Abstract
The lomaiviticins are dimeric genotoxic metabolites that contain unusual diazocyclopentadiene functional groups and 2–4 deoxyglycoside residues. Because only 6 of 19 carbon atoms in the monomeric aglycon unit are proton-attached, their structure determination by NMR spectroscopic analysis is difficult. Prior structure elucidation efforts established that the two halves of the lomaiviticins are joined by a single carbon–carbon bond appended to an oxidized cyclohexenone ring. This ring was believed to comprise a 4,5-dihydroxycyclohex-2-ene-1-one. The bridging bond was positioned at C6. This structure proposal has not been tested because no lomaiviticin has been prepared by total chemical synthesis or successfully analyzed by X-ray crystallography. Here, we disclose microED studies which establish that (−)-lomaiviticin C contains a 4,6-dihydroxycyclohex-2-ene-1-one residue, that the bridging carbon–carbon bond is located at C5, and that the orientation of the cyclohexenone ring and configuration of the secondary glycoside are reversed, relative to their original assignment. High-field (800 MHz) NMR analysis supports the revised assignment and suggests earlier efforts were misled by a combination of a near-zero 3JH4,H5 coupling constant and a 4JC,H coupling interpreted as a 3JC,H coupling. DFT calculations of the expected 13C chemical shifts and C–H coupling constants provide further robust support for the structure revision. Because the interconversion of lomaiviticins A, B, and C has been demonstrated, these findings apply to each isolate. These studies clarify the structures of this family of metabolites and underscore the power of microED analysis in natural product structure determination.
Graphical Abstract

INTRODUCTION
The lomaiviticins are cytotoxic bacterial metabolites characterized by the presence of the deoxyglycosides β-N,N-dimethyl-l-pyrrolosamine and α-l-oleandrose, one or two diazotetrahydro[b]benzofluorene (diazofluorene) residues, and a single carbon–carbon bond linking two highly oxygenated cyclohexene rings (Figure 1).1 The C2-symmetric derivatives known as (−)-lomaiviticins A and B were first reported in 2001. The structures 1a and 2a were advanced by He and co-workers based on HRMS and extensive NMR analysis (Figure 1a).2 The relative stereochemistry (including anomeric configuration) of the carbohydrates was determined by analysis of 3JH,H coupling constants and by comparison to other natural glycosides.3 The presence of the diazofluorene was inferred from the kinamycins (e.g, (−)-kinamycin F (4)), monomeric diazofluorenes whose structures are known with certainty.4 The structure elucidation of the 2,2′-bicyclohexyl core was more challenging. In aggregate, only 6 of 19 carbon atoms in the monomeric aglycon unit are proton-attached, and two of these carbon atoms (C8, C9) are separated from the core by the diazofluorene. Thus, C2, C4, C12, and C13 constitute the only proton-attached centers to guide NMR analysis.
Figure 1.

(a) Original structures advanced for (−)-lomaiviticins A, B, and C (1a, 2a, and 3a, respectively), and the structures of (−)-kinamycin F (4), the synthetic hydroxyfulvene 5, and the aglycon 6a, which was first prepared in 2011. (b) Revised structures of (−)-lomaiviticins A, B, and C (1b, 2b, and 3b, respectively) based on the data reported herein. The assignments that are exchanged in the revised structures are shown in orange in 1b. (c) The structural homology between (−)-lomaiviticins A–C has been established by semisynthesis.5,6
The structures 1a and 2a were based on NOESY, HMQC, and HMBC analysis. The data points from the initial report2 that are germane to the work below are as follows. H2 and H4 were observed as singlets in the 1H NMR spectra (500 MHz) of (−)-lomaiviticins A and B but were correlated in the COSY spectrum of (−)-lomaiviticin A. This correlation was attributed to W-plane coupling, thus forming the basis for assignment of a syn-facial arrangement of the C2 and C4 hydrogens, and the number of bonds separating them (four). A weak HMBC correlation between H4 and the diazo carbon (C5) in (−)-lomaiviticin A was reportedly observed; this correlation was attributed to 3JC,H coupling and used to anchor the location of the aminosugar with respect to the diazocyclopentadiene. Finally, a correlation between H2 and C2 was observed in both the HMQC and HMBC spectra of (−)-lomaiviticin A. These correlations supported assignment of the location of the bridging carbon–carbon bond as that shown. While the biosynthetic relationship between (−)-lomaiviticins A and B has not been rigorously established, the 2,6-dideoxy α-l-oleandrose residues of (−)-lomaiviticin A are acid-labile. Hydrolytic cleavage of these residues, followed by 2-fold hemiketal formation, converts natural (−)-lomaiviticin A to semisynthetic (−)-lomaiviticin B (Figure 1c).5
In 2012, the Herzon laboratory reported the isolation of (−)-lomaiviticin C, the product of formal hydrodediazotization of (−)-lomaiviticin A.6 The C1-symmetric structure 3a was assigned to (−)-lomaiviticin C based upon the doubling of many signals in the NMR spectra, a reduction in 26 mass units relative to (−)-lomaiviticin A, and the successful conversion of natural (−)-lomaiviticin C to semisynthetic (−)-lomaiviticin A (Figure 1c). New signals observed at 6.72 and 120.9 ppm in the 1H and 13C NMR spectra, respectively, of (−)-lomaiviticin C were attributed to the hydroxfulvene shown in red. These shifts are in agreement with fully synthetic hydroxyfulvenes, such as 5 (δ H = 6.74, δ C = 120.0). Acidic digestion of (−)-lomaiviticin C allowed for isolation of the carbohydrate residues. Their absolute stereochemistry was determined by comparison of optical rotation data to synthetic standards (both were found to be of the l-form). ROESY analysis was then used to relay this assignment to the cyclohexane core. However, to the best of our knowledge, single crystal X-ray crystallographic data has not been obtained for any lomaiviticin. Moreover, none of the structures 1a–3a have yet been prepared by chemical synthesis.5,7 The Herzon laboratory reported a synthesis of the aglycon structure 6a in 2011.5,7j Unfortunately, attempted hydrolytic cleavage of the carbohydrate residues in natural (−)-lomaiviticin A could not be effected without accompanying decomposition of the aglycon.5,8 Thus, a definitive link between fully synthetic and natural lomaiviticins has not been established in the 20 years since their initial disclosure.
Here, we report microED analysis of (−)-lomaiviticin C. This analysis indicates the structure of (−)-lomaiviticin C is that shown as 3b (Figure 1b). In this structure, the C1/C4 and C2/C3 positions are exchanged, and the configuration of C4 is inverted, relative to 3a. Since structural homology between (−)-lomaiviticins A, B, and C has been definitively established (vide supra, see also Figure 1c) these findings indicate the structures of (−)-lomaiviticins A and B are correctly depicted as 1b and 2b, respectively. High-field (800 MHz) NMR analysis of (−)-lomaiviticin C supports the structure 3b and indicates that earlier efforts may have been misled by a fortuitous combination of a near-zero 3JH4,H5 coupling constant and a 4JH4,C5 HMBC correlation that was interpreted as a 3JH4,C5 coupling.
RESULTS
We chose (−)-lomaiviticin C for study because the isolate is more abundant (liquid fermentation yields of (−)-lomaiviticins A and C are ~0.5–1 and 37–62 mg per liter, respectively),6 it is more stable than (−)-lomaiviticin A,9 and it is >2 orders of magnitude less cytotoxic.6 Additionally, the C1-symmetry of (−)-lomaiviticin C provides a nearly 2-fold increase in the number of NMR resonances and allows for detection of correlations between the two halves of the isolate, thereby greatly amplifying the utility of NMR analysis. We envisioned that microED, an emerging cryoEM method that allows for structural characterization of complex molecules from crystals that fall below the size required for common crystallographic techniques,10 might be uniquely effective in providing crystallographic data for (−)-lomaiviticin C.
In our initial studies, we found that deposition of 2.0 μL of a solution of (−)-lomaiviticin C (~200 ng) in tetrahydrofuran–acetonitrile (~0.1 mg/mL) onto a carbon TEM grid, followed by room temperature evaporation, yielded microcrystalline domains that diffracted to >2 Å. While these were insufficient for ab initio structural determination, we were encouraged by these initial findings. Ultimately, after extensive screening of sample preparation conditions, it was found that slow evaporation (30 d) of this solution at −20 °C led to microcrystals that diffracted to 1.05 Å (complete crystallographic parameters are presented in Table S1). Cryogenic data collection was required to mitigate radiation damage during collection of rotation data. Merging data from four crystals afforded an ab initio solution (R1 = 11.9%, refined anisotropically; Figure 2; see also Figure S1). From the preliminary structure, which was obtained without any human input other than the molecular formula, it was apparent that there was extensive structural variation from previous reports that relied exclusively on NMR analysis. Notably, the C1/C4 (C1′/C4′) and C2/C3 (C2′/C3′) positions were exchanged, and the configurations of C4 and C4′ were inverted. The connectivity of the atoms was determined by microED analysis. However, the positions of the heteroatoms in the aminosugar residues could not be assigned using the microED data alone. These heteroatoms were manually introduced to the locations established by earlier chemical degradation studies.6 The experimental and calculated powder diffraction data for structure 3b were in agreement (Figure S2).
Figure 2.

Preliminary (left) and fully refined (right) microED structures of (−)-lomaiviticin C (3b).
In light of these findings, we acquired high-field (800 MHz) NMR spectroscopic data for (−)-lomaiviticin C in acetone-d6, which was found to provide the optimal balance of solubility and consistency of shimming among a number of solvents examined. A Monte Carlo molecular dynamics simulation11 was carried out on structures 3a and 3b; the lowest energy conformers are shown in Figure 3. Notably, the calculated geometry of the lowest-energy conformer of 3b was in qualitative agreement with the geometry obtained by microED analysis (compare Figures 2 and 3). Additionally, the lowest-energy conformer of 3a contains an equatorial disposition of H2 and H4, which is consistent with the conformational analysis in the original structural studies.2
Figure 3.

Lowest-energy conformers (top) and line structures (bottom) of structures 3a and 3b. For clarity, the naphthoquinone residues are shaded in the lowest-energy conformers and omitted from the line structures.
NMR signals that were used to distinguish between the structures 3a and 3b are discussed below and presented in Figure 4 (for complete data, see Table S2). 1H NMR analysis revealed H2 and H2′ each as a doublet of doublets at δ 3.95 (J = 4.2, 1.5 Hz) and δ 3.85 (J = 4.1, 1.3 Hz), respectively. These were clearly coupled to each other and to H4 and H4′ in the COSY spectrum. H4 and H4′ were observed as narrow doublets at δ 5.51 (J = 1.1 Hz) and δ 5.33 (J = 1.3 Hz), respectively. A TOCSY experiment showed correlations between (H2, H4′) and (H2′, H4), thereby confirming these protons are part of the same spin system (see the Supporting Information).
Figure 4.

Diagnostic ROESY and HMBC correlations in the NMR spectra of (−)-lomaiviticin C supporting structure 3b. (a) ROESY correlations between (H4, H2), (H4′, H2′), (H4, H2′), and (H4′, H2). (b) ROESY correlation between (H4, H5′). (c) HMBC correlations between (H4, C2′), (H4, C2), (H4′, C2′), and (H4′, C2). (d) HMBC correlations between (H12, C1) and (H12′, C1′).
2D-ROESY analysis supported the assignment 3b but not 3a. We observed a clear correlation between (H2, H4) and (H2′, H4′) (Figure 4a). These correlations are consistent with structure 3b wherein the pairs of protons are vicinal to each other. The structure 3a necessitates that H2, H4, H2′, and H4′ are equatorially disposed to account for the putative W-plane coupling discussed above. However, this spatial arrangement would not be expected to give rise to the ROESY correlation shown in Figure 4a. Figure 4a also shows weaker but definitive correlations between (H2′, H4) and (H2, H4′). These correlations across the monomeric units are accommodated by the structure 3b, but would not be anticipated in structure 3a, wherein the pairs of protons are oriented on opposing sides of each cyclohexane ring. We also observed a ROESY correlation between (H5′, H4) (Figure 4b). This is readily accommodated by the position and configuration of H4 in 3b but not the alternate configuration found in structure 3a. Additional ROESY correlations support the revised structure assignment (see the Supporting Information).
HMBC analysis (using a nJC,H threshold of 8.0 Hz) provided further support for the structure 3b. We observed correlations between (H4, C2) and (H4′, C2′), which are consistent with either structure 3a or 3b (3JC,H coupling in 3a, 2JC,H coupling in 3b; Figure 4c). However, we detected additional correlations between (H4, C2′) and (H4′, C2). These latter correlations require 3JC,H coupling in structure 3b but would necessitate a 4JC,H coupling in structure 3a. However, a 4JC,H coupling is inconsistent with the ROESY data shown in Figure 4a. Finally, correlations between (H12, C1) and (H12′, C1′) were observed (Figure 4d). Interpretation of these correlations as 3JC,H couplings support the location of the carbonyl as that shown in 3b but would necessitate invoking a 4JC,H coupling through an aliphatic system to achieve consistency with structure 3a. On reducing the nJC,H threshold to 3.0 Hz, we detected a faint HMBC correlation between (H4, C5), as noted in the original characterization of (−)-lomaiviticin A,2 but not between (H4′, C5′) or (H5′, C4′) (see the Supporting Information).
To provide further support for the structure revision, we carried out computational modeling of 13C NMR chemical shifts according to the method of Hehre et al.12 We focused on lomaiviticin aglycon and (−)-lomaiviticin B because their caged structures significantly decrease the conformational flexibility of the molecules, thereby reducing the computational cost. To benchmark the computational approach, we first calculated the expected 13C NMR chemical shifts of the aglycon structure 6a (shown in Figure 1a), for which experimental 13C shifts have been recorded.5,7j Figure 5a depicts the absolute difference between the experimental and calculated chemical shifts as a function of position. We observed good agreement between the calculations and experiment, with all but three positions (C4, C11, and C11a) within 5 ppm of experimental values, and an overall root-mean-square (RMS) = 3.39. These results support this computational approach as a viable strategy to determine the expected 13C NMR spectra of the lomaiviticins.
Figure 5.

(a) Absolute difference between calculated and experimental 13C chemical shifts of the aglycon structure 6a. RMS = 3.39. (b) Absolute difference between experimental 13C chemical shifts of the aglycon structure 6a and natural (−)-lomaiviticin B. RMS = 5.82. (c) Absolute difference between calculated and experimental 13C chemical shifts of the aglycon structure 6b and (−)-lomaiviticin B, respectively. RMS = 1.99.
We then compared the experimental 13C chemical shifts of the aglycon structure 6a with those of natural (−)-lomaiviticin B (Figure 5b). The experimental shifts were in poor agreement (RMS = 5.82). A 17.9 ppm difference was observed between the bridging carbon atoms (C2), and large differences were also observed at the all-substituted C3 stereocenter (Δδ = 10.6 ppm) and the center bearing the aminosugar residue (C4; Δδ = 9.6 ppm). We then determined the expected 13C chemical shifts of the revised aglycon structure 6b and compared these to the experimental chemical shifts of natural (−)-lomaiviticin B (Figure 5c). Among the three data sets shown in Figure 5, these data were in closest agreement, with an overall RMS = 1.99. All experimental chemical shifts of (−)-lomaiviticin B were within 5 ppm of the values calculated for the structure 6b.
We also calculated the expected 13C NMR chemical shifts of the structures 2a and 2b and compared these to the experimental chemical shifts of natural (−)-lomaiviticin B (Figure 6). With the exception of C4, all chemical shifts calculated for 2b were within 5 ppm of the corresponding shifts of natural (−)-lomaiviticin B (RMS = 2.83). By comparison, the structure 2a showed large discrepancies in the critical C2, C3, and C4 positions of the cyclohexene ring, at two adjacent carbon atoms (C11b, C12) and at the hemiketal carbon (C1; RMS = 6.63).
Figure 6.

Absolute difference between calculated 13C chemical shifts of structures 2a and 2b and natural (−)-lomaiviticin B. Gray series: 2a (RMS = 6.63); blue series: 2b (RMS = 2.83).
Finally, we measured nJC,H coupling constants for H2, H2′, H4, and H4′ of natural (−)-lomaiviticin C using 2D-EXSIDE experiments. We determined the expected values of these couplings for structures 3a and 3b by DFT calculations. To render these calculations tractable, the single lowest energy conformers of 3a and 3b identified by molecular dynamics were used as input. These were each optimized by DFT (ωB97X-D/6–31G*; higher levels of theory do not lead to more accurate results, see ref 12) and the couplings were determined. Figure 7 depicts the absolute difference between the calculated and experimental coupling constants (because the 2D-EXSIDE experiment does not provide the sign of the coupling, the absolute values of the theoretical couplings were employed). As can be seen, the theoretical couplings derived from structure 3b were in better agreement (RMS = 1.80 vs RMS = 3.87 for 3a). The couplings between (H2, C4′) and (H2′, C4) are perhaps most diagnostic. The calculated 3JC,H couplings for structure 3b were 6.3 and 6.1 Hz for (H2, C4′) and (H2′, C4), respectively, in reasonable agreement with the experimental values of 10.2 and 10.8 Hz. By comparison, structure 3a requires these to be 4JC,H couplings, and we calculated 4JC,H = 0 for both (H2, C4′) and (H2′, C4) in structure 3a. Additionally, the calculated 3JH,H coupling constants for (H2, H4) and (H2′, H4′) in 3b were 1.6 and 1.1 Hz, respectively. These are in agreement with the experimental values of 1.1 and 1.3 Hz (vide supra). The calculated 4JH,H couplings for structure 3a were 0.5 and 0 Hz for (H2, H4) and (H2′, C2′), respectively.
Figure 7.

Absolute difference between experimental nJC,H coupling constants for H2, H2′, H4, and H4′ of natural (−)-lomaiviticin C and those expected for structures 3a and 3b based on DFT calculations. Experimental couplings were obtained by 2D-EXSIDE experiments, which do not provide the sign of the couplings. Therefore, the absolute values of the theoretical couplings were employed.
DISCUSSION
The lomaiviticins are complex cytotoxic bacterial metabolites that were first disclosed in 2001. They contain an unusual diazotetrahydrobenzo[b]fluorene functional group and the deoxyglycosides l-oleandrose and N,N-dimethyl-l-pyrrolosamine. The high degree of unsaturation of the lomaiviticins makes their de novo structure assignment exceedingly difficult using NMR analysis alone. In the 20 years since their initial disclosure, substantial efforts have been devoted toward their chemical synthesis. Yet, because none of the proposed structures 1a–3a have been synthesized, a comparison of spectroscopic data derived from fully synthetic material to that of the natural samples has not been possible. Moreover, single crystal X-ray crystallographic data of any lomaiviticin has not been disclosed.
Here, we used microED analysis of (−)-lomaiviticin C to revise the core structure of this family of metabolites. The results of the microED studies are supported by high-field NMR spectroscopy and computational modeling. The central issue at hand is the orientation of the cyclohexenone ring with respect to the diazofluorene residue. As best as we can discern from the original report, the connectivity shown in structures 1a and 2a rested on two observations. First, it was noted that a weak correlation was detected from H4 to C5 in the HMBC spectrum of (−)-lomaiviticin A (see Figure 8). At the time, this was quite reasonably interpreted as a 3JC,H coupling, and this served to anchor the position of the secondary ether (bearing the aminosugar) with respect to the diazofluorene. In addition, both H2 and H4 were observed as singlets in the 500 MHz 1H NMR spectrum of (−)-lomaiviticin A, but a COSY correlation was observed between these two protons. This COSY correlation was attributed to a syn diequatorial disposition of H2 and H4 (placing the aminosugar and bridging bonds in an axial orientation) giving rise to a 4-bond W-plane coupling. This coupling, in turn served to orient the bridging carbon atom (C2), both in position and stereochemistry, relative to C4, and the remaining substituents could be placed logically within the ring. It is worth noting that X-ray analysis of a 6,6′-bicyclohexenone structurally related to that found in (−)-lomaiviticin A revealed an axial disposition of the conjoining bond.13 In our analysis of natural (−)-lomaiviticin C, we did not observe an HMBC correlation from H4 to C5 using a nJC,H threshold of 8 Hz; a faint coupling was observed when the nJC,H threshold was reduced to 3 Hz. We attribute this correlation to a weak 4JC,H coupling in the structure 3b, rather than a 3JC,H coupling as originally proposed in structure 1a. Such four-bond couplings in unsaturated systems are well-precedented.14 Our molecular dynamics simulations indicate that the dihedral angles between (H2, H4) and (H2′, H4′) in (−)-lomaiviticin C are 78° and 86°, respectively. Consistent with this, the calculated 3JH,H coupling constants for (H2, H4) and (H2′, H4′) in 3b were 1.6 and 1.1 Hz, respectively, which is in agreement with experiment (1.1–1.3 Hz). Such small couplings could give rise to an apparent singlet at lower frequency, as originally observed. In sum, the original assignment appears to have gone awry due to a limited number of H–H and C–H couplings deriving from the high degree of unsaturation of the isolates, and the combination of two unusual spin-coupling effects.
Figure 8.

Original (left) and revised (right) core structures of the lomaiviticins.
Because the interconversion of lomaiviticins A–C has been demonstrated (Figure 1c), the findings here apply to each of these metabolites. The oxidation pattern found in the revised structures 1b, 2b, and 3b has been identified in other isolates, such as nenestatin A (Figure 9),15 and in intermediates in kinamycin biosynthesis.4c Interestingly, biosynthetic precursors to nenestatin A were found to undergo nucleophilic addition and dimerization,15b a pathway that may have implications for lomaiviticin biosynthesis. The lomaiviticin biosynthetic gene cluster has been identified,16 but the mechanism of dimerization remains poorly understood.
Figure 9.

Structure of nenestatin A.
Studies from the Herzon laboratory have established that the cytotoxic effects of (−)-lomaiviticin A derive from the induction of double-strand breaks (DSBs) in DNA.17 Because these studies were conducted using semisynthetic samples of (−)-lomaiviticin A, derived from diazotransfer to natural (−)-lomaiviticin C, the conclusions of these studies remain valid. However, the mode of interaction of (−)-lomaiviticin A with DNA18 will need to be reinterpreted.
CONCLUSION
In conclusion, we have used a combination of microED, NMR, and computational analysis to advance revised structures for the lomaiviticins. In the revised structures 1b, 2b, and 3b the positions of the ketone and aminosugar (C1/C4), and the bridging carbon–carbon bond and the center bearing the oleandrose residues (C2/C3) are transposed. The configuration of the secondary ether bearing the aminosugar residue is also inverted. These studies highlight the power of the microED method to elucidate natural product structure in instances where NMR spectroscopy alone may be insufficient.
Supplementary Material
ACKNOWLEDGMENTS
The authors thank Duilio Cascio (UCLA-DOE Institute) for assistance in data processing. Financial support from the National Institutes of Health (R35-GM131913 to S.B.H), the David and Lucile Packard Foundation (Fellowship to H.M.N.), Bristol Myers Squibb (Unrestricted Grant in Synthetic Organic Chemistry to H.M.N.), and Yale University is gratefully acknowledged.
Footnotes
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c01729
Procedures for microED analysis, crystallization methods, and graphical reproductions of NMR spectroscopic data (PDF)
Coordinates for structure 6a (PDB)
Coordinates for structure 6b (PDB)
Coordinates for structure 2a (PDB)
Coordinates for structure 2b (PDB)
Coordinates for structures 3a and 3b (PDB)
Accession Codes
CCDC 2062671 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
Complete contact information is available at: https://pubs.acs.org/10.1021/jacs.1c01729
The authors declare no competing financial interest.
Contributor Information
Lee Joon Kim, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States.
Mengzhao Xue, Department of Chemistry, Yale University, New Haven, Connecticut 06511, United States.
Xin Li, Department of Chemistry, Yale University, New Haven, Connecticut 06511, United States.
Zhi Xu, Department of Chemistry, Yale University, New Haven, Connecticut 06511, United States.
Eric Paulson, Department of Chemistry, Yale University, New Haven, Connecticut 06511, United States; Chemical and Biological Instrumentation Center, Yale University, New Haven, Connecticut 06511, United States.
Brandon Mercado, Department of Chemistry, Yale University, New Haven, Connecticut 06511, United States; Chemical and Biological Instrumentation Center, Yale University, New Haven, Connecticut 06511, United States.
Hosea M. Nelson, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States.
Seth B. Herzon, Department of Chemistry, Yale University, New Haven, Connecticut 06511, United States; Department of Pharmacology, Yale School of Medicine, New Haven, Connecticut 06510, United States.
REFERENCES
- (1).For a recent review, see:; Herzon SB; Woo CM The diazofluorene antitumor antibiotics: Structural elucidation, biosynthetic, synthetic, and chemical biological studies. Nat. Prod. Rep 2012, 29, 87. [DOI] [PubMed] [Google Scholar]
- (2).He H; Ding WD; Bernan VS; Richardson AD; Ireland CM; Greenstein M; Ellestad GA; Carter GT Lomaiviticins a and b, potent antitumor antibiotics from micromonospora lomaivitiensis. J. Am. Chem. Soc 2001, 123, 5362. [DOI] [PubMed] [Google Scholar]
- (3).For example, oleandrose is well-known and is found in the avermectins, see:; (a) Burg RW; Miller BM; Baker EE; Birnbaum J; Currie SA; Hartman R; Kong YL; Monaghan RL; Olson G; Putter I; Tunac JB; Wallick H; Stapley EO; Oiwa R; Omura S Avermectins, new family of potent anthelmintic agents: Producing organism and fermentation. Antimicrob. Agents Chemother 1979, 15, 361. [DOI] [PMC free article] [PubMed] [Google Scholar]; Pyrrolosamine is found in pyrrolosporin, see:; (b) Lam KS; Hesler GA; Gustavson DR; Berry RL; Tomita K; MacBeth JL; Ross J; Miller D; Forenza S Pyrrolosporin a, a new antitumor antibiotic from micromonospora sp. C39217-r109–7. I. Taxonomy of producing organism, fermentation and biological activity. J. Antibiot 1996, 49, 860. [DOI] [PubMed] [Google Scholar]; (c) Schroeder DR; Colson KL; Klohr SE; Lee MS; Matson JA; Brinen LS; Clardy J Pyrrolosporin a, a new antitumor antibiotic from micromonospora sp. C39217-r109–7. Ii. Isolation, physico-chemical properties, spectroscopic study and x-ray analysis. J. Antibiot 1996, 49, 865. [DOI] [PubMed] [Google Scholar]
- (4).Isolation of (−)-kinamcyin F (4):; (a) Cone MC; Seaton PJ; Halley KA; Gould SJ New products related to kinamycin from streptomyces murayamaensis. I. Taxonomy, production, isolation and biological properties. J. Antibiot 1989, 42, 179. [DOI] [PubMed] [Google Scholar]; (b) Seaton PJ; Gould SJ New products related to kinamycin from streptomyces murayamaensis. Ii. Structures of pre-kinamycin, keto-anhydrokinamycin, and kinamycins e and f. J. Antibiot 1989, 42, 189. [DOI] [PubMed] [Google Scholar]; For reviews of the kinamycins, see:; (c) Gould SJ Biosynthesis of the kinamycins. Chem. Rev 1997, 97, 2499. [DOI] [PubMed] [Google Scholar]; (d) Marco-Contelles J; Molina MT Naturally occurring diazo compounds: The kinamycins. Curr. Org. Chem 2003, 7, 1433. [Google Scholar]; (e) Herzon SB The kinamycins. In Total synthesis of natural products. At the frontiers of organic chemistry;, Li JJ, Corey EJ, Eds.; Springer-Verlag: Berlin Heidelberg, 2012; pp 39. [Google Scholar]
- (5).Woo CM; Gholap SL; Lu L; Kaneko M; Li Z; Ravikumar PC; Herzon SB Development of enantioselective synthetic routes to (−)-kinamycin f and (−)-lomaiviticin aglycon. J. Am. Chem. Soc 2012, 134, 17262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (6).Woo CM; Beizer NE; Janso JE; Herzon SB Isolation of lomaiviticins c-e. Transformation of lomaiviticin c to lomaiviticin a, complete structure elucidation of lomaiviticin a, and structure-activity analyses. J. Am. Chem. Soc 2012, 134, 15285. [DOI] [PubMed] [Google Scholar]
- (7).For synthetic studies toward the lomaiviticins, see:; (a) Freed JD Toward a synthesis of the lomaiviticins. Ph.D. Thesis, Harvard University, 2005. [Google Scholar]; (b) Nicolaou KC; Denton RM; Lenzen A; Edmonds DJ; Li A; Milburn RR; Harrison ST Stereocontrolled synthesis of model core systems of lomaiviticins a and b. Angew. Chem., Int. Ed 2006, 45, 2076. [DOI] [PubMed] [Google Scholar]; (c) Morris WJ; Shair MD Stereoselective synthesis of 2-deoxy-β-glycosides using anomeric o-alkylation/arylation. Org. Lett 2009, 11, 9. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Krygowski ES; Murphy-Benenato K; Shair MD Enantioselective synthesis of the central ring system of lomaiviticin a in the form of an unusually stable cyclic hydrate. Angew. Chem., Int. Ed 2008, 47, 1680. [DOI] [PubMed] [Google Scholar]; (e) Zhang W; Baranczak A; Sulikowski GA Stereocontrolled assembly of the c3/c3′ dideoxy core of lomaiviticin a/b and congeners. Org. Lett 2008, 10, 1939. [DOI] [PubMed] [Google Scholar]; (f) Gholap SL; Woo CM; Ravikumar PC; Herzon SB Synthesis of the fully glycosylated cyclohexenone core of lomaiviticin a. Org. Lett 2009, 11, 4322. [DOI] [PubMed] [Google Scholar]; (g) Nicolaou KC; Nold AL; Li H Synthesis of the monomeric unit of the lomaiviticin aglycon. Angew. Chem., Int. Ed 2009, 48, 5860. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Morris WJ; Shair MD Synthesis of the n-(tert-butyloxycarbonyl)-o-triisopropylsilyl-d-pyrrolosamine glycal of lomaiviticins a and b via epimerization of l-threonine. Tetrahedron Lett. 2010, 51, 4310. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Lee HG; Ahn JY; Lee AS; Shair MD Enantioselective synthesis of the lomaiviticin aglycon full carbon skeleton reveals remarkable remote substituent effects during the dimerization event. Chem. - Eur. J 2010, 16, 13058. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Herzon SB; Lu L; Woo CM; Gholap SL 11-step enantioselective synthesis of (−)-lomaiviticin aglycon. J. Am. Chem. Soc 2011, 133, 7260. [DOI] [PMC free article] [PubMed] [Google Scholar]; (k) Scully SS; Porco JA Asymmetric total synthesis of the epoxykinamycin fl-120 b′. Angew. Chem., Int. Ed 2011, 50, 9722. [DOI] [PMC free article] [PubMed] [Google Scholar]; (l) Baranczak A; Sulikowski GA Synthetic studies directed toward dideoxy lomaiviticinone lead to unexpected 1,2-oxazepine and isoxazole formation. Org. Lett 2012, 14, 1027. [DOI] [PMC free article] [PubMed] [Google Scholar]; (m) Scully SS; Porco JA Studies toward the synthesis of the epoxykinamycin fl-120b′: Discovery of a decarbonylative photocyclization. Org. Lett 2012, 14, 2646. [DOI] [PMC free article] [PubMed] [Google Scholar]; (n) Feldman KS; Selfridge BR Enantioselective synthesis of the ent-lomaiviticin a bicyclic core. Org. Lett 2012, 14, 5484. [DOI] [PubMed] [Google Scholar]; (o) Lee AS; Shair MD Synthesis of the c4-epi-lomaiviticin b core reveals subtle stereoelectronic effects. Org. Lett 2013, 15, 2390. [DOI] [PubMed] [Google Scholar]; (p) Feldman KS; Selfridge BR Synthesis studies on the lomaiviticin a aglycone core: Development of a divergent, two-directional strategy. J. Org. Chem 2013, 78, 4499. [DOI] [PubMed] [Google Scholar]; (q) Rose JA; Mahapatra S; Li X; Wang C; Chen L; Swick SM; Herzon SB Synthesis of the bis(cyclohexenone) core of (−)-lomaiviticin a. Chem. Sci 2020, 11, 7462. [DOI] [PMC free article] [PubMed] [Google Scholar]; (r) Nicolaou KC; Chen Q; Li R; Anami Y; Tsuchikama K Total synthesis of the monomeric unit of lomaiviticin a. J. Am. Chem. Soc 2020, 142, 20201. [DOI] [PubMed] [Google Scholar]; (s) Kaneko M; Li Z; Burk M; Colis L; Herzon SB Synthesis and biological evaluation of (2s,2′s)-lomaiviticin a. J. Am. Chem. Soc 2021, 143, 1126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).As noted in the text, the oleandrose residues of (−)-lomaiviticin A (1) are readily-removed. However, forcing conditions are required to remove the aminosugar residues, presumably due to preferential N-protonation, which impedes ionization at the anomeric position. See ref 6.
- (9).Xue M; Herzon SB Mechanism of nucleophilic activation of (−)-lomaiviticin a. J. Am. Chem. Soc 2016, 138, 15559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Jones CG; Martynowycz MW; Hattne J; Fulton TJ; Stoltz BM; Rodriguez JA; Nelson HM; Gonen T The cryoem method microed as a powerful tool for small molecule structure determination. ACS Cent. Sci 2018, 4, 1587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Spartan 18; Wavefunction, Inc., Irvine, CA. [Google Scholar]
- (12).Hehre W; Klunzinger P; Deppmeier B; Driessen A; Uchida N; Hashimoto M; Fukushi E; Takata Y Efficient protocol for accurately calculating 13c chemical shifts of conformationally flexible natural products: Scope, assessment, and limitations. J. Nat. Prod 2019, 82, 2299. [DOI] [PubMed] [Google Scholar]
- (13).Konkol LC; Guo F; Sarjeant AA; Thomson RJ Enantioselective total synthesis and studies into the configurational stability of bismurrayaquinone-a. Angew. Chem., Int. Ed 2011, 50, 9931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Parella T; Espinosa JF Long-range proton-carbon coupling constants: Nmr methods and applications. Prog. Nucl. Magn. Reson. Spectrosc 2013, 73, 17. [DOI] [PubMed] [Google Scholar]
- (15).(a) Jiang X; Zhang Q; Zhu Y; Nie F; Wu Z; Yang C; Zhang L; Tian X; Zhang C Isolation, structure elucidation and biosynthesis of benzo[b]fluorene nenestatin a from deep-sea derived micromonospora echinospora scsio 04089. Tetrahedron 2017, 73, 3585. [Google Scholar]; (b) Huang C; Yang C; Zhang W; Zhang L; De BC; Zhu Y; Jiang X; Fang C; Zhang Q; Yuan C-S; Liu H.-w.; Zhang C Molecular basis of dimer formation during the biosynthesis of benzofluorene-containing atypical angucyclines. Nat. Commun 2018, 9, 2088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).(a) Kersten RD; Lane AL; Nett M; Richter TKS; Duggan BM; Dorrestein PC; Moore BS Bioactivity-guided genome mining reveals the lomaiviticin biosynthetic gene cluster in salinispora tropica. ChemBioChem 2013, 14, 955. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Janso JE; Haltli BA; Eustaquio AS; Kulowski K; Waldman AJ; Zha L; Nakamura H; Bernan VS; He H; Carter GT; Koehn FE; Balskus EP Discovery of the lomaiviticin biosynthetic gene cluster in salinispora pacifica. Tetrahedron 2014, 70, 4156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).For a review, see: Herzon SB The mechanism of action of (−)-lomaiviticin a. Acc. Chem. Res 2017, 50, 2577.
- (18).Woo CM; Li Z; Paulson EK; Herzon SB Structural basis for DNA cleavage by the potent antiproliferative agent (−)-lomaiviticin a. Proc. Natl. Acad. Sci. U. S. A 2016, 113, 2851. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
