Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Dec 18.
Published in final edited form as: J Org Chem. 2021 Jun 7;86(12):7904–7919. doi: 10.1021/acs.joc.0c03069

The Formation of Impossible Rings in Macrocyclooligomerizations for Cyclodepsipeptide Synthesis: The 18-from-12 Paradox

Abigail N Smith 1, Jeffrey N Johnston 2
PMCID: PMC8611659  NIHMSID: NIHMS1757375  PMID: 34097410

Abstract

A reinvestigation into the macrocyclooligomerization (MCO) of a tetradepsipeptide is reported, uncovering a paradox in which the MCO of depsipeptide monomers can produce “impossible” ring sizes: a 12-atom chain produced the expected 24-membered ring, alongside unexpected 18- and 30-membered cyclic oligomeric depsipeptides (CODs). We report an alternative preparation of authentic 18- and 36-membered macrocycles for this case using a stepwise synthesis that provides definitive analytical characterization for each ring size. Our investigation yields a recharacterization and reassignment of two macrocycles originally reported in this MCO series, along with updated yields and isothermal titration calorimetry data after implementation of new critical protocols for purification and subsequent analysis. Initial studies to probe this mechanistic conundrum are described.

Graphical Abstract

graphic file with name nihms-1757375-f0001.jpg

INTRODUCTION

We recently reported an approach to small collections of cyclic depsipeptides in a single reaction from simple monomers using a controlled oligomerization/macrocyclization process termed macrocyclooligomerization (MCO).1,2 The salient features of this process were the use of didepsipeptides and tetradep-sipeptides to prepare (macro)cyclic oligomeric depsipeptides (CODs) with 18, 24, 36, and higher rings sizes. Two distinct advantages were highlighted: the formation of (typically) three CODs in significant yield with a remarkably small unused mass balance and rapid access to large ring sizes in as few as six steps from readily available starting materials. A Mitsunobu reaction was leveraged for the chain elongation and macrocyclization steps, and evidence for ion templation (sodium, potassium, and cesium) was documented for several cases.

These studies were originally driven by our interest in (−)-verticilide,3 an insect ryanodine receptor (RyR) antagonist (Figure 1A, 24-membered COD from ent-1).4 Our discovery that ent-verticilide, but not nat-verticilide, is a potent and selective inhibitor of mammalian RyR25 led us to wonder whether the lower and higher oligomers of this structure might also exhibit activity. To accomplish this while developing a size-selective synthesis of the (unnatural) 36-membered oligomer (Figure 1, 36-membered COD), we explored a traditional stepwise synthesis strategy. We report here the successful realization of this initial goal and the finding that the data originally assigned to this oligomer is consistent instead with the 18-membered oligomer. Importantly, such a product is ostensibly precluded by the MCO of a tetradepsipeptide that should produce only ring sizes that are multiples of 12 (i.e., 12-, 24-, 36-, and 48-membered rings). Our in-depth study of this paradox has led to the conclusion that, in this case,6 for reasons not entirely clear, a competitive chain-cleaving step occurs to produce the unexpected 18- and 30-membered rings (Figure 1B). This study provides incontrovertible data for each COD ring in this tetradepsipeptide MCO case and tests several mechanistic hypotheses for its formation. Evidence suggests that a depsipeptide fragmentation contributes the needed monomer to access the paradoxical ring sizes.

Figure 1.

Figure 1.

Macrocyclooligomerization (MCO) approach to cyclic oligomeric depsipeptides (CODs): (A) partial revision of original findings and (B) documenting the formation of paradoxical ring-size oligomers.

RESULTS AND DISCUSSION

Preparation and investigation of molecules with sizes larger than most “small molecule” drugs but smaller than typical “small peptide” drugs have drawn increasing attention. This size regime has adopted the moniker “beyond the rule of five”, alluding to Lipinski’s rules7 but highlighting the potential for discovery when considering this largely unexplored chemical space. This endeavor is not without limits, including the need for concise access to conformationally well-defined molecules that might also exhibit favorable drug-like properties. Our effort has targeted cyclic depsipeptides for these reasons, culminating in the finding that small depsipeptide monomers could be converted to cyclic oligomeric depsipeptides by a straightforward oligomerization/cyclization process using the Mitsunobu reaction. This approach was further refined by the addition of Lewis acids with the anticipation that the cyclization step might be templated and therefore accelerated relative to further chain elongation.

Our use of this approach complements substantial prior work by others who used an MCO strategy to prepare oligomeric esters and amides. Most notable is the use of activated seco-acids to form macrodiolides (and triolides)814 and amidations of activated peptidic amino acids to construct the macrocyclic peptides valinomycin and westiellamide.1517 A common feature to these amidation reactions is a five-membered heterocycle, such as a thiazole or oxazole, installed in the peptide backbone.1827 This presumably enhances a conformation more favorable for cyclization.

A specific tetradepsipeptide (1, Figure 1) was noted to form unusually large cyclic depsipeptides that were 36- and 60-membered rings in size, alongside the 24-membered ring. Assignments to these ring sizes were made on the basis of fractionation in the workflow of purification, and the observation of the corresponding mass ions, albeit with relatively weak intensity. This case was reinvestigated to better understand the factors leading to larger ring-size production. A key factor was assumed to be the length of the depsipeptide monomer, with the tetradepsipeptide (1, contributing 12 backbone atoms) leading to allowed ring sizes that included 24-, 36-, 48-, and 60-membered rings. A route to the 36-membered ring using a stepwise chain-extension sequence and final macrocyclization was planned and executed to acquire larger amounts of this target. The first chemical synthesis of nat-verticilide employed a linear route in which the stereochemistry of the α-hydroxyheptanoic acid moiety was installed using a chiral auxiliary, requiring six steps overall to form the didepsipeptide.4 They recently developed a second-generation synthesis, which uses serine as a starting material to shorten the step-count to five steps.28 Our synthesis of the hydroxyheptanoic acid leveraged enantioselective catalysis and Umpolung Amide Synthesis to arrive at the didepsipeptide in three steps.

Our approach relies on O-deprotected benzyl ester 4 and carboxylic acid 6 prepared using optimized procedures of those previously reported (Scheme 1).1 These didepsipeptide units were coupled using a Mitsunobu protocol to form the tetradepsipeptide 7 (Scheme 2). This material was divided, with half that was O-deprotected to 9, and half that was hydrogenolyzed to 8 prior to segment coupling [4+4] using another Mitsunobu reaction. The MOM-protecting group of the resulting octadepsipeptide 10 was removed, and the resulting alcohol was reacted with tetradepsipeptide carboxylic acid 8 using Mitsunobu conditions. This furnished the dodecadepsipeptide 11 in 60% yield. This material was deprotected at both termini, subjected to a Mitsunobu macrocyclization protocol (1 mM concentration), and then isolated as a 36-membered COD (12) (70% for three steps).

Scheme 1.

Scheme 1.

Enantioselective Synthesis of Differentially Protected and Unprotected Didepsipeptide Monomers

Scheme 2.

Scheme 2.

Linear, Stepwise Synthesis of the 36-Membered Cyclic Dodecadepsipeptide.

Analysis of 12 revealed a 1H NMR spectrum with the characteristic number and type of peaks in this COD series, but it was distinct from the 1H NMR spectrum of the material originally assigned to the 36-membered ring (Figure 1A). More telling was the observation of an intense parent ion peak ([M + Na]+ 1217.7146) by HRMS consistent with a 36-membered COD. This parent ion was far more intense than that observed in the original material (Figure 2). A second distinguishing feature in the HRMS spectrum for the authentic 36-membered ring is the absence of a singly charged 620 Da ion [M18 + Na]+. Instead, the ion peak at 620.3518 corresponds to [M36 + 2Na]2+.

Figure 2.

Figure 2.

Comparison of HRMS data from originally assigned “‘36” COD (A) formed by MCO and authentic 36-membered COD (B) formed through stepwise linear synthesis. [exact mass calculated for [M36 + Na]+ = 1217.7148; [M18 + Na]+ = 620.3523; [M36 + 2Na]2+ = 620.3523].29

The successful synthesis of 12 using a stepwise route provided compelling data to fingerprint the 36-membered COD 12. Given that all other aspects of the spectroscopic data were consistent with a highly symmetrical oligomeric depsipeptide, the assignment of this material was considered sound, and the identity of the material originally assigned as the 36-membered COD was reevaluated. We ultimately postulated that the original material could be an 18-membered COD, an “impossible” product of the 12-mer oligomerization (Figure 1B). Data consistent with this assignment included (a) a high degree of symmetry by 1H NMR, indicating a product of the Mitsunobu reaction that produces homochiral α-oxy amide subunits, and (b) intense adducts of the parent ions including [M + H]+ at 598 and [M + Na]+ at 620 consistent with an 18-membered COD (Figure 2). Verification of this structure by independent synthesis was straightforward, leveraging a stepwise route analogous to the synthesis of the cyclo-dodecadepsipeptide 12. Hence, treatment of an O-deprotected tetradepsipeptide 9 with a benzyl deprotected didepsipeptide 6 under Mitsunobu conditions yielded the linear hexapeptide in 93% yield (Scheme 3). Final MOM deprotection, hydrogenolysis, and cyclization under dilute conditions gave the authentic 18-membered COD in a 75% yield over three steps. The 1H NMR (A, Figure 3) and HRMS (Figure 2) data mirrored the data associated with the compound originally assigned as the 36-membered ring (D, Figure 3). It should be noted that broadening and slight shifting of α-oxy amide methine peaks in the proton NMR spectra are evident when any amount of trifluoroacetic acid (TFA) from HPLC purification is present (see discussion on residual TFA below and in the Supporting Information). Even small amounts of TFA appear to affect the chemical shift and coupling, as illustrated in Figure 3, where the originally characterized 18-membered ring (spectrum C) appears reasonably distinct from the other samples (spectra B and D). This difference contributed to the initial belief that two different compounds were forming in the MCO reactions. We have since confirmed that upon complete removal of TFA, by exhaustive base washes, the spectra overlap (see the Supporting Information). From this data, we are able to conclude that the 18-membered COD is the major product when using didepsipeptide monomer.

Scheme 3.

Scheme 3.

Linear, Stepwise Synthesis of the 18-Membered Cyclic Hexadepsipeptide.

Figure 3.

Figure 3.

Full-scale 1H NMR spectrum of (A) the authentic 36-membered COD (12, Scheme 2), (B) an authentic sample of the 18-membered COD (16, Scheme 3), (C) the originally assigned 18-membered ring formed in the MCO reaction,30 and (D) the originally assigned 36-membered ring formed in the MCO reaction (Figure 1A). Inset: expansion from 5.2 to 4.4 ppm.31

Although it is a minor COD formed from the tetradepsipeptide MCO, the 18-membered COD is clearly formed. The 36-membered COD is not formed in either MCO reaction, and attempts to change the reaction conditions to favor its formation have been unsuccessful to date.

A stepwise synthesis of the 60-membered ring was considered but not pursued. However, a parallel investigation of the data led to the conclusion that this material is most likely another paradoxical product: the 30-membered COD. This reassignment is made based on an analysis of HRMS data exhibiting a more intense singly charged mass ion for the 30-membered ring relative to the mass ion corresponding to a 60-membered COD, as well as MS/MS data (see the Supporting Information). Figure 4 illustrates our final assignments as well as the general similarities and differences of each ring-size congener. The use of 1H NMR analysis can clearly distinguish each COD, but the irregular movement of specific peak types undermines its utility as a predictor of ring size.

Figure 4.

Figure 4.

Full-scale 1H NMR spectra of authentic samples of all COD rings sizes.

High-Resolution Mass Spectrometry Investigation.

At the time of the original investigation, the HRMS data supported the conclusions but required the perceived mechanistic limitations of a 12-atom monomer leading to lengths of only 24/36/48/60 to be decisive. In every spectra of the 18-membered ring and “36”-membered ring formed in the MCO reactions, only singly charged ion peaks at both 620 and 1217 Da were observed. For an 18-membered ring, these ion peaks would correspond to [M18 + Na]+ and [2M18 + Na]+. For a 36-membered ring, these would correspond to [1/2M36 + Na]+ and [M36 + Na]+.32 However, in all cases, the ion at 620 Da was more intense. Only after the synthesis of the authentic samples of each ring size through a stepwise route was a clear difference apparent in the MS spectra (Figure 5). To further probe behaviors for these oligomers by HRMS, MS/MS experiments were applied to authentic samples of the 18- and 36-membered rings.

Figure 5.

Figure 5.

Comparison of the HRMS spectra and corresponding MS/MS spectrum of 620 and 1217 ion peaks in each sample of authentic 18-membered and 36-membered rings (z = charge).

When the 620 Da ion is fragmented in each experiment, the results are remarkably different. From the 18-membered ring, the singly charged 620 Da ions gave no fragments above m/z of 620. However, when the doubly charged 620 Da ion for 12 [M36+Na]2+ was fragmented, an intense 1217 ion was observed, corresponding to [M36 − Na]+. Additionally, there were a handful of fragments from the ring itself. When fragmenting the 1217 Da ion in each sample, again the MS2 patterns were distinct. The MS-MS of the 1217 ion from 16 showed the largest ion at 620 Da. This would correspond to the dissociation of the noncovalent dimer [2M18 + Na]+ with loss of one macrocycle (M18). Perhaps most intriguing is the presence of fragments heavier than 620 Da. This would imply that the non-covalent interaction that forms the singly charged dimer ion is rather strong as some of these dimers are experiencing fragmentation of the covalent bonds in the backbone of the macrocycle first.

Rationalizing the 18-from-12 Paradox.

The formation of an 18-membered COD from a 12-atom chain could result from several general pathways: (1) the presence of shorter monomers as impurities that co-oligomerize (e.g., 12+6) during MCO, (2) tetradepsipeptide monomer (12-atom chain) decomposition during reaction but pre-MCO, generating 6-atom chains, and (3) decomposition of oligomers to 6-atom monomers that then react with tetradepsipeptide monomer (e.g., 6+12). In the synthesis of these macrocyclic depsipeptides, the key oligomerization step occurs through a Mitsunobu reaction in which the stereochemistry of the α-hydroxy amide is inverted. When probing the different mechanistic possibilities for simple tetradepsipeptide cleavage, almost all of the options result in two linear peptides, heterochiral at the terminal α-hydroxy amide carbon. If this occurred, the isolated macrocycles would be diastereomers heterochiral at one α-oxy amide, and the dissymmetry would likely be evident by 1H NMR. Furthermore, these stereochemical analogues are separable by HPLC purification based on our previous work.1 Throughout this reinvestigation, the combined yields of the three major MCOs were high, suggesting that any chain cleavage process was likely not random.

To rule out the impure starting material as the source of adventitious didepsipeptide, the tetradepsipeptide was purified via reversed-phase HPLC. There was no indication of didepsipeptide during this purification, and the material purified in this manner produced the same array of products from MCO (Scheme 4, eq 1). HPLC-purified tetradepsipeptide was subjected to the same purification again, and no HPLC-induced decomposition was noted. We next tested the hypothesis that residual water in the reaction may cleave the central ester bond in the tetradepsipeptide in the MCO reaction or, separately, affect the level of possible anhydride species formed during the reaction. The formation of anhydrides in the Mitsunobu reaction has been studied,33 and this pathway would be expected to form non-symmetrical CODs that we have yet to detect. Reactions with the tetradepsipeptide under dry conditions (Scheme 4, eq 2) or in the presence of a variety of drying agents did not decrease the amount of 18-membered ring or 30-membered ring formed in the reaction. Moreover, when varying amounts of water were added to the reaction (ranging from 0.25 to 5 equivalents), it yielded only the starting material (no macrocycles) (Scheme 4, eq 3). To explore the possibility that DIAD or PPh3 could cause cleavage of the tetradepsipeptide, we added each reagent to tetradepsipeptide in benzene for 12 h. Both of these experiments returned only unreacted tetradepsipeptide with no sign of decomposition (Scheme 4, eqs 4 and 5). Additionally, no decomposition was noted after stirring the starting material alone in benzene for several days. Hypothesizing that the rings could decompose after formation and then recyclize to make smaller macrocycles, the authentic 36-membered ring was re-subjected to MCO reaction conditions (Scheme 4, eq 6). Again, no sign of decomposition was observed, and the starting material was recovered quantitatively. Finally, an experiment was investigated in which partially protected MOMO–tetradepsipeptide–CO2H (8) was subjected to MCO reactions to determine if it was self-cleaving from the carboxylic acid terminus concomitant with the MCO reaction (Scheme 5). While this experiment did result in mostly recovered starting material (65%), the protected didepsipeptide 6 was isolated by HPLC purification in 13% yield.34

Scheme 4.

Scheme 4.

Summary of Experiments Probing the Correlation of Conditions to the Formation of Paradoxical Ring-Size CODs.

Scheme 5.

Scheme 5.

Evidence That MCO Conditions Can Promote Depsipeptide Self-Cleavage.

The observation of protected depsipeptide instability under the typical Mitsunobu reaction conditions employed in MCO provides a tangible basis on which to consider the 18-from-12 paradox. The formation of 6 from 8 does not readily explain the formation of the 18-membered ring (16) since an analogous cleavage of 1 would lead to 5 and epi-5. While didepsipeptide 5 would give the expected heterochiral CODs (l,d), didepsipeptide epi-5 would lead to either homochiral CODs (d,d) or an epimer of 12 or 16, both distinguishable by 1H NMR. However, our best rationale for the formation of 6 from 8 (Scheme 6) involves carboxylate addition to the central ester, collapse of this tetrahedral intermediate (a hemiorthoester)35 to the anhydride isomer 18, and then either anhydride hydrolysis, or similar transacylation to form an active ester. Despite our careful reexamination, this is our best evidence that small amounts of the linear depsipeptide chain are self-cleaving under the MCO reaction conditions.

Scheme 6.

Scheme 6.

Mechanistic Hypothesis for C-Terminal Didepsipeptide Self-Cleavage from the Tetradepsipeptide Prior to MCO.

This occurrence, however, raises the question of why evidence of diastereomer formation is not observed. While the protected depsipeptide 6 is observed, didepsipeptide epi-5, where the α-hydroxy chiral center has already been inverted in a Mitsunobu reaction, should accompany it. The α-hydroxy amide configuration, therefore, is a label to determine the source of an MCO product; the expected heterochiral CODs (l,d) result from single-inversion Mitsunobu reactions. While the exact mechanism of cleavage is not certain, we outline a general pathway to explain this. These mechanistic proposals all stem from the formation of an anhydride (e.g., 18). In one scenario, the pendant alcohol resulting from the anhydride formation could intramolecularly cleave the anhydride, resulting in diketomorpholine 19 and the dipeptide (5) with the correct stereochemistry to participate in the MCO reaction to generate the heterochiral CODs (l,d) observed. The 18-and 30-membered ring formation support the presence of didepsipeptide 5, which could react with the tetradepsipeptide and octadepsipeptide, respectively, to form the requisite COD linear precursors (pre-18, pre-30). Unfortunately, we have yet to isolate any diketomorpholine products from these reactions, and it is not clear if any are present in the crude reaction mixtures (1H NMR) for these reactions, given their complexity.36 We speculate that this occurs with the tetradepsipeptide, but it is plausible that an analogous cleavage process could occur on a variety of linear depsipeptides.

Alternatively, a C-terminal didepsipeptide cleavage could occur on any linear oligomer (12, 24, 36, etc.). If an anhydride were to form at the C-terminal end (similar intermediate to 17 in Scheme 6), one might imagine a disproportionation mechanism, in which a single didepsipeptide unit is transferred to another linear depsipeptide (Figure 6). This possibility provides an explanation why the diketomorpholine, homochiral macrocycles, and diastereomeric macrocycles are not observed under these reaction conditions. This type of mechanism could be related to a competitive transesterification process responsible for chain scission in the polymerization of hydroxy acids.37,38

Figure 6.

Figure 6.

Possible disproportionation mechanism to paradoxical ring sizes.

Isothermal Titration Calorimetry and Correlations to Ring-Size Influences during MCO by Salt Additives.

The preparation of authentic 36-membered COD 12 and the reassignment of mass initially attributed to 12, but reallocated to 18-membered COD 16, demanded a reevaluation of the isothermal titration calorimetry (ITC) measurements from this series. We noted that our reported ion binding measurements for these macrocycles exhibited very different patterns for the 18-membered ring and putative 36-membered ring.2 With the knowledge that these compounds should be identical, we sought an understanding of the measurement difference with a prejudice toward possible contaminant(s). This careful analysis led to the finding that some of the original samples used for ITC contained (small) residual amounts of trifluoroacetic acid (TFA) from reversed-phase HPLC purification. Detection of this residual TFA was possible only via 19F NMR since the 1H NMR of these samples can appear identical when there are only small amounts of TFA in the sample. Furthermore, this finding was unexpected since all of these samples had been subjected to a standard base wash in a work-up after the HPLC purification. Hence, these macrocycles exhibit a strong propensity to bind TFA.

Armed with this knowledge, the original macrocycle samples were extensively base-washed until TFA removal was confirmed by 19F NMR, prior to repeating the ITC experiments. Truly, TFA-free samples labeled 18- or “36”-membered ring from the MCO reactions then provided the same reproducible isotherms (see the Supporting Information).

A complete series of TFA-free samples of 18-, 24-, 30-, and 36-membered rings were next analyzed by ITC. This data is provided in Figure 7. Overall, while the remeasured data still delineates cases of strong, selective binding, the binding data for this particular depsipeptide does not support40 a templating effect with sodium, potassium, or cesium. The benefit of adding cesium is clear, however, as the combined yield of three CODs is 86%, compared to 57% without cesium salt additive. This increase in yield results from the significant increase in production of the paradoxical 18-membered ring (46% vs 29%).

Figure 7.

Figure 7.

Comparison of MCO reaction yields (di- and tetradepsipeptide monomers) and ITC data (corrected). Association constant (Ka), binding stoichiometry (N), and enthalpy of binding (ΔH either positive (+) or negative (−)) were determined by ITC using methanol solutions, whereas MCO was performed in benzene. See the Supporting Information for complete details. K salt is KSCN for ITC measurements and KCl for MCO reaction conditions, Na is NaPF6 for ITC measurements, NaBF4 for MCO reaction conditions, and Cs salt is CsCl for both ITC data and MCO reaction conditions.39 A value of <1.00 × 103 indicates that no binding was observed, and therefore fields denoted with N/A (not applicable) were not measured. ITC data for the 36-membered ring was collected using material prepared via the synthesis in Scheme 2 since this ring size is not formed in MCO reaction. Yields are averages of three or more replicates for each case, based on a theoretical yield of 18.3 mg (from didepsipeptide) and 19.1 mg (from tetradepsipeptide). The maximal observed yield variation was 12% (~2 mg) in the didepsipeptide series and 13.5% (2.6 mg) in the tetradepsipeptide series. ITC data for the 24-membered ring is the same as that previously reported;2 these macrocycle samples contained no TFA and were replicated 10 times.

As we noted originally, the ITC measurements could only be comprehensively obtained by the use of methanol solvent. This contrasts the use of benzene for MCO reactions, although the salt additives are used at saturating concentrations in benzene. Attempts to prepare ITC runs in other organic solvents or solvent mixtures led to insufficient solubility of either the salts or the salt–macrocycle complexes. At the time of initial investigation with this series, we noted that the solvent can have a strong effect on the magnitude of binding affinity; however, the goal was to observe possible trends in relative binding interactions. The trends observed in methanol, with the corrected ring sizes, still hold true. While it is disappointing that a correlation of affinity and MCO selectivity cannot be concluded for this particular MCO reaction (Figure 1A),40 the potential ion binding abilities of this series of CODs remain unchanged.

We have carefully reviewed other series of MCO reactions and have no reason to suspect TFA impurities in any other cases. First and foremost, there are no other series of macrocycles that display the same data that lead us to believe there was a mischaracterization. Each COD reported in other series have very distinct NMRs and significantly different HPLC retention times. For ITC data, one factor that indicated the presence of TFA in samples was isotherms that exhibited atypical S-shaped curves. While this is characteristic of aggregate formation at the beginning of titration41,42 and our initial suspicion, we now know that TFA impurity was responsible. No other series of COD isotherms appear to display this behavior. Additionally, some samples used in original experiments were still on hand and confirmed to be TFA-free by 19F NMR. The potential application of ion binding in future chemistry remains promising, and this study highlights an important consideration for any depsipeptide MCO.

The ability of residual TFA, despite routine base wash of HPLC fractions, to adversely affect COD behavior is evident. While this was a problem specifically with depsipeptides here, routine 19F NMR analysis should be a consideration when using reversed-phase HPLC purification buffered with TFA. Several patterns of behavior have been documented here, which could be essential when 19F NMR is not accessible.

CONCLUSIONS

In summary, we have re-examined the single case in which we isolated a material from an MCO reaction, utilizing a tetradepsipeptide, with the original assignment as a 36-membered ring. Through synthesis of an authentic 36-membered ring using a stepwise route and analysis of its product, we have reassigned the “original 36” material to be the 18-membered ring, a size not possible by direct oligomerization of a tetradepsipeptide (12-atom chain). While its relative stereochemistry confirms that it is formed during an MCO, its size is paradoxical since a hexadepsipeptide cannot form directly from a tetradepsipeptide. Our reanalysis of this case illuminates the paradox and explores several possible mechanisms, narrowing to an elongation that succumbs to an unusually competitive self-cleavage step. Importantly, the additive effects that have been shown to template certain MCO reactions are re-evaluated in this case, highlighting the possible insensitivity40 of this specific tetradepsipeptide conversion to added sodium or potassium salts.

EXPERIMENTAL SECTION

General Information.

Glassware was flame-dried under vacuum for all non-aqueous reactions. All reagents and solvents were commercial grade and purified prior to use when necessary. Benzene and dichloromethane (CH2Cl2) were dried by passage through a column of activated alumina as described by Grubbs.43 Flash column chromatography was performed using Sorbent Technologies 230–400 mesh silica gel with solvent systems indicated. Analytical thin layer column chromatography was performed using Sorbent Technologies 250 μm glass backed UV254 silica gel plates and were visualized by fluorescence upon 250 nm radiation and/or the by use of ceric ammonium molybdate (CAM), phosphomolybdic acid (PMA), or potassium permanganate (KMnO4). Solvent removal was affected by rotary evaporation under vacuum (~25–40 mm Hg). All extracts were dried with MgSO4 or NaSO4 unless otherwise noted. Preparative HPLC was performed on an Agilent 1260 system (column: Zorbax Eclipse XDB-C18; 21.2 mm × 150 mm, 5 μm, flow rate 8 mL/min) with a 210 nm monitoring wavelength and acetonitrile/water (+0.1% TFA) gradient as indicated. Nuclear magnetic resonance (NMR) spectra were acquired on a Bruker AV-400 (400 MHz), Bruker DRX500 (500 MHz), or Bruker AV II-600 (600 MHz) instrument. Chemical shifts are measured relative to residual solvent peaks as an internal standard set to δ 7.26 and δ 77.16 (CDCl3), unless otherwise specified. Mass spectra were recorded on a high-resolution Thermo Electron Corporation MAT 95XP-Trap mass spectrometer by use of the ionization method noted by the Indiana University Mass Spectrometry Facility. IR spectra were recorded on a Nicolet Avatar 360 spectrophotometer and are reported in wave-numbers (cm−1) as neat films on a NaCl plate (transmission). Melting points were measured using an OptiMelt automated melting point system (Stanford Research Systems) and are not corrected. Chiral HPLC analysis was conducted on an Agilent 1100 series Infinity instrument using the designated ChiralPak column. Optical rotations were measured on either a Jasco P-2000 polarimeter or an AUTOPOL III (Rudolph Research Analytical) polarimeter. ITC measurements were performed using a Microcal PEAQ-ITC instrument (MicroCal Inc. Northampton, MA). All experiments were performed at 25 °C in anhydrous methanol. Optimal settings for consistent results in methanol include reference power set to 5 μcal/mol, stir speed set to 1000 rpm, and instrument feedback set to low.

General Procedures.

Methoxymethylene Ether (MOM) Deprotection.

A flame-dried round-bottom flask was charged with the depsipeptide (1 equiv) dissolved in dry dichloromethane (0.05 M), pentamethyl benzene (3 equiv), and BF3·Et2O (3 equiv). The reaction was allowed to stir for 50 min at ambient temperature. The crude reaction mixture was quenched with satd aq NaHCO3, washed with brine, dried, concentrated, and subjected to flash column chromatography to afford the alcohol.

Benzyl Deprotection.

A round-bottom flask was charged with the depsipeptide (1 equiv) dissolved in methanol (0.1 M) and treated with 10% Pd/C (10 mol %). The reaction flask was evacuated with a light vacuum (~40 Torr). Hydrogen (balloon) was added, and then the flask was cycled through a light vacuum three times. The reaction was stirred for 1.5 h. The crude reaction mixture was filtered through Celite and concentrated to afford the carboxylic acid.

Didepsipeptide MCO.

A flame-dried round-bottomed flask under inert atmosphere was charged with seco-acid (1 equiv), PPh3 (6 equiv), and benzene (0.02 M). DIAD (5 equiv) was then added to the stirred solution in 15 aliquots over 120 min. The reaction was allowed to stir at ambient temperature for 24 h and then concentrated to afford a residue that was subjected to the MCO purification protocol.

Tetradepsipeptide MCO.

A flame-dried round-bottomed flask under inert atmosphere was charged with seco-acid (1 equiv), PPh3 (3 equiv), and benzene (0.005 M). DIAD (2.5 equiv) was then added to the stirred solution in 5 aliquots over 40 min. The reaction was allowed to stir at ambient temperature for 24 h and then concentrated to afford a residue that was subjected to the MCO purification protocol.

General MCO Purification Information.

The crude residue was purified following the procedure reported previously with minor changes.1 The crude residue was filtered through a silica gel plug (2 cm × 9 cm) to remove excess Mitsunobu reagents. A stepwise MeOH/DCM gradient was used, and the mixture was collected into two fractions (Fraction 1: 0.5–1% MeOH in DCM; Fraction 2: 20% MeOH in DCM). After analysis by 1H NMR, fraction 1 was discarded as Mitsunobu byproducts only. Fraction 2 was subjected to HPLC purification. Preparative HPLC was performed on an Agilent 1260 system (column: Zorbax Eclipse XDB-C18; 21.2 mm × 150 mm, 5 μm, 8 mL/min flow rate) with 210, 254, and 198 nm monitoring wavelength. The gradient used is listed in the Supporting Information. HPLC fractions containing depsipeptides (macrocyclic or linear) were then subjected to extractive workup with ethyl acetate. These compounds are sensitive to cleavage under the high-heat conditions necessary to remove the acidic water–acetonitrile solvent system via rotary evaporation. Additionally, these compounds retain polar solvents and TFA, so the washes (water for acidic depsipeptides and satd aq NaHCO3 for non-acidic depsipeptides) are necessary for complete removal.

Experimental and Characterization Data for Reported Compounds.

((S)-2-((((R)-2-Hydroxyheptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alanine (1).

graphic file with name nihms-1757375-f0002.jpg

Following the general benzyl deprotection procedure, the benzyl protected tetradepsipeptide (405 mg, 799 μmol) afforded the acid (312 mg, 95%) as a pale-yellow oil. All spectral data are in agreement with literature values.1

(R)-2-(Methoxymethoxy)-1-nitroheptane (2).

graphic file with name nihms-1757375-f0003.jpg

Following the Evans protocol,44 IndaBOX ((3aR,3a′R,8a-S,8a′S)-2,2′-(propane-2,2-diyl)bis(3a,8a-dihydro-8H-indeno-[1,2-d]oxazole)) (321 mg, 897 μmol) and Cu(OAc)2·H2O (162 mg, 813 μmol) were stirred at ambient temperature in isopropanol (32.6 mL) for 1 h. The cerulean blue solution was then cooled to 0 °C, and hexanal (2.00 mL, 16.3 mmol) was added and allowed to stir for 10 m before nitromethane (9.95 mL, 163 mmol) addition. After stirring for 4 days at ambient temperature, the reaction was quenched dropwise at 0 °C with 1 N HCl and the aqueous layer was extracted with CH2Cl2. Following drying and concentration under reduced pressure, the crude alcohol was dissolved in CHCl3 (81.6 mL), treated with P2O5 (23.1 g, 163 mmol) and dimethoxymethane (33.9 mL, 326 mmol), and stirred at ambient temperature overnight. The reaction mixture was diluted with DCM and decanted from the solid. The organic layer was then washed with NaHCO3 and brine. The organic layers were dried and concentrated to afford an oil that was subjected to flash column chromatography (SiO2, 3–6% diethyl ether in hexanes) to afford the title compound as a pale-yellow oil (2.54 g, 76%, 2 steps). All spectral data are in agreement with literature values.1

Benzyl ((R)-2-(Methoxymethoxy)heptanoyl)-d-alaninate (3).

graphic file with name nihms-1757375-f0004.jpg

A round-bottom flask was charged with 2 (289 mg, 1.41 mmol), benzyl d-alaninate (515 mg, 2.87 mmol), DME (5.2 mL), and H2O (127 μL, 7.04 mmol). The mixture was then treated with DBTCE (550 mg, 1.69 mmol) followed by NaI (21.1 mg, 141 μmol), K2CO3 (397 mg, 2.87 mmol), and O2 (balloon). The heterogeneous solution was vigorously stirred for 2 days at ambient temperature and then treated at 0 °C with 1 N HCl and poured into CH2Cl2. The aqueous layer was extracted with CH2Cl2, and the combined organic layers were washed with satd aq Na2S2O3, dried, and concentrated. The crude residue was subjected to flash column chromatography (SiO2, 15% ethyl acetate in hexanes) to afford the amide as an orange solid (231 mg, 47%).45 All spectral data are in agreement with literature values.1

Benzyl ((R)-2-Hydroxyheptanoyl)-d-alaninate (4).

graphic file with name nihms-1757375-f0005.jpg

Following the general MOM deprotection procedure, 3 (1.07 g, 3.02 mmol) afforded a crude pale-yellow solid. Flash column chromatography (SiO2, 15–40% ethyl acetate in hexanes) afforded the alcohol (805 mg, 87%) as a pale-yellow oil. All spectral data are in agreement with literature values.1

((R)-2-Hydroxyheptanoyl)-d-alanine (5).

graphic file with name nihms-1757375-f0006.jpg

Following the general benzyl deprotection procedure, 4 (441 mg, 1.44 mmol) afforded the deprotected acid (310 mg, 99%) as a pale-yellow oil. All spectral data are in agreement with literature values.1

((R)-2-(Methoxymethoxy)heptanoyl)-d-alanine (6).

graphic file with name nihms-1757375-f0007.jpg

Following the general benzyl deprotection procedure, 5 (776 mg, 2.21 mmol) afforded the deprotected acid (566 mg, 99%) as a pale-yellow oil. All spectral data are in agreement with literature values.1

Benzyl ((S)-2-((((R)-2-(Methoxymethoxy)heptanoyl)-d-alanyl)-oxy)heptanoyl)-d-alaninate (7).

graphic file with name nihms-1757375-f0008.jpg

A flame-dried round-bottom flask was charged with PPh3 (1.04 g, 3.94 mmol), DIAD (775 μL, 3.94 mmol), and benzene (39.4 mL). The reaction was allowed to stir for 30 min at ambient temperature. Compound 4 (604 mg, 1.97 mmol) was added followed by 6 (566 mg, 2.17 mmol), and the reaction was allowed to stir for 24 h. The crude reaction mixture was concentrated and subjected to flash column chromatography (SiO2, 15–60% ethyl acetate in hexanes) to afford the tetradepsipeptide (857 mg, 80%) as a pale-yellow solid. All spectral data are in agreement with literature values.1

((S)-2-((((R)-2-(Methoxymethoxy)heptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alanine (8).

graphic file with name nihms-1757375-f0009.jpg

Following the general benzyl deprotection procedure, the protected tetradepsipeptide (200 mg, 363 μmol) afforded the acid (165 mg, 99%) as a pale-yellow oil. All spectral data are in agreement with literature values.1

Benzyl ((S)-2-((((R)-2-Hydroxyheptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alaninate (9).

graphic file with name nihms-1757375-f0010.jpg

Following the general MOM deprotection procedure, the protected tetradepsipeptide (500 mg, 908 μmol) afforded a crude pale-yellow solid. Flash column chromatography (SiO2, 15–40% ethyl acetate in hexanes) afforded the alcohol (405 mg, 88%) as a pale-yellow oil. All spectral data are in agreement with literature values.5

Benzyl ((S)-2-((((S)-2-((((S)-2-((((R)-2-(Methoxymethoxy)-heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alaninate (10).

graphic file with name nihms-1757375-f0011.jpg

A flame-dried round-bottom flask was charged with PPh3 (85.5 mg, 326 μmol), DIAD (64.2 μL, 326 μmol), and benzene (3.3 mL). The reaction was allowed to stir for 30 min at ambient temperature. Compound 9 (82.6 mg, 163 μmol) was added, followed by 8 (82.6 mg, 179 μmol), and the reaction was allowed to stir for 24 h. The crude reaction mixture was concentrated and subjected to flash column chromatography (SiO2, 20–40% ethyl acetate in hexanes) to afford the octadepsipeptide (124 mg, 80%) as a colorless oil. All spectral data are in agreement with literature values.5

Benzyl ((S)-2-((((S)-2-((((S)-2-((((S)-2-((((S)-2-((((R)-2-(Methoxy-methoxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alaninate (11).

graphic file with name nihms-1757375-f0012.jpg

A flame-dried round-bottom flask was charged with PPh3 (56.7 mg, 216 μmol), DIAD (43 μL, 216 μmol), and benzene (2.2 mL). The reaction was allowed to stir for 30 m at ambient temperature. The octadepsipeptide free alcohol (98.0 mg, 108 μmol) was added followed by 8 (54.8 mg, 119 μmol), and the reaction was allowed to stir for 24 h. The crude reaction mixture was concentrated and subjected to flash column chromatography (SiO2, 20–50% ethyl acetate in hexanes) to afford the dodecadepsipeptide (108 mg, 72%) as a colorless oil. [α]D2023(c0.86,CHCl3); Rf = 0.37 (40% EtOAc/hexanes); IR (film) 3311, 2930, 2862, 1752, 1657, 1539, 1456, 1380, 1311, 1195, 1155 cm−1; 1H NMR (600 MHz, CDCl3) δ 7.75 (d, J = 5.3 Hz, 2H), 7.71 (d, J = 4.1 Hz, 1H), 7.66 (d, J = 6.0 Hz, 1H), 7.39 (d, J = 7.4 Hz, 1H), 7.36–7.28 (m, 5H), 7.00 (d, J = 5.7 Hz, 1H), 5.21–5.11 (m, 5H), 5.19 (d, J = 12.9 Hz, 1H), 5.13 (d, J = 12.4 Hz, 1H), 4.70 (dd, J = 8.8, 6.7 Hz, 2H), 4.63 (dq, J = 7.3, 7.3 Hz, 1H), 4.33–4.17 (m, 5H), 4.04 (dd, J = 6.3, 4.8 Hz, 1H), 3.42 (s, 3H), 1.96–1.69 (series of m, 12H), 1.50 (d, J = 7.4 Hz, 3H), 1.498 (d, J = 7.0 Hz, 3H), 1.496 (d, J = 7.1 Hz, 3H), 1.491 (d, J = 7.4 Hz, 3H), 1.486 (d, J = 7.2 Hz, 3H), 1.45 (d, J = 7.4 Hz, 3H), 1.41–1.27 (m, 36H), 0.90–0.84 (m, 18H); 13C{1H} NMR (150 MHz, CDCl3): Due to extensive overlap of methylene peaks, line listing is not given. Refer to the image of the 13C spectrum;46 HRMS (ESI) m/z: [M + H]+ calcd for C69H115N6O20 1347.8161; Found 1347.8120.

(3R,6S,9R,12S,15R,18S,21R,24S,27R,30S,33R,36S)-3,9,15,21,27-,33-Hexamethyl-6,12,18,24,30,36-hexapentyl-1,7,13,19,25,31-hexaoxa-4,10,16,22,28,34-hexaazacyclohexatriacontane-2,5,8,11,14-,17,20,23,26,29,32,35-dodecaone (12).

graphic file with name nihms-1757375-f0013.jpg

A flame-dried round-bottom flask was charged with S1 (10.0 mg, 8.2 μmol) dissolved in dichloromethane (100 μL). The seco-acid was diluted with benzene (8.2 mL) and PPh3 (12.9 mg, 41 μmol) was added. DIAD (8.1 μL, 50 μmol) was added over 40 min in 5 aliquots. The reaction was allowed to stir at ambient temperature for 12 h. The crude mixture was concentrated and passed through a silica plug (F1: 0.5% MeOH/DCM, 1% MeOH/DCM; F2: 20% MeOH/DCM). Fraction 2 was purified by preparative HPLC (45–95% aqueous acetonitrile, 210 nm, flow rate: 20 mL/min, Rt = 9.8 m) to afford the 36-membered macrocycle (4.2 mg, 42%) as a colorless oil. [α]D2313(c0.40,CHCl3); Rf = 0.60 (4% MeOH/DCM); IR (film) 3316, 2955, 2926, 2857, 1750, 1657, 1544, 1457, 1379, 1193, 1156, 1064 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.25 (d, J = 7.8 Hz, 1H), 5.16 (dd, J = 8.5, 4.1 Hz, 1H), 4.45 (dq, J = 7.0, 7.0 Hz, 1H), 1.93–1.70 (series of m, 2H), 1.47 (d, J = 7.3 Hz, 3H), 1.40–1.28 (m, 6H), 0.87 (t, J = 6.7 Hz, 3H); 13C{1H} NMR (150 MHz, CDCl3) ppm 171.6, 170.6, 74.4, 48.7, 31.5, 31.4, 24.7, 22.5, 17.2, 14.1; HRMS (ESI) m/z: [M + H]+ calcd for C60H103N6O18 1195.7323; Found 1195.7355.

Benzyl ((S)-2-((((S)-2-((((R)-2-(Methoxymethoxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alaninate (13).

graphic file with name nihms-1757375-f0014.jpg

A flame-dried round-bottom flask was charged with PPh3 (97.5 mg, 372 μmol), DIAD (73.2 μL, 372 μmol), and benzene (3.7 mL). The reaction was allowed to stir for 30 min at ambient temperature. Compound 9 (94.0 mg, 186 μmol) was added followed by 6 (53.3 mg, 204 μmol), and the reaction was allowed to stir for 24 h. The crude reaction mixture was concentrated and subjected to flash column chromatography (SiO2, 20–30% ethyl acetate in hexanes) to afford the hexadepsipeptide (130 mg, 93%) as a colorless oil. [α]D20+2.3(c1.42,CHCl3); Rf = 0.33 (40% EtOAc/hexanes); IR (film) 3295, 2952, 2956, 2857, 1749, 1655, 1541, 1457, 1378, 1155 cm−1; 1H NMR (600 MHz, CDCl3) δ 7.51 (d, J = 4.8 Hz, 1H), 7.36–7.30 (m, 5H), 7.34 (d, J = 7.0 Hz, 1H), 6.99 (d, J = 5.6 Hz, 1H), 5.20 (d, J = 12.4 Hz, 1H), 5.20–5.18 (m, 2H), 5.13 (d, J = 12.4 Hz, 1H), 4.70 (d, J = 6.7 Hz, 1H), 4.68 (d, J = 6.7 Hz, 1H), 4.62 (dq, J = 7.3, 7.3 Hz, 1H), 4.32 (dq, J = 7.0, 5.6 Hz, 1H), 4.31 (dq, J = 7.1, 5.4 Hz, 1H), 4.04 (dd, J = 6.4, 4.8 Hz, 1H), 3.42 (s, 3H), 1.97–1.71 (series of m, 6H), 1.48 (d, J = 7.2 Hz, 3H), 1.48 (d, J = 7.2 Hz, 3H), 1.44 (d, J = 7.3 Hz, 3H), 1.40–1.21 (br m, 18H), 0.89–0.85 (m, 9H); 13C{1H} NMR (150 MHz, CDCl3) ppm 173.6, 172.6, 172.5, 172.4, 170.5, 169.8, 135.8, 128.6, 128.3, 128.2, 96.5, 77.7, 74.4, 74.3, 67.0, 56.4, 49.4, 49.0, 48.1, 32.7, 31.6, 31.5, 31.39, 31.36, 29.8, 24.7, 24.5, 24.4, 22.6, 22.52, 22.47, 17.5, 16.9, 16.4, 14.12, 14.09, 14.0; HRMS (ESI) m/z: [M + Na]+ calcd for C39H63N3NaO11 772.4360; Found 772.4361.

Benzyl ((S)-2-((((S)-2-((((R)-2-Hydroxyheptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alaninate (14).

graphic file with name nihms-1757375-f0015.jpg

Following the general MOM deprotection procedure, the hexadepsipeptide (130 mg, 173 μmol) afforded a crude pale-yellow solid. Flash column chromatography (SiO2, 20–40% ethyl acetate in hexanes) afforded the alcohol (104.0 mg, 85%) as a colorless oil. [α]D204.1(c1.27,CHCl3); Rf = 0.20 (40% EtOAc/hexanes); IR (film) 3312, 2955, 2927, 2859, 1749, 1655, 1538, 1457, 1381, 1155 cm−1; 1H NMR (600 MHz, CDCl3) δ 7.54 (d, J = 5.1 Hz, 1H), 7.37–7.30 (m, 5H), 7.30 (d, J = 7.6 Hz, 1H), 7.04 (d, J = 5.6 Hz, 1H), 5.20 (d, J = 12.4 Hz, 1H), 5.19–5.16 (m, 2H), 5.13 (d, J = 12.4 Hz, 1H), 4.63 (dq, J = 7.3, 7.3 Hz, 1H), 4.32 (dq, J = 7.0, 5.6 Hz, 1H), 4.31 (dq, J = 7.1, 5.4 Hz, 1H), 4.11 (ddd, J = 8.2, 8.2, 4.3 Hz, 1H), 2.85 (d, J = 4.9 Hz, 1H), 1.97–1.59 (series of m, 6H), 1.47 (d, J = 7.2 Hz, 3H), 1.47 (d, J = 7.1 Hz, 3H), 1.44 (d, J = 7.3 Hz, 3H), 1.40–1.22 (br m, 18H), 0.90–0.85 (m, 9H); 13C{1H} NMR (150 MHz, CDCl3) ppm 174.9, 172.7, 172.6, 172.4, 170.5, 169.9, 135.7, 128.7, 128.4, 128.2, 74.42, 74.36, 72.2, 67.0, 49.3, 49.0, 48.1, 34.6, 31.7, 31.6, 31.5, 31.40, 31.37, 24.69, 24.67, 24.5, 22.6, 22.52, 22.48, 17.5, 16.8, 16.5, 14.11, 14.08, 14.0; HRMS (ESI) m/z: [M + Na]+ calcd for C37H59N3NaO10 728.4098; Found 728.4083.

((S)-2-((((S)-2-((((R)-2-Hydroxyheptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanine (15).

graphic file with name nihms-1757375-f0016.jpg

Following the general benzyl deprotection procedure, the hexadepsipeptide (140 mg, 198 μmol) afforded deprotected acid (119 mg, 98%) as a colorless oil. [α]D200.7(c0.98,CHCl3); Rf = 0.17 (80% EtOAc/hexanes); IR (film) 3310, 3073, 2930, 2863, 1749, 1655, 1540, 1457, 1380, 1203, 1155, 1064 cm−1; 1H NMR (400 MHz, (CD3)2SO) δ 8.30 (d, J = 7.3 Hz, 1H), 8.24 (d, J = 7.0 Hz, 1H), 7.93 (d, J = 5.6 Hz, 1H), 4.94 (dd, J = 7.6, 4.7 Hz, 1H), 4.91 (dd, J = 7.5, 4.9 Hz, 1H), 4.35 (dq, J = 7.1, 4.5 Hz, 1H), 4.34 (dq, J = 7.0, 4.6 Hz, 1H), 4.12 (dq, J = 7.0, 7.0 Hz, 1H), 3.88 (dd, J = 7.7, 4.0 Hz, 1H), 1.75–1.39 (series of m, 6H), 1.34 (d, J = 7.2 Hz, 6H), 1.33–1.16 (br m, 18H), 1.25 (d, J = 7.2 Hz, 3H), 0.87–0.83 (m, 9H), [OH and CO2H not observed]; 13C{1H} NMR (150 MHz, (CD3)2SO) ppm 174.6, 173.7, 171.8, 171.5, 169.5, 168.4, 73.6, 73.4, 70.6, 47.8, 47.73, 47.67, 34.1, 31.3 (2C), 31.1, 30.79, 30.77, 24.2, 24.0, 23.8, 22.0, 21.88, 21.86, 17.4, 17.0, 16.7, 13.9, 13.8 (2C); HRMS (ESI) m/z: [M + Na]+ calcd for C30H53N3NaO10 638.3629; Found 638.3624.

(3R,6S,9R,12S,15R,18S)-3,9,15-Trimethyl-6,12,18-tripentyl-1,7,13-trioxa-4,10,16-triazacyclooctadecane-2,5,8,11,14,17-hexaone (16).

graphic file with name nihms-1757375-f0017.jpg

A flame-dried round-bottom flask was charged with the fully deprotected hexadepsipeptide (10.0 mg, 16.2 μmol), PPh3 (12.8 mg, 48.7 μmol), and benzene (1.62 mL). DIAD (8.00 μL, 40.5 μmol) was added in 5 aliquots over 40 min. The reaction was allowed to stir at ambient temperature for 24 h. The crude reaction mixture was concentrated and subjected to the general MCO purification protocol to afford the 18-membered macrocycle (6.0 mg, 62%, Rt = 23.8 min) as an amorphous white solid. [α]D23+25(c0.95,CHCl3); Rf = 0.30 (6% MeOH/DCM); IR (film) 3208, 3046, 2922, 2852, 1756, 1669, 1556, 1456, 1379, 1260, 1211, 1171, 1105 cm−1; 1H NMR (400 MHz, CDCl3) δ 6.98 (d, J = 7.8 Hz, 1H), 5.13 (dd, J = 7.6, 4.7 Hz, 1H), 4.58 (dq, J = 7.4, 7.3 Hz, 1H), 1.93–1.74 (m, 2H), 1.47 (d, J = 7.1 Hz, 3H), 1.40–1.22 (series of m, 6H), 0.88 (t, J = 6.8 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3 ) ppm 171.6, 169.9, 75.0, 48.7, 31.6, 31.5, 24.7, 22.6, 17.5, 14.1; HRMS (ESI) m/z: [M + Na]+ calcd for C30H51N3NaO9 620.3523; Found 620.3518.

((2S)-2-((((2S)-2-((((2S)-2-(((2-((((S)-2-((((R)-2-Hydroxyheptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)-heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanyl)oxy)heptanoyl)-d-alanine (S1).

graphic file with name nihms-1757375-f0018.jpg

A flame-dried vial was charged with N-H depsipeptide (48.0 mg, 35.6 μmol), pentamethyl benzene (26.4 mg, 178 μmol), and dry DCM (712 μL). BF3·Et2O (22 μL, 180 μmol) was then added to the reaction mixture. The reaction was allowed to stir at ambient temperature for 45 m, and it was then quenched by the dropwise addition of satd aq NaHCO3. The aqueous layer was extracted with DCM. The organic layer was dried and concentrated. The crude reaction mixture was concentrated and subjected to flash column chromatography (SiO2, 40% ethyl acetate in hexanes) to afford the alcohol as a colorless oil (39.4 mg, 30.2 μmol). This material was loaded into a vial and then treated with 10% Pd/C (39.4 mg, l mass equiv.) and dissolved in a 1:1 MeOH/CH2Cl2 mixture (302 μL). The reaction vial was evacuated and backfilled with hydrogen (balloon), and this cycle was repeated. The reaction was allowed to stir for 35 min. After purging the flask with nitrogen, the crude reaction mixture was filtered through Celite and concentrated. The reaction mixture was then purified by preparative HPLC (45–95% aqueous acetonitrile, 210 nm, flow rate: 20 mL/min, Rt = 9.8 m) to afford the depsipeptide (35 mg, 95% 2-step) as a yellow oil. [α]D20+2.0(c0.34,CHCl3); Rf = 0.26 (4% MeOH/DCM); IR (film) 3315, 2928, 2861, 1752, 1657, 1541, 1457, 1380, 1196, 1155, 1065 cm−1; 1H NMR (600 MHz, CDCl3) δ 7.74 (d, J = 4.8 Hz, 2H), 7.66 (d, J = 5.5 Hz, 1H), 7.64 (d, J = 6.2 Hz, 1H), 7.31 (d, J = 6.9 Hz, 1H), 7.12 (d, J = 5.3 Hz, 1H), 5.23 (dd, J = 8.1, 4.3 Hz, 1H), 5.20 (dd, J = 5.9, 4.0 Hz, 1H), 5.18 (dd, J = 4.9, 3.9 Hz, 1H), 5.16 (dd, J = 7.9, 3.8 Hz, 1H), 5.12 (dd, J = 8.1, 3.7 Hz, 1H), 4.55 (dq, J = 7.2, 7.2 Hz, 1H), 4.45 (dq, J = 6.9, 6.9, 1H), 4.38 (dq, J = 7.2, 7.2 Hz, 1H), 4.31 (dq, J = 7.2, 7.2 Hz, 2H), 4.23 (dq, J = 7.1, 7.0 Hz, 1H), 4.12 (dd, J = 7.9, 3.8 Hz, 1H), 1.97–1.76 (series of m, 12H), 1.67–1.59 (m, 1H), 1.50 (d, J = 7.2 Hz, 6H), 1.49 (d, J = 7.0 Hz, 3H), 1.48 (d, J = 7.9 Hz, 3H), 1.47 (d, J = 7.3 Hz, 3H), 1.46 (d, J = 7.2 Hz, 3H), 1.45–1.27 (m, 36H), 0.89 (t, J = 7.0 Hz, 3H), 0.88 (t, J = 6.9 Hz, 6H), 0.87 (t, J = 6.7 Hz, 6H), 0.86 (t, J = 6.8 Hz, 3H), [CO2H not observed]; 13C{1H} NMR (150 MHz, CDCl3) Due to extensive overlap of methylene peaks, line listing is not given. Refer to the image of the 13C spectrum;46 HRMS (ESI) m/z: [M + H]+ calcd for C60H105N6O19 1213.7429; Found 1213.7387.

(3R,6S,9R,12S,15R,18S,21R,24S)-3,9,15,21-Tetramethyl-6,12,18,24-tetrapentyl-1,7,13,19-tetraoxa-4,10,16,22-tetraazacy-clotetracosan-2,5,8,11,14,17,20,23-octaone (S2).

graphic file with name nihms-1757375-f0019.jpg

Following the tetradepsipeptide MCO general procedure, seco-acid 1 (20.0 mg, 48.0 μmol) was stirred for 24 h at ambient temperature to afford a pale-yellow oil. Preparative HPLC following the general MCO purification protocol afforded the 24-membered macrocycle (8.1 mg, 44%, Rt = 24.8 min) as a white solid. All spectral data are in agreement with literature values.1

(3R,6S,9R,12S,15R,18S,21R,24S,27R,30S)-3,9,15,21,27-Penta-methyl-6,12,18,24,30-pentapentyl-1,7,13,19,25-pentaoxa-4,10,16-,22,28-pentaazacyclotriacontane-2,5,8,11,14,17,20,23,26,29-decaone (S3).

graphic file with name nihms-1757375-f0020.jpg

Following the tetradepsipeptide MCO general procedure, seco-acid 1 (20.0 mg, 48.0 μmol) was stirred for 24 h at ambient temperature to afford a pale-yellow oil. Preparative HPLC following the general MCO purification protocol afforded the 30-membered macrocycle (2.8 mg, 15%, Rt = 35.6 min) as a colorless oil. [α]D2315(c0.65,CHCl3); Rf = 0.35 (4% MeOH/DCM); IR (film) 3315, 2957, 2924, 2853, 1747, 1664, 1547, 1458, 1380, 1260, 1199, 1158, 1064 cm−1; 1H NMR (400 MHz, CDCl3) δ 7.10 (d, J = 6.4 Hz, 1H), 5.10 (dd, J = 8.1, 4.0 Hz, 1H), 4.45 (dq, J = 6.9, 6.9 Hz, 1H), 1.93–1.83 (m, 1H), 1.83–1.68 (m, 1H), 1.45 (d, J = 7.2 Hz, 3H), 1.41–1.20 (m, 6H), 0.88 (t, J = 6.6 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3) ppm 172.1, 170.4, 74.4, 48.7, 31.5, 31.4, 24.8, 22.6, 17.1, 14.1; HRMS (ESI) m/z: [M + Na]+ calcd for C50H85 NaN5O15 1018.5934; Found 1018.5936.

Supplementary Material

Supplementary Material

ACKNOWLEDGMENTS

The authors are grateful to Suzanne Batiste for careful record-keeping, sample preservation, and critical feedback. Research reported in this publication was supported by the National Institute of General Medical Sciences of the National Institutes of Health (GM 063557 and HL151125 (F31 support for ANS)). The Indiana University Mass Spectrometry Facility acknowledges support from the NSF (CHE1726633), and we thank Dr. Jon Karty and Angela Hansen (IUMSC) for helpful HRMS data acquisition and analysis.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.0c03069.

Full experimental details and characterization data for new compounds, additional comparative spectroscopic data, new isothermal titration calorimetry data and isotherms, and a discussion of the 30- vs 60-membered rings (PDF)

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.joc.0c03069

The authors declare no competing financial interest.

Contributor Information

Abigail N. Smith, Department of Chemistry and Vanderbilt Institute of Chemical Biology, Vanderbilt University, Nashville, Tennessee 37235, United States;.

Jeffrey N. Johnston, Department of Chemistry and Vanderbilt Institute of Chemical Biology, Vanderbilt University, Nashville, Tennessee 37235, United States;.

REFERENCES

  • (1).(a) Batiste SM; Johnston JN Rapid Synthesis of Cyclic Oligomeric Depsipeptides with Positional, Stereochemical, and Macrocycle Size Distribution Control. Proc. Natl. Acad. Sci. U. S. A 2016, 14893. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) A correction has been submitted for this manuscript. [Google Scholar]
  • (2).(a) Batiste SM; Johnston JN Evidence for Ion-Templation during Macrocyclooligomerization of Depsipeptides. J. Am. Chem. Soc 2018, 140, 4560–4568. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Batiste SM; Johnston JN Correction to “Evidence for Ion-Templation during Macrocyclooligomerization of Depsipeptides,”. J. Am. Chem. Soc 2021, 143, 6701–6701. [DOI] [PubMed] [Google Scholar]
  • (3).Shiomi K; Matsui R; Kakei A; Yamaguchi Y; Masuma R; Hatano H; Arai N; Isozaki M; Tanaka H; Kobayashi S; Turberg A; Ōmura S Verticilide, a New Ryanodine-Binding Inhibitor, Produced by Verticillium Sp. FKI-1033. J. Antibiot 2010, 63, 77–82. [DOI] [PubMed] [Google Scholar]
  • (4).Monma S; Sunazuka T; Nagai K; Arai T; Shiomi K; Matsui R; Ōmura S Verticilide: Elucidation of Absolute Configuration and Total Synthesis. Org. Lett 2006, 8, 5601–5604. [DOI] [PubMed] [Google Scholar]
  • (5).Batiste SM; Blackwell DJ; Kim K; Kryshtal DO; Gomez-Hurtado N; Rebbeck RT; Cornea RL; Johnston JN; Knollmann BC Unnatural Verticilide Enantiomer Inhibits Type 2 Ryanodine Receptor-Mediated Calcium Leak and Is Antiarrhythmic. Proc. Natl. Acad. Sci. U. S. A 2019, 116, 4810–4815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6). The remaining cases studied in ref. 1 are not suspected to be in error. The MCO of tetradepsipeptide 1 is the only MCO that appeared to produce large macrocycles (36, 60-membered CODs).
  • (7).Lipinski CA; Lombardo F; Dominy BW; Feeney PJ Experimental and Computational Approaches to Estimate Solubility and Permeability in Drug Discovery and Development Settings. Adv. Drug Delivery Rev 1997, 23, 3–25. [DOI] [PubMed] [Google Scholar]
  • (8).Burke SD; O’Donnell CJ; Porter WJ; Song Y Cyclic Hydropyran Oligolides as Preorganized Ligand Arrays: Cumulative Effects of Structural Elements on Shape and Cation Binding. J. Am. Chem. Soc 1995, 117, 12649–12650. [Google Scholar]
  • (9).Burke SD; Heap CR; Porter WJ; Song Y Cyclic Hydropyran Oligolides as Preorganized Ligand Arrays: Modular Assembly of 18-, 24-, 36-, 48-, 54- and 72-Membered Rings Via Iteration and Cyclooligomerization. Tetrahedron Lett. 1996, 37, 343–346. [Google Scholar]
  • (10).Burke SD; McDermott TS; O’Donnell CJ Template Macrolactonization of Trichloroethyl Ester Derivatives Catalyzed by Potassium Salts. J. Org. Chem 1998, 63, 2715–2718. [DOI] [PubMed] [Google Scholar]
  • (11).Barth R; Mulzer J Total Synthesis of Efomycine M. Angew. Chem., Int. Ed 2007, 46, 5791–5794. [DOI] [PubMed] [Google Scholar]
  • (12).Mulzer J; Berger M Total Synthesis of the Boron-Containing Ion Carrier Antibiotic Macrodiolide Tartrolon B. J. Org. Chem 2004, 69, 891–898. [DOI] [PubMed] [Google Scholar]
  • (13).Beeler AB; Acquilano DE; Su Q; Yan F; Roth BL; Panek JS; Porco JA Synthesis of a Library of Complex Macrodiolides Employing Cyclodimerization of Hydroxy Esters. J. Comb. Chem 2005, 7, 673–681. [DOI] [PubMed] [Google Scholar]
  • (14).Su Q; Beeler AB; Lobkovsky E; Porco JA; Panek JS Stereochemical Diversity through Cyclodimerization: Synthesis of Polyketide-Like Macrodiolides. Org. Lett 2003, 5, 2149–2152. [DOI] [PubMed] [Google Scholar]
  • (15).Rothe M; Kreiss W Synthesis of Valinomycin by the Phosphite Method. Angew. Chem., Int. Ed 1973, 12, 1012–1013. [DOI] [PubMed] [Google Scholar]
  • (16).Rothe M; Kreiss W Synthesis of Cyclodepsipeptides by Cyclooligomerization. Angew. Chem., Int. Ed 1977, 16, 113–114. [Google Scholar]
  • (17).Wipf P; Miller CP Total Synthesis of Westiellamide. J. Am. Chem. Soc 1992, 114, 10975–10977. [Google Scholar]
  • (18).Wipf P; Miller CP; Grant CM Synthesis of Cyclopeptide Alkaloids by Cyclooligomerization of Dipeptidyl Oxazolines. Tetrahedron 2000, 56, 9143–9150. [Google Scholar]
  • (19).Black RJG; Dungan VJ; Li RYT; Young PG; Jolliffe KA A Cyclooligomerisation Approach to Backbone-Modified Cyclic Peptides Bearing Guanidinium Arms. Synlett 2010, 2010, 551–554. [Google Scholar]
  • (20).Chakraborty TK; Tapadar S; Raju TV; Annapurna J; Singh H Cyclic Trimers of Chiral Furan Amino Acids. Synlett 2004, 2004, 2484–2488. [Google Scholar]
  • (21).Sokolenko N; Abbenante G; Scanlon MJ; Jones A; Gahan LR; Hanson GR; Fairlie DP Cyclooligomerization of Thiazole-Containing Tetrapeptides. Symmetrical Macrocycles with up to 76 Amino Acids. J. Am. Chem. Soc 1999, 121, 2603–2604. [Google Scholar]
  • (22).Singh Y; Sokolenko N; Kelso MJ; Gahan LR; Abbenante G; Fairlie DP Novel Cylindrical, Conical, and Macrocyclic Peptides from the Cyclooligomerization of Functionalized Thiazole Amino Acids. J. Am. Chem. Soc 2001, 123, 333–334. [DOI] [PubMed] [Google Scholar]
  • (23).Bertram A; Blake AJ; Turiso F. G.-L. d.; Hannam JS; Jolliffe KA; Pattenden G; Skae M Concise Synthesis of Stereodefined, Thiazole—Containing Cyclic Hexa- and Octapeptide Relatives of the Lissoclinums, Via Cyclooligomerisation Reactions. Tetrahedron 2003, 59, 6979–6990. [Google Scholar]
  • (24).Jayaprakash S; Pattenden G; Viljoen MS; Wilson C Synthesis of Novel Proline–Thiazole Based Cyclic Hexa- and Octapeptides. Tetrahedron 2003, 59, 6637–6646. [Google Scholar]
  • (25).Bertram A; Pattenden G Marine Metabolites: Metal Binding and Metal Complexes of Azole-Based Cyclic Peptides of Marine Origin. Nat. Prod. Rep 2007, 24, 18–30. [DOI] [PubMed] [Google Scholar]
  • (26).Bertram A; Hannam JS; Jolliffe KA; González-López de Turiso F; Pattenden G The Synthesis of Novel Thiazole Containing Cyclic Peptides Via Cyclooligomerisation Reactions. Synlett 1999, 1999, 1723–1726. [Google Scholar]
  • (27).Dudin L; Pattenden G; Viljoen MS; Wilson C Synthesis of Novel N-Methylated Thiazole-Based Cyclic Octa- and Dodeca-peptides. Tetrahedron 2005, 61, 1257–1267. [Google Scholar]
  • (28).Watanabe A; Noguchi Y; Hirose T; Monma S; Satake Y; Arai T; Masuda K; Murashima N; Shiomi K; Ōmura S; Sunazuka T Efficient Synthesis of a Ryanodine Binding Inhibitor Verticilide Using Two Practical Approaches. Tetrahedron Lett. 2020, 61, 151699. [Google Scholar]
  • (29).These ions for the 18-membered ring in the original “36”-membered ring characterization data are not within a 5 ppm standard error—this is due to the tuning of the instrument for the larger mass. When tuned for the smaller mass, the error is within the standard range.
  • (30). Spectrum C is the original spectrum from the characterization data in ref 1, now known to contain residual TFA. This spectrum is reproduced to emphasize its subtle differences when compared to the authentic 18-membered ring (spectrum B) and the originally characterized 36-membered ring (spectrum D). This contributed to the belief that two different ring sizes were isolated in that instance.
  • (31). Spectrum D is the original spectrum from the characterization data for the 36-membered ring in ref 1 that is reassigned here as the 18-membered ring.
  • (32).It is important to distinguish that [1/2M36 + Na]+ is referring to a fragmented half of the 36-membered ring, not to be confused with [M36 + 2Na]2+/2, which would indicate a doubly charged ion evident by isotopic peaks differing by only 0.5 mass units.
  • (33).Harvey PJ; von Itzstein M; Jenkins ID the Formation of Anhydrides in the Mitsunobu Reaction. Tetrahedron 1997, 53, 3933–3942. [Google Scholar]
  • (34).Although the protected didepsipeptide was isolated from the tetradepsipeptide cleavage, the corresponding cleavage product has eluded isolation.
  • (35).Literature studies of hemiorthoesters appear limited to five- and six-membered examples. Leading references:; (a) Austin KAB; Banwell MG; Willis AC Synthesis and X-Ray Crystal Structure of a Cyclic Hemiorthoester. ARKIVOC 2006, xiii, 1–7. [Google Scholar]; (b) McClelland RA; Seaman NE; Cramm D Observation of a Stable Hemiortho Ester Anion. Acidity Constant for a Tetrahedral Intermediate. J. Am. Chem. Soc 1984, 106, 4511–4515. [Google Scholar]; (c) Capon B; Ghosh AK Tetrahedral Intermediates. 2. The Detection of Hemiorthoesters in the Hydration of Ketene Acetals and the Mechanism of their Breakdown. J. Am. Chem. Soc 1981, 103, 1765–1768. [Google Scholar]; (d) Deslong-champs P; Chênevert R; Taillefer RJ; Moreau C; Saunders JK the Hydrolysis of Cyclic Orthoesters. Stereoelectronic Control in the Cleavage of Hemiorthoester Tetrahedral Intermediates. Can. J. Chem 1975, 53, 1601–1615. [Google Scholar]; (e) Antonov VK; Shkrob AM; Shemyakin MM Cyclol Formation in Peptide Systems. III. Rearrangement of N-(β-Hydroxypropionyl)-Piperidone into a 10-Membered Cyclodepsipeptide. Tetrahedron Lett. 1963, 4, 439–443. [Google Scholar]; (f) Lin S-Y; Yu H-L; Li M-J Formation of Six-Membered Cyclic Anhydrides by Thermally Induced Intramolecular Ester Condensation in Eudragit E Film. Polymer 1999, 40, 3589–3593. [Google Scholar]; (g) Chandrasekhar S; Kumar HV the Reaction of Aspirin with Base. Tetrahedron Lett. 2011, 52, 3561–3564. [Google Scholar]
  • (36). A homochiral (d,d)-diketomorpholine is observed in the MCO of epi-5 (ref 1) as the major product. While we would expect a heterochiral variant (l,d) in our case, if either were formed here, it should be detectable, owing to the remarkably distinct 1H NMR peaks compared to other CODs.
  • (37).Culkin DA; Jeong W; Csihony S; Gomez ED; Balsara NP; Hedrick JL; Waymouth RM Zwitterionic Polymerization of Lactide to Cyclic Poly(Lactide) by Using N-Heterocyclic Carbene Organocatalysts. Angew. Chem., Int. Ed 2007, 46, 2627–2630. [DOI] [PubMed] [Google Scholar]
  • (38).Szymanski R Molecular Weight Distribution in Living Polymerization Proceeding with Reshuffling of Polymer Segments Due to Chain Transfer to Polymer with Chain Scission, 1. Determination of Kp/Ktr Ratio from DPw/DPn Data. Ideal Reproduction of Polymer Chain Activities, Macromol. Theory Simul. 1998, 7, 27–39. [Google Scholar]
  • (39).Different salts were used for the ITC experiments due to a lack of solubility in MeOH for the salts used in the MCO reaction conditions.
  • (40).It is important to note that one explanation for the paradox involves disproportionation of the “pre-24” chain. This complicates interpretation of the ion binding trends as some (or all) of the paradoxical ring-formation may stem from a rate that involves templation.
  • (41).Darby SJ; Platts L; Daniel MS; Cowieson AJ; Falconer RJ an Isothermal Titration Calorimetry Study of Phytate Binding to Lysozyme. J. Therm. Anal. Calorim 2017, 127, 1201–1208. [Google Scholar]
  • (42).Clavería-Gimeno R; Velazquez-Campoy A; Pey AL Thermodynamics of Cooperative Binding of FAD to Human NQO1: Implications to Understanding Cofactor-Dependent Function and Stability of the Flavoproteome. Arch. Biochem. Biophys 2017, 636, 17–27. [DOI] [PubMed] [Google Scholar]
  • (43).Pangborn AB; Giardello MA; Grubbs RH; Rosen RK; Timmers FJ Safe and Convenient Procedure for Solvent Purification. Organometallics 1996, 15, 1518–1520. [Google Scholar]
  • (44).Evans DA; Seidel D; Rueping M; Lam HW; Shaw JT; Downey CW a New Copper Acetate-Bis(Oxazoline)-Catalyzed, Enantioselective Henry Reaction. J. Am. Chem. Soc 2003, 125, 12692–12693. [DOI] [PubMed] [Google Scholar]
  • (45).The rest of the starting material is recovered as the mono-brominated nitroalkane, which can be resubjected to Umpolung Amide Synthesis:; (a) Shen B; Makley DM; Johnston JN Nature 2010, 465, 1027. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Leighty MW; Shen B; Johnston JN J. Am. Chem. Soc 2012, 134, 15233. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Schwieter KE; Shen B; Shackleford JP; Leighty MW; Johnston JN Org. Lett 2014, 16, 4714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Carbon spectrum has labeled peaks for this compound.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Material

RESOURCES