnection option 2.2. Both leverage the same strategic disconnections: the C9–C10 bond of the maximally bridged ring, followed by two-bond disconnection of the CD ring system to give a fused precursor. Qin's two-bond disconnection cleaved the D ring, whereas Sarpong's removed the C15–C16 bridge common to the C and D rings. This unusual disconnection led to the use of a Wessely-type oxidative dearomatization to append a dienophile that facilitated the formation of the CD bicycle. However, Diels–Alder adduct 65 contained superfluous oxygenation at C9, which could not be converted into a suitable substrate for ketyl–olefin coupling (5) and required deoxygenation in two subsequent concession steps (see Scheme 5B, 45 → 54 and Scheme 6C, 65 → 69).
While the choice to disconnect the CD C15–C16 bridge in the retrosynthetic sense is what enabled the completion of arcutinidine, it was not the original plan. The Sarpong group originally sought to use option 1 from the bond-network analysis, as realized by Xu in the context of atropurpuran (Scheme 6C). There, the maximally bridged ring would have been forged through cycloaddition of a fused precursor, ideally allylic alcohol 68. However, all attempts to introduce a vinyl group at C10 from intermediate 67 were unsuccessful, likely due to steric encumbrance from the C5 quaternary center, as well as the concavity of the molecule created by the cis-fused AF ring system and the C20 stereocenter. The lack of success in this context showcases how bicyclization transforms, while rapidly increasing structural complexity, also rapidly increase steric congestion, which can introduce new challenges in the forward synthesis. Turning this impasse into an opportunity, the concavity of the molecule was leveraged to effect a diastereoselective Wessely oxidation. Changing the disconnection option ultimately enabled the completion of a synthesis of arcutinidine, however at the cost of a significantly longer synthesis.
There is a lot to be learned from Li and coworkers' bioinspired synthesis. First, it provides experimental evidence for the plausibility of Sarpong's proposed biosynthesis (Scheme 1B). In addition to proving that the rearrangement between the hetidine and arcutane skeletons is feasible, Li also showed that it is possible to introduce nitrogen and complete the pyrrolidine ring from the corresponding keto-aldehyde (Scheme 7, 80 → 3). The Qin group had been unable to effect the same transformation from atropurpuran, citing the C4 quaternary center and the C1–C10 alkene as likely impediments. On the contrary, Li showed that the C4 quaternary center does not prevent N incorporation (see 80 → 3); it is possible that the C1–C10 alkene impedes cyclization. That said, the Li group did not advance intermediate 80 to atropurpuran. This was not commented upon in their report.
Scheme 7 Bond-network analysis (A) and key bond forming steps (B) in Li's synthesis of the arcutane alkaloids. |
In addition to lending credence to the proposed arcutane biogenesis, the Li synthesis also exposes a potential limitation of bond-network analysis. In network analysis, emphasis is placed on disconnection at the maximally bridged ring, which is believed to represent structural complexity. As skeletal rearrangements typically involve both building and breaking sigma bonds, network analysis does not align itself well with identifying potential rearrangements. The value of such rearrangements is difficult to define; however, they can provide the ability to trace back a complex bridged framework to a more readily accessible one and have been used in other diterpenoid alkaloid syntheses.18,41–45 In the case of the Li synthesis of the arcutane-type alkaloids, the bioinspired rearrangement is central to what is the shortest reported syntheses to date. That said, bioinspired syntheses and retrosyntheses borne out of bond-network analysis are not necessarily mutually exclusive; Li's rearrangement forges the maximally bridged ring at a late stage and all other disconnections follow the tenets of bond-network analysis.
The Li synthesis is also noteworthy in the use of a convergent46 Diels–Alder reaction to bring together fragments 73 and 74 (see Scheme 7), allowing for a reduced linear step count. By using this atom economical28 Diels–Alder cycloaddition, the Li group avoided the introduction of additional heteroatom-containing functional groups. This strategy was also utilized in the construction of the CD bicycle (see 71 → 76), further contributing to the overall efficiency of the synthesis. Li's Prins/Wagner–Meerwein cascade also takes advantage of functionality and unsaturation inherent within the arcutanes. The Xu group similarly relied on C–C π bonds to effect the desired bond forming transforms (see Scheme 3, 30 → 28 → 31). While the Li and Xu syntheses took very different approaches to arrive at the arcutane framework, both groups were able to minimize the introduction of heteroatom-containing functional groups that would later need to be removed, which allowed them to achieve the shortest syntheses of arcutinidine and atropurpuran, respectively.
Scheme 8 Bond-network analysis (A) and key bond forming steps (B) in Kobayashi's synthesis of the arcutane framework. |
While Kobayashi and coworkers did not ultimately complete a total synthesis of atropurpuran, this work appears to have laid part of the groundwork for Xu and coworkers' synthesis of atropurpuran (see Section 3.2). In both syntheses, an oxidative dearomatization/intramolecular Diels–Alder cycloaddition is employed to simultaneously construct all the bridged ring systems in the arcutane scaffold (see option 1, Fig. 2). Comparison of the fused precursors also provides further insight into the similarities of the two plans. Xu's ketone 29 resembles Kobayashi's 84. Ketone 29 is, however, commercially available as opposed to ketone 84, which was synthesized in 7 steps from ortho-eugenol (83). Both groups then append unsaturated groups alpha to the carbonyl to be utilized in a Grubbs ring-closing metathesis to forge the A ring. For this purpose, Kobayashi uses two equivalent allyl groups whereas Xu employs more highly functionalized substituents. Kobayashi's inspiring report of the synthesis of the arcutane ABCDE ring system thus represented a major advancement toward the total synthesis of atropurpuran, the completion of which would not be reported for another five years by others.
The Hsung retrosynthesis of atropurpuran highlights the cyclization methodology developed: a unique use of a triene as a masked diene for an intramolecular Diels–Alder to produce the BCD ring system. While this method provides an expedient route to three of the five rings of atropurpuran, it did not prove fruitful for the complete synthesis of atropurpuran. Furthermore, the transformations reported by Hsung did not feature functional groups at the key bridging sites (C5, C10, and C14) of 95 nor did it provide ways to introduce functionality, which would be required for a successful total synthesis. However, the Hsung approach does highlight the potential of reactions that forge multiple C–C bonds in a single step, as well as the utility of pericyclic reactions in constructing heteroatom-poor natural products. In addition, the Hsung approach would minimize functional group manipulation, allowing production of the BCD carbon scaffold in only three steps from known materials (seven steps from commercial in the longest linear sequence).
Examination of the Singh synthesis through the lens of bond-network analysis is impeded by the fact that the A ring was never installed. However, by analogy to a full arcutane scaffold and the other reported syntheses, some insights can be gleaned. Similar to Qin's synthesis of atropurpuran, the Singh approach began with an oxidative dearomatization/Diels–Alder cycloaddition sequence to build the CD bicycle. From there, the Singh group made an interesting choice of excising the C17 carbon (see Scheme 10, 97 → 98) which would be required to complete a synthesis of atropurpuran. The maximally bridged ring was then forged before the last bridging element was secured, in stark contrast to the other syntheses where closure of the maximally bridged ring completed the installation of bridging elements.
There are several clues as to why this synthesis was not advanced to the full arcutane framework. First, dibarrelane-like 103 lacked the required functionality at C20, as well as at the C5 bridgehead, which would be challenging to install later in the synthesis. While such elements could be installed earlier in the synthesis when the intermediates were more highly functionalized, the Singh group instead removed functional handles that could have been leveraged for this purpose (see the loss of C17 in 97 → 98 and loss of unsaturation in 99 → 100). Additionally, the three functional groups resident in 103 are all ketone moieties that are not readily distinguishable. The synthesis also relied on different heteroatom-based functional groups for each stage of the synthesis, which would have led to a significantly longer synthesis than many of those discussed previously.
Scheme 11 (A) Retrosynthetic analysis of atropurpuran. (B) Key bond forming steps (B) in Qin's synthesis of the arcutane ABC ring system. |
The Qin retrosynthesis, which did not align with the bond-network analysis for this scaffold, was designed to probe the validity of the ethylene Diels–Alder reaction that Wang had proposed for the biosynthesis of atropurpuran.6 Unfortunately, given the undesired stereochemistry at C8 in tricycle 115, Wang's hypothesis remained untested. It is worth noting that this study from Qin is the only reported attempt to synthesize the arcutane scaffold in which the authors began with one of the central rings (E or B) and worked outward, annealing the other rings sequentially. This appears to have reduced overall synthetic efficiency; synthesis of tricycle 115 took 18 steps (21 from commercial), as opposed to the 27 total steps (28 from commercial) needed for their total synthesis of pentacyclic atropurpuran. That said, the disparity in step count was exacerbated by the reliance of the route outlined in Scheme 11 on carbonyl chemistry, which necessitated repeated manipulation of protecting groups. Still, the unviability of a synthesis of atropurpuran from 115, regardless of the number of steps involved, further validates the efficacy of bond network analysis as a tool in retrosynthesis.
Taken together, these syntheses show that bond-network analysis can provide an effective starting point in retrosynthesis, guiding the synthetic chemist in the development of an efficient synthesis. The tenets of bond-network analysis were by-and-large followed in all of the completed arcutane syntheses, whether intentionally or not. While both one- and two-bond disconnections can be considered strategic, two-bond disconnections proved particularly effective. All successful syntheses included at least one bicyclization step. This fact highlights not only the reliability and the ubiquity of the Diels–Alder transform, the most common method used herein for two-bond disconnection of [2.2.2]bicycles, but also the inherent preference for the application of bicyclization transforms to achieve a rapid decrease in complexity in the retrosynthetic direction. Not only was this transform utilized in the total synthesis of the arcutane-type natural products, but also in many of the efforts toward the arcutane scaffold. However, a remaining challenge in the synthesis of such highly bridged polycycles is to leverage these disconnections for convergent synthesis. Of the syntheses surveyed here, only Li's arcutane alkaloid synthesis involves coupling of two cyclic components.
Additionally, although two-bond disconnections tended to correlate positively to step efficiency, the functional handles utilized to forge skeletal bonds in the forward sense were also shown to have a large impact. Functional group-oriented analysis is thus an important secondary consideration in retrosynthesis. The challenge is to minimize use of functional handles that one would need to remove later, leveraging functionality present in the natural product wherever possible. In the context of arcutane natural products, which contain relatively few heteroatoms, this presents an additional challenge: how to forge skeletal C–C bonds with minimal reliance on heteroatom-based functional groups. While alkenes are often utilized throughout the syntheses discussed herein—enabling various Diels–Alder cycloadditions, Xu's enyne metathesis, and more—there remains significant room for advancement in this area as the majority of these transformations still involved at least one heteroatom that did not appear in the target structure. The exception to this was Hsung's series of C–C bond-forming pericyclic reactions en route to the arcutane scaffold.
As bond-network analysis is based on bond disconnection, it necessarily focuses on structural—rather than synthetic—complexity and, thus, emphasizes the closure of new rings over rearrangement processes. This is an area where transform-oriented and bioinspired retrosynthesis can augment the topology-oriented bond-network analysis. The potential benefits of such a hybrid approach become apparent when one considers the Li synthesis, which features a bioinspired Prins/Wagner–Meerwein cascade and is the shortest synthesis of the arcutane alkaloids to date.
Bond-network analysis, unbiased by known reactions, can identify strategic disconnections for which no precedent exists and can thus challenge chemists to find creative solutions and drive development of new reactions. This critical examination of the arcutane syntheses underscores this idea, as well as the inventiveness of the synthetic groups towards solving problems. While there were not many examples of new reactivity—such as Sarpong's use of the oxopyrollium Diels–Alder reaction or Hsung's pericyclic cyclization sequence—these syntheses contribute to the paradigm shift in the use of “protecting groups”. A number of the syntheses utilized traditional protecting groups to enforce the desired diastereoselectivity in reactions (e.g., the use of a TMS group in Qin's arcutinine syntheses) or as carbon sources (e.g., Li's use of a MOM group in their arcutinidine synthesis). The creativity displayed is also highlighted by the fact that multiple groups chose similar disconnection types but arrived at the natural products through very different means.
Footnotes |
† In the case of the arcutane alkaloids, a single bond disconnection across the C5–C20 bond would not be considered strategic due to the macrocyclic intermediate it would form. |
‡ The term efficiency, for the purpose of this review, will only refer to step count. A “one-pot” reaction will count as a single step only if none of the components of the reaction mixture are removed (including removal of solvent). The step counts for the following syntheses are adjusted from their original reports to follow this criterion. |
This journal is © The Royal Society of Chemistry 2020 |