Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Jun 5.
Published in final edited form as: Org Biomol Chem. 2019 Jun 5;17(22):5615–5632. doi: 10.1039/c9ob01076a

Functionalised Bicyclic Tetramates Derived from Cysteine as Antibacterial Agents

Tharindi D Panduwawala , Sarosh Iqbal †,#, Amber L Thompson , Miro Genov , Alexander Pretsch , Dagmar Pretsch , Shuang Liu §, Richard H Ebright §, Alison Howells ¥, Anthony Maxwell &, Mark G Moloney
PMCID: PMC6686852  NIHMSID: NIHMS1031883  PMID: 31120090

Abstract

Routes to bicyclic tetramates derived from cysteine permitting ready incorporation of functionality at two different points around the periphery of a heterocyclic skeleton are reported. This has enabled the identification of systems active against Gram-positive bacteria, some of which show gyrase and RNA polymerase inhibitory activity. In particular, tetramates substituted with glycosyl side chains, chosen to impart polarity and aqueous solubility, show high antibacterial activity coupled with modest gyrase/polymerase activity in two cases. An analysis of physicochemical properties indicates that the antibacterially active tetramates generally occupy physicochemical space with MW of 300–600, clogD7.4 of −2.5 to 4 and rel. PSA of 11–22%. This work demonstrates that biologically active 3D libraries are readily available by manipulation of a tetramate skeleton.

Introduction

Tetramates are of interest14 principally for their antibacterial activity,5, 6 and we have reported recently that systems derived from the amino acids, serine 1a,7 threonine 1b,8 and cysteine 1c9 provide useful templates for application in synthetic and medicinal chemistry, which in some cases exhibit potent antibacterial activity. Amongst various methodologies to these derivatives,10 our route makes use of the preferential formation of the malonamides of t-butyl oxazolidine/thiazolidine templates, formed as the cis-2,5-isomer 2a-c rather than the alternative trans-2,5- 3a-c, which mimises the steric impact of the bulky t-butyl group, to direct a very efficient chemoselective Dieckmann ring closure (Route a, Scheme 1) in preference to the alternative (Route b, Scheme 1).11 Of interest, though, is that the analogous thiazolidine system derived from cysteine and an aromatic aldehyde behaves differently (Scheme 2), not least because there is a shift in preference away from the cis-thiazolidine 6 to the trans- isomer 7, the latter of which now preferentially cyclizes to give tetramate 8, equivalent to Route b in Scheme 1; this shift is, however, not complete and some of the cis-isomer 6, after epimerization at C5, closes under these conditions leading to the enantiomer of 8, and therefore giving some erosion of enantioselectivity.12 The overall outcome is that acyl tetramates 8 (Scheme 2),13 known for their metal chelating14 and antibacterial activity,14 are accessed directly, rather than by acylation of an unfunctionalised tetramate nucleus.15 Since we had earlier explored in detail the structure activity relationship of tetramates substituted at C-7 but limited to t-butyl substitution at C-2, these thiazolidine compounds provided access for the first time to compounds with an aryl substituent at C-2 and this gave the opportunity to explore the scope of antibacterial bioactivity at this position. The results of that work are reported here.

Scheme 1.

Scheme 1

Scheme 2.

Scheme 2

Results and Discussion

Synthesis of the required tetramate proceeded from cysteine 1c to a mixture of the cis- and trans- diastereomeric malonamides 6a-g and 7a-g, with a preference for the latter trans- isomers in three cases, followed by Dieckmann cyclisation to 7-ethoxcarbonyltetramates 8a-g (Scheme 2); there is up to 20% attrition of e.e. during this process, arising from cyclisation of the minor but unseparated cis-isomer 6a-g.12 These esters were expected to be suitable for conversion to the amide by direct aminolysis, a process which had been used successfully previously,16 and this was immediately successful for synthesis of systems 9a-v (see Table 1) despite the high level of functional group density around the ring system which might not have given a chemoselective process, although elevated temperature was needed for successful reaction.

Table 1:

Synthesis of tetramates 9a-v according to Scheme 2.

Compound Ar R Yield (%) Tautomeric ratioa
9a Ph 1-Adamantyl graphic file with name nihms-1031883-t0001.jpg 45 2.7:1
9b p-BrC6H4 70 2.6:1
9c m-BrC6H4 38 2:1
9d p-FC6H4 56 2.7:1
9e p-NO2C6H4 40 2.4:1
9f 2Cl-4-FC6H3 60 2.5:1
9g 2-C4H3O 92 3:1
9h Ph 4-Cyclohexylphenyl graphic file with name nihms-1031883-t0002.jpg 54 15:1
9i p-FC6H4 45 13:1
9j p-NO2C6H4 59 12.5:1
9k 2Cl-4-FC6H3 73 15:1
9l Ph 4-Chloro-2-methylphenyl graphic file with name nihms-1031883-t0003.jpg 53 only 1 form
9m p-FC6H4 40 only 1 form
9n p-NO2C6H4 37 only 1 form
9o p-BrC6H4 40 only 1 form
9p m-BrC6H4 38 only 1 form
9q 2Cl-4-FC6H3 35 only 1 form
9r p-BrC6H4 4-Morpholinophenyl graphic file with name nihms-1031883-t0004.jpg 69 only 1 form
9s 2-C4H3O 75 only 1 form
9t p-BrC6H4 4-Aminotetrahydropyranyl graphic file with name nihms-1031883-t0005.jpg 70 5.2:1
9u Ph Cyclohexyl graphic file with name nihms-1031883-t0006.jpg 49 3.1:1
9v 2Cl-4-FC6H3 40 3.1:1
a

Major form is tautomer pair AB -see Figure 4 and Table 2.

Chemoselective formation of the amide product, and not the alternative enamine arising by attack at the ketone, was evident by loss of the ester function, and also confirmed by HMBC correlation (Figure 1); thus, C-6 correlates with H-4 and H-5, while C-8 correlates with H-2 and H-5 as expected. For compounds 9s and 9t, C-9 correlates with the β-CH in the amine with respect to C-9, confirming the assignment of the carbonyl functional groups.

Figure 1.

Figure 1.

HMBC correlation data of selected carboxamide tetramates.

As anticipated, the stereochemical relationship of the starting bicyclic tetramate esters 8 was conserved in the amide systems, and a trans- relationship across the bicyclic ring between H-2 and H-5 was determined from NOE analysis (Figure 2); this outcome was further confirmed by single crystal X-ray diffraction studies of adamantyl derivative 9f (Figure 3).17

Figure 2.

Figure 2.

NOE analysis of selected carboxamide tetramates.

Figure 3:

Figure 3:

The structure of 9f from single crystal X-ray diffraction studies. Displacement ellipsoids drawn at 50% probability.

The products were generally found to exist as tautomeric forms, a phenomenon which has been observed previously, and which could be assigned from careful NMR examination.15, 18 In their NMR spectra, the internal tautomeric pairs (A and B, C and D) could not distinguished and were observed as an averaged hybrid, while the external tautomeric pairs (AB and CD) had distinct 13C chemical shift differences of C-6, C-8 and C-9 (Figure 4), so that the tautomeric forms could be readily distinguished (Table 2).

Figure 4.

Figure 4.

NOESY correlation data for 9k suggesting tautomer A as the major tautomer. The distance between H-5 and enol O-H for each tautomer was calculated for the MM2-energy minimized molecular model using Chem3D v.15.

Table 2.

13C chemical shifts of tautomers of selected examples of carboxamide tetramates.

Compound Number Solvent 13C chemical shifts (ppm) Tautomeric form
C-6 C-8 C-9
9a CDCl3 187.8 172.2 165.8 major (AB)
191.1 178.0 166.4 minor (CD)
9b CDCl3 188.4 172.4 165.9 major (AB)
191.1 178.2 166.4 minor (CD)
9f CDCl3 188.7 172.0 166.0 major (AB)
191.0 177.8 166.5 minor (CD)
MeOD 189.5 175.2 166.9 only 1 form (AB)
9k CD2Cl2 185.2 171.7 164.3 major (AB)
191.5 178.2 165.4 minor (CD)
9q CD2Cl2 184.7 171.7 164.3 only 1 form (AB)
9t CD2Cl2 186.6 172.5 165.9 major (AB)
191.6 178.8 166.5 minor (CD)

graphic file with name nihms-1031883-f0052.jpg

The 13C chemical shifts of C-6, C-8 and C-9 of the minor tautomers were invariably more downfield than their major tautomers, with the smallest Δδ13C observed for the exocyclic amide C-9. The tautomeric behaviour is also solvent dependent, as seen from the different chemical shift values of 9f in CDCl3 and CD3OD. In polar solvents (e.g. methanol-d4, acetone-d6, DMSO-d6 and CD3CN), only one set of resonances was observed for each carbon, and appears to be that of an averaged hybrid structure arising from facile equilibrium of the tautomeric forms. NOESY data obtained for 9k in dry CD2Cl2 revealed that the major tautomer for 9k was A (Figure 4), based on the correlation from H-5 to proximal enol O-H, which would not have been possible in other tautomeric forms B, C and D (Figure 4). This conclusion is supported by single crystal X-ray diffraction studies of tetramate 9f (Figure 3, vide supra), which was found to exist in the endoenolic form A. Since the chemical shift pattern observed for the major/minor tautomers is similar for all compounds, they are all assumed to adopt the major tautomeric form A, and this finding is similar to that found in related systems.15 Minimum energy conformations were generated for the four tautomeric forms A-D for compounds 9k using MM2 methodology and confirmed tautomer A to be the thermodynamically more stable (Table 3 and Table 1, SI).

Table 3.

MM2-minimized energy of tautomeric forms A-D for selected examples 9f, 9k and 9q.

Compound Number Tautomeric ratio Minimum Energy of tautomeric forms (kcal/mol)
A B C D
9f 2.5:1 18.93 39.86 26.83 39.32
9k 15:1 6.93 31.28 13.04 30.75
9q only 1 form 0.80 25.54 7.05 26.23

Having demonstrated the generality and selectivity of this approach, of interest was its application to more elaborate side chain systems; to this end, the products 9w-g’ were accessed by direct aminolysis of esters 8a, b and f using the required amines (Scheme 2 and Table 4), mostly available from commercial sources. In order to incorporate a higher level of polarity, mono-Fmoc protected 10b, prepared from dapsone 10a, an inhibitor of dihydropteroate synthetase,19 was used for the aminolysis of tetramate ester 8a to give carboxamide tetramate 9a’, which on subsequent deprotection gave 9b’ in 40 % yield (Scheme 3). The design of carboxamide 9d’ was inspired by naturally occurring tetramates kibdelomycin20 and amycolamicin,21 and was synthesised by aminolysis of tetramate 8a with glycosylated amine derivative 12b (Scheme 4). Although it has been previously reported that the desired glycosylated amine 12b could be directly synthesised from 4-aminophenol 11a and bromogalactose,22, 23 Fmoc-protected amine 11b was found to be more suitable for this purpose by Lewis acid-mediated (BF3.OEt2) glycosylation; the equatorial substitution of the aryl group was confirmed by the trans-diaxial 3JH1-H2 coupling constant (8.0 Hz). Base-mediated deprotection of the amine gave the desired product 12b which was reacted with tetramate 8a to yield carboxamide 9d’ in 60 % yield, and methanolysis under basic conditions afforded 9e’.

Table 4:

Synthesis of tetramates 9w-g’

Compound Ar R Yield (%)
9w Ph graphic file with name nihms-1031883-t0007.jpg 30
9x graphic file with name nihms-1031883-t0008.jpg 74
9y graphic file with name nihms-1031883-t0009.jpg 75
9z graphic file with name nihms-1031883-t0010.jpg 72
9a’ graphic file with name nihms-1031883-t0011.jpg 30
9b’ graphic file with name nihms-1031883-t0012.jpg 40
9c’ p-BrC6H4 graphic file with name nihms-1031883-t0013.jpg 40
9d’ Ph graphic file with name nihms-1031883-t0014.jpg 60
9e’ Ph graphic file with name nihms-1031883-t0015.jpg 54
9f’ 2Cl-4-FC6H3 graphic file with name nihms-1031883-t0016.jpg 30
9g’ 2Cl-4-FC6H3 graphic file with name nihms-1031883-t0017.jpg 36

Scheme 3.

Scheme 3

Scheme 4.

Scheme 4

The carboxamidotetramates 9a-g’, when purified by silica gel column chromatography, gave broad signal resonances in their 1H NMR spectra, consistent with metal chelation, as has been observed previously.7 As a result, carboxamides were routinely purified on silica column with 1 % Et3N in the eluent. Residual Et3N was removed by washing with 5 % citric acid, giving the purified product with sharp and well-resolved 1H spectra; an example is given for compound 9d (Figure 5). Alternatively, it was also possible to purify these carboxamides on silica column without any Et3N in the eluent, provided that they were subsequently washed with 2 M HCl to give the metal-free form.

Figure 5.

Figure 5.

1H NMR spectra of tetramate 9d (a) post-column purification, before acid-wash and (b) post-column purification, after acid-wash.

Since peak broadening was a recurrent observation after column purification, the metal-chelated tetramate and acid-washed tetramate for compound 9a were analysed by Inductively Coupled Plasma Mass Spectrometry. The metals Na, Mg, Fe and Zn but especially Ca were found in high abundance in the metal-chelated tetramate, which were substantially removed by washing with 2M HCl (Table 5 and Table 2, SI). Clearly these tetramates are easily capable of forming metal salts with diverse cations; the metal chelating behaviour of 3-acyltetramates has been reviewed recently,2, 6, 13, 14, 24 and chelation both of natural products2529 especially with antibacterial activity30 and of simpler systems3136 has also been reported while in our own work, calcium salt formation (but not other metals) during the isolation of tetramates had been observed earlier.37

Table 5.

Metal content observed in metal-chelated and acid-washed forms of carboxamide 9a.

Metal content (ppm)
Mg Ca Fe Zn Na
Metal-chelated 7026.1 20483.7 38.2 71.2 3385.6
Acid-washed 15.9 24.8 34.8 9.9 18.8

In an attempt to make use of the C-2 bromoaryl substituent for Suzuki-Miyaura (SM) cross-coupling reactions for further derivatisation, reaction of both thiazolidines 8b and 9b with 2-furanylboronic acid was examined, but only unreacted starting material was recovered; this was in contrast to oxazolidine 13a (Scheme 5) which gave the expected coupled product 13b under the same conditions. This failure of SM cross-coupling on thiazolidine-derived tetramic acids was attributed to catalyst deactivation by the cyclic thioether function, and alternative reaction conditions were clearly necessary. While successful SM cross-coupling on thiophene systems had been documented using PPh3 and AsPh3,38 no reports on systems containing non-aromatic thioethers such as those of the thiazolidine-derived tetramates could be found in the literature. However, a more recent report on Suzuki cross-coupling on biotin-derived systems39 together with a detailed review by Buchwald et al.40 suggested XPhos as an alternative, and this ligand proved to be successful, with the desired product 14a being formed along with the return of some unreacted starting material (Scheme 5 and Table 6). Although it has been reported that the active form of the catalyst for SM cross-coupling involving dialkylbiaryl phosphine ligands is the monoligated L1Pd(0) species rather than the more highly coordinated L2Pd(0),40, 41 a ratio of L:Pd=1.5:1 was not sufficient to drive the reaction to completion even after a reaction time of 48 h. Increasing the L:Pd ratio to 3:1 led to completion of the reaction and this approach was successfully applied to the synthesis of a range of carboxamide tetramates 14a-l with elaborated pendant groups at C-2 derived from 9b and 9s (Table 6). The boron reagents used for the above transformations were commercially available acids or their pinacol esters, except for 14c where the boron reagent was the glycol ester derived from the corresponding boronic acid, prepared according to literature procedure.42 Carboxamide tetramate 14k was synthesised as a mixture of diastereomers from the corresponding commercially available racemic 1-(tetrahydropyran-2-yl)-1H-pyrazole-5-boronic acid pinacol ester. It would appear the additional nucleophilicity of XPhos gives a higher electron density at Pd(0) in the monoligated L1Pd(0) species, and increases the rate of oxidative addition leading to overall successful coupling.40

Scheme 5.

Scheme 5

Table 6:

Suzuki-Miyaura cross-coupling of carboxamide tetramates 9b and 9t.

Carboxamide tetramate (R=) Compound Number R1 % Yield Compound Number R1 % Yield
graphic file with name nihms-1031883-t0018.jpg 14a graphic file with name nihms-1031883-t0019.jpg 72 14g graphic file with name nihms-1031883-t0020.jpg 70
14b graphic file with name nihms-1031883-t0021.jpg 74 14h graphic file with name nihms-1031883-t0022.jpg 78
14c graphic file with name nihms-1031883-t0023.jpg 80 14i graphic file with name nihms-1031883-t0024.jpg 53
14d graphic file with name nihms-1031883-t0025.jpg 45 14j graphic file with name nihms-1031883-t0026.jpg 70
14e graphic file with name nihms-1031883-t0027.jpg 78 14k graphic file with name nihms-1031883-t0028.jpg 60
14f graphic file with name nihms-1031883-t0029.jpg 74
graphic file with name nihms-1031883-t0030.jpg 14l graphic file with name nihms-1031883-t0031.jpg 30

Several naturally occurring tetramic acids bearing glycosyl residues are known, including virgineone,43 kibdelomycin,44 amycolamicin,45 aurantoside46, 47 and streptolydigin.48 While the exact role of the sugar moiety of these tetramates is not known, it may give improved bioavailability with increased solubility and transportation, or help modulate toxicity and aid in bacterial target interaction.49 Thus, investigating the effect of glycosylated tetramates on antibacterial activity was of interest, and the systems developed thus far offered the feasibility of C-2 substitution of the tetramate ring. In order to demonstrate an alternative process using late-stage ring formation, bicyclic tetramates with C-2 pendant glycones were derived from β-D-galactose pentaacetate (Scheme 6). Acetobromo-α-D-galactose 15b was prepared from β-D-galactose pentaacetate 15a according to the literature procedure;50 axial bromination at the anomeric carbon was confirmed from the 3J(H1-H2)-coupling constant of H-1 (d, J=4.1 Hz). Solid-liquid phase transfer-catalysed aryl glycosylation of 4-hydroxybenzaldehyde with acetobromo-α-D-galactose 15b gave the desired aldehyde 15c, according to a modified literature procedure; BnNBu3Cl proved to be a better phase transfer catalyst than Bu4NBr while the presence of H2O led to a remarkable reduction in yield.51 The β-anomer was obtained exclusively, where the equatorial substitution of the phenoxide at the anomeric carbon was confirmed by 3J-coupling constant of the anomeric C-H (d, J=8.0 Hz). The glycosylated aldehyde 15c was then condensed with ʟ-cysteine methyl ester hydrochloride to yield the expected thiazolidine in 85 % yield with a cis:trans diastereomeric ratio of 1.6:1, and this material was N-acylated to give the corresponding malonamide in 52 % yield, which was in turn cyclised to afford the glycosylated tetramate 16 but only if the crude product was isolated without acidic work-up. Base-catalysed methanolysis gave tetraol 17 in quantitative yield (Scheme 6); neither hydrolysis nor transesterification of the ethyl ester at C-7 was observed under these reaction conditions. Aminolysis of tetramate ester 16 afforded carboxamide tetramates 18 with the acyl groups of the sugar moiety intact; methanolysis of the former resulted in the very polar carboxamide tetramate 19.

Scheme 6.

Scheme 6

A tetramate with a disaccharide pendant group was derived from commercially available lactose octaacetate (Scheme 7); this followed a similar approach to the monosaccharide derivative above. Commercially available lactose octaacetate 20a (as a 2.9:1 mixture of α- and β- anomers at the glucopyranosyl end) was subjected to BiBr3-catalysed bromination and afforded 20b as the α-anomer in 92 % yield, confirmed by 3J-coupling constant of the anomeric C-H. Aryl glycosylation gave aldehyde 20c, which was condensed with ʟ-cysteine methyl ester hydrochloride, acylated and cyclised to afford tetramate 21.

Scheme 7.

Scheme 7

Due to the acid-labile nature of phenyl glycosides prepared above and the resulting difficulty associated with the acidic work-up procedure required to remove chelating metal ions, an alternative route to the incorporation of glycones via a benzyl alcohol linking group was explored (Scheme 8). The controlled reduction of terephthalaldehyde with NaBH4 between 0–2°C afforded 4-(hydroxymethyl)benzaldehyde in 70 % yield according to a modified literature procedure.52 Lewis acid-catalysed glycosylation with β-D-galactosepentaacetate 15a gave aldehyde 22 as the β-anomer in 71% yield along with the byproduct 4-(acetoxymethyl)benzaldehyde in 28 % yield. Following the usual synthetic route to access the bicyclic tetramate core, condensation of aldehyde 22 with cysteine ethyl ester yielded the expected thiazolidine, which was acylated and cyclized under basic conditions to afford benzyl glycosylated tetramate 23, which proved to be stable to acidic work-up.

Scheme 8.

Scheme 8

In order to permit late stage C-2 substituent manipulation, tetramates 26a and 26b were derived from tetramates 25a,b by silyl deprotection, which were obtained from the silyl ether of 4-hydroxybenzaldehyde 24a or 4-(hydroxymethyl)benzaldehyde 24b respectively (Scheme 9). Conversion of 25a to the corresponding carboxamides 27a-c by aminolysis with the required amine proceeded using the approach described above, in modest to good yield. The deprotection of silyl ether 27a to obtain tetramate 28a, however, required some optimisation. While TBAF was a successful desilylating agent, chromatographic purification was difficult, possibly due to the basicity of fluoride ion in an aprotic solvent53 leading to deprotonation of the product tetramic acid. However, alternative conditions using KF as a fluoride source were more successful. Two other carboxamide tetramates 28b and 28c were also synthesised using this approach.

Scheme 9.

Scheme 9

The presence of a phenolic group at the C-2 position provided a platform for further functionalization of carboxamide tetramates via Williamson ether synthesis. In the presence of K2CO3, nucleophilic substitution of ethyl bromoacetate and 1-bromo-2-butyne by the phenoxide derived from 28a gave carboxamide tetramates 29a and 30 respectively (Scheme 9). HMBC correlation data confirmed the substitution at the phenolic but not at the enolic alcohol group. Ester hydrolysis of 29a afforded the carboxylic acid bearing tetramate 29b with improved hydrophilicity.

Since the pyroglutamate skeleton is found in antibacterial natural products (e.g. oxazolomycin54, 55 and pramanicin56), the synthesis of a pyroglutamate-tetramate hybrid system was also examined. This was achieved by condensing pyroglutaminol 31,57 prepared by the literature procedure,58, 59 with terephthalaldehyde by reflux in toluene using a Dean-Stark apparatus to give a mixture of products 32 and 33 in 1:2.4 ratio (Scheme 10); the lactam-functionalized aldehyde 33 was isolated by column chromatography as a single diastereomer, whose stereochemical assignment was in agreement with previous reports on N,O-acetal formation of 31 with benzaldehyde.60, 61 Aldehyde 33 was in turn converted to the desired bicyclic tetramate ester 34a in the usual way, which was converted to carboxamide 34b with 1-adamantylamine. The stereochemical assignment of the tetramate system was confirmed by NOE experiments (Scheme 10). The relative stereochemistry of both heterocyclic systems was found to be trans-, in keeping with structures reported in the literature.62

Scheme 10.

Scheme 10

Of interest is that direct Buchwald-Hartwig amination of bromo-substituted tetramates 8b and 9b using any of XPhos, RuPhos and XantPhos was not successful. Instead, p-bromobenzaldehyde 35a was functionalized via Buchwald amination/amidation to give aldehydes 35b,c and these were then used to synthesise the tetramate core de novo using a similar route to that described above, giving esters 36a,b which were converted to carboxamides 37a,b as before (Scheme 11).

Scheme 11.

Scheme 11

In one case, deprotection of the N,S-acetal was examined; in comparison to oxazolidine-based N,O-acetal systems, thiazolidine-based N,S-acetals have been shown to have higher stability to acids,63 and their deprotection often requires harsh conditions such as heating under reflux with 5M HCl for 72 h.64 However, Onoda et al. have shown that Boc-protected thiazolidines can be deprotected at the N,S-acetal in good yield with 15 eq of TFA in water-saturated CH2Cl2 under ambient conditions within 2 h.65 Deprotection of carboxamide tetramate 9a was investigated with TFA in CH2Cl2, but even after a reaction time of 24 h, only starting material remained, and while the addition of water accelerated the reaction, complete deprotection was not observed in 24 h. However, deprotection of thiazolidine via the Corey-Reichard protocol66 was successful giving carboxamidotetramate 38 in 70 % yield (Scheme 12). This observation highlights the need for stronger conditions for the deprotection of the more stable thiazolidine-derived carboxamide tetramates.

Scheme 12.

Scheme 12

Antibacterial properties

Mode of action of tetramates

Naturally-occurring tetramates, such as kibdelomycin and amycolamycin, are inhibitors of bacterial topoisomerase IV (topo IV) and DNA gyrase,20, 44, 67 whereas the tetramates streptolydigin and tirandamycin are inhibitors of RNA polymerase (RNAP).68, 69 Some tetramates have been shown to provide dual inhibition of both RNAP and undecaprenyl pyrophosphate synthase (UPPS).7 All four of theses enzymes are essential for viability. The primary function of topo IV is the decatenation of daughter chromosomes that are multiply-linked, during the final stages of DNA replication,70 and the role of DNA gyrase is catalysis of negative supercoiling of DNA.71 Bacterial RNAP is an attractive target in antibacterial chemotherapy as the bacterial RNAP subunit sequences are highly conserved while being different from eukaryotic RNAP subunit sequences. This permits broad-spectrum activity and therapeutic selectivity.72 An examination of the mode of action of the tetramates prepared in the current work was therefore of interest.

Inhibition of possible bacterial target enzymes by some bicyclic tetramates was studied for Staphylococcus aureus topo IV and RNAP, Escherichia coli RNAP and Mycobacterium tuberculosis gyrase. A selection of compounds (full data is given in Tables 3–5, SI) initially was tested at a fixed concentration of 100 μM for the inhibition of topo IV and gyrase, and the percentage of DNA decatenated or supercoiled in the presence of each test compound was calculated. A lower percentage of decatenation or supercoiling indicated a higher level of inhibition of the target enzymes. The data were compared to those for the known topo IV and gyrase inhibitor, ciproflaxacin. Concentrations resulting in half-maximal inhibition (IC50s) for some of the more potent inhibitors (vide infra) were determined and are presented in Figure 6(a). Tetramates 9a, 9h and 9m were tested for their inhibition of S. aureus and E. coli RNAP,7375 and the data obtained are presented in Figure 6(b).

Figure 6.

Figure 6.

(a) IC50 values of tetramates against M. tuberculosis gyrase and S. aureus topoisomerase IV. Compounds 9a and 9m were weak inhibitors of M. tuberculosis gyrase and 9d’ was a weak inhibitor of S. aureus topoisomerase IV. (b) IC50 values of tetramates against S. aureus RNAP and E. coli RNAP.

Comparison of mimimal inhibitory concentrations (MICs) for antibacterial activity against methicillin-resistant S. aureus (Table 10, SI; see also Figure 7; vide infra) with IC50s for inhibition of S. aureus RNAP and topo IV (Figure 6) indicates that compounds exhibiting high antibacterial activities also exhibit high S. aureus RNAP and topo IV inhibitory activities, consistent with the possibility that antibacterial activity may be attributable to RNAP and topo IV inhibitory activity. Compounds 9a, 9h and 9m inhibited both RNAP and S. aureus topo IV, consistent with a possible dual mode of action for these compounds. Some compounds that showed high efficacy in enzyme inhibition assays did not show high efficacy in whole-cell assays (e.g 9d’, 19). This potentially may be due to low cell permeability or elimination via efflux mechanisms, leading to reduced bioavailability in the bacterial cell. In general, E. coli RNAP was less sensitive to tetramates than S. aureus RNAP, and the M. tuberculosis gyrase was less sensitive to tetramates than S. aureus topo IV.

Figure 7.

Figure 7.

Bioactivity of carboxamide tetramates with/without HSA.

Antibacterial activity: hole-plate method

Whole-cell antibacterial assays were performed using the hole-plate method with Gram-positive S. aureus DS267 or Gram-negative E. coli X580, using Cephalosporin C as a positive control. The samples were prepared as 4 mg/mL solutions of 70% DMSO in MeOH, with serial dilution to the desired concentrations where necessary. The relative potency was estimated by reference to positive controls prepared with Cephalosporin C.76 Many of the carboxamide analogues 9a-v showed antibacterial activity against S. aureus, but none showed activity against E. coli (Table 6–9, ESI). The presence of an adamantyl carboxamide side chain on the tetramate core resulted in a significant enhancement of antibacterial activity and such tetramates 9a-f were active at a concentration of 1 μg/mL. Tetramates 9h-9k with a 4-cyclohexylphenyl group or 9l-9q with a 4-chloro-2-methylphenyl group demonstrated good antibacterial activity, although much less potent compared to their adamantyl analogues, with the 4-cyclohexylphenyl group being the more active. The decrease in antibacterial activity observed for 9r, 9s and 9w further confirms the requirement for a bulky pendant group at C-7 of the bicyclic tetramate core.

Since some tetramates have been shown to lose antibacterial activity in the presence of human serum albumin (HSA),7 further assays were run in the presence of HSA. These data were compared with the bioactivity of each sample in the absence of HSA and selected examples are given in Figure 7 along with full data in Table 6–9 (ESI). While 9a-9f showed good antibacterial activity at 1 μg/mL in the absence of HSA, there was a complete loss of activity in the presence of 4 mg/mL of HSA (where the tetramate:HSA ratio is approximately 1:30, data not shown). These data suggest that carboxamides are possible ligands for HSA. Nevertheless, of interest was the return of bioactivity for these tetramates when the concentration of the test compounds was increased to 5 μg/mL (in which tetratmate:HSA ratio is approximately 1:6). Similarly, tetramates 9h-9q exhibited no antibacterial activity at their original concentrations when tested in the presence of HSA, but their activity was regained by increasing the compound concentrations. Interestingly, tetramates 9g,9r,9s and 9w which were less potent compared to the other tetramates discussed above retained their potency in the presence of HSA.

The assay samples were prepared in 70% DMSO in MeOH for reasons of solubility, followed by mixing with albumin solution, and resulted in a final solvent composition of 35:15:50 of DMSO:MeOH:H2O in the well (35% DMSO). It is known that higher concentrations of DMSO can lead to protein denaturation,77, 78 and this might have interfered with the above analysis. Carboxamides 9a-9f showed good antibacterial activity at 1 μg/mL, and these samples could be dissolved in a minimum amount of DMSO at such low concentrations. Thus, the antibacterial activity of carboxamides 9a-9f in the presence of HSA was tested where the final solvent composition was 2% DMSO in H2O. The zones of inhibition measured were in agreement with the data obtained in the presence of 35% DMSO, confirming that the antibacterial activity observed for the tetramates in the presence of HSA was not a result of the compound unbound to denatured HSA.

Broth dilution method

Broth assays were performed to determine MICs of compounds (Table 10, ESI). The carboxamido tetramates 9a-g’ showed either weak (MIC= 250 μg/mL) or no activity against Gram-(−) strains, in agreement with the data obtained from the hole-plate method, while some exhibited promising antibacterial activity against Gram-(+) strains. Among the C-7 functionalized tetramates, the most potent carboxamide tetramates 9b-f, h, l-m, q, u-v, z showed an MIC of 0.49 μg/mL against at least one organism, significantly better than the tetramate ester analogues 8a-g,12 highlighting the importance of a carboxamide side chain group for whole-cell antibacterial activity. However, the presence of more polar carboxamides at C-7 led to a loss of activity (see 9d’ and 9e’). Since an MIC value of 16–32 μg/mL against a test organism is considered a suitable minimum for optimisation,79 several tetramates therefore offer suitable starting points for core scaffold elaboration. The selectivity of these tetramates for prokaryotic cells over mammalian cells was analysed via a primary cytotoxicity screening against HaCaT cell line (see ESI). Compounds that are worth progressing from this stage should possess ≥10-fold higher antibacterial activity over cytotoxicity,79 and on this basis, carboxamido tetramates 9e, 9f, 9g, 9h, 9l, 9m and 9p are of interest.

The tetramates with C-2 functionalization obtained from SM cross-coupling were assayed for antibacterial activity against Gram-(−) bacterial strains E. coli, K. pneumoniae and P. aeruginosa and Gram-(+) strains MRSA, Enterococcus and Streptococcus pneumonia but no activity was observed. There was, however, potent antibacterial activity against Gram-(+) strains MRSA, Enterococcus and Streptococcus pneumoniae. In particular, compounds 14d, 14h and 14k displayed improved antibacterial activity against MRSA while 14a, 14d and 14h showed improved activity against Enterococcus sp. compared to 9a. The glycosylated tetramates were either inactive or only weakly active (MIC= 250 μg/mL) against Gram-(−) strains tested, but displayed potent antibacterial activity against Gram-(+) strains MRSA, Enterococcus and Streptococcus pneumoniae. Contrary to the trend in antibacterial activity observed for the tetramate esters 8a-g,12 tetramate ester 16 proved to be more active than its adamantyl analogue 18. Increasing polarity at C-2 as in 17 conferred no bioactivity. Introduction of a disaccharide analogue also led to the complete loss of antibacterial activity of tetramate ester 21. Further, it was interesting to note that the aglycone analogue 26a was slightly more potent than the glycosylated tetramate 16. In the case of glycosyl tetramate 23, in which an acylated galactose is linked to the tetramate via a benzyl ether, improved antibacterial activity was observed compared to 16. Both 23 and its aglycone analogue 26b displayed similar level of antibacterial activity. Thus, while the presence of a glycone does not lead to a significant compromise in bioactivity, it seems that a significant increase in polarity at C-2 leads to a detrimental reduction in the antibacterial activity (e.g. 17). The exact role of the glycosidic residue is not known, but it may assist in improving polarity, bioavailability, modulating toxicity and bacterial target interaction.49 Conversion of the tetramate ester 16 to its carboxamide analogues 18 led to a loss in activity. However, the absence of the glycone moiety in these carboxamide tetramates restored antibacterial activity as observed for 28a, indicating that the presence of a glycone is not favourable for antibacterial activity of carboxamide tetramates. This outcome for amides is in contrast to that observed for tetramate esters.

Etherification of 28a to 29a resulted in a 64-fold reduction in antibacterial activity. Ethyl ester hydrolysis of 29a to give carboxylic acid 29b resulted in a further 8-fold reduction in potency (MIC of 29b against MRSA was 250 μg/mL). However, etherification of 28a with 1-bromo-2-butyne afforded 30 without much compromise in antibacterial activity. It was also interesting to note that replacement of the adamantyl side chain with either phenyl (28b) or tetrahydropyranyl (28c) groups led to a reduction in potency, with no antibacterial activity being observed for the latter. An evaluation of adamantyl carboxamide tetramates with aliphatic heterocycles 34b, 37a and 37b appended at C-2 showed good antibacterial activity when compared to the corresponding tetramate esters. This observation is in agreement with the general trend observed for tetramate esters and their carboxamide analogues, highlighting the need for a bulky, hydrophobic C-7 pendant group. In a separate comparison of C-7 adamantyl carboxamides bearing morpholine, pyroglutamate and β-lactam moieties at C-2, morpholine substituted tetramate 37a was the most active while pyroglutamate substituted tetramate 34b was the least potent.

In the absence of a C-2 substituent, antibacterial activity of tetramate 38 against Gram-positive S. aureus is significantly reduced, and to obtain a similar zone of inhibition to that of 9a, a 1000-fold increase in concentration of 38 was necessary. However, upon deprotection, activity against Gram-negative E. coli could be observed; thus, at a concentration of 4 mg/mL, 38 showed activity, while 9a was completely inactive. It is noteworthy that deprotected tetramate 38 does not exhibit significant binding to albumin as its antibacterial activity is retained when treated with HSA. This improvement in activity observed for E. coli could be due at least in part to its increased polarity, which in turn leads to increased cell membrane permeability.

O’Shea and Moser have analysed the major antibacterial classes of compounds and have shown that mean values for physicochemical properties define distinct classes;80 antibacterial efficacy does not, however, necessarily correspond to oral bioavailability codified in the Lipinski Rule of 5,81 primarily as a result of bacterial cell wall penetration.82, 83 A correlation of calculated physicochemical properties with the antibacterial activity of some known tetramates was undertaken (using Marvin (16.4.18.0), 2016, ChemAxon (http://www.chemaxon.com)) (Table 11, SI), and their physicochemical properties show general resemblance to fluoroquinolones, with relative PSA values of 13–25% and a MW range of 300–600Da. A similar comparison of physicochemical properties (MW, clogD7.4 and rel. PSA) and antibacterial activity (where ‘activity’ was defined as MIC of ≤ 8 μg/mL80 against at least one bacterial strain) of the tetramate compound library (Figure 8 and Table 11, SI) showed compounds that are active largely occupy physicochemical space with MW of 300–600, clogD7.4 of −2.5 to 4 and rel. PSA of 11–22; in this, they also generally resemble fluoroquinolones.80 The ‘active’ tetramate esters 16 and 23 with glycone pendants appear as outliers due to their higher MW in the region of 650–670. Compounds that are more hydrophilic (that is, with lower clogD7.4 and higher relative PSA) did not exhibit antibacterial activity against Gram-positive strains, even though the mean MW of known Gram-positive antibacterial agents is around 800.80 However, not all polar tetramates that fit in to the optimal chemical space of antibacterials with Gram-negative activity showed the expected activity, although the antibacterial activity of N,S-acetal deprotected tetramate 38 against E. coli suggests suitable polarity improves Gram-negative activity of tetramates, since uptake/penetration through the cell membrane might be better.70 Tetramates with antibacterial activity possessed desired physicochemical properties required by Lipinsky’s Ro5, and therefore have potential to be developed as oral drug candidates, but these highly hydropbobic compounds are likely to be penetrative of the blood brain barrier on the basis of calculation of log BB values using a published algorithm.84 Furthermore, the Fsp3 value (Fsp3 = sp3 hybridized carbon count/ total carbon count85) of the tetramate scaffold 8 is 0.43; Lovering and co-workers have demonstrated that complexity as measured by Fsp3 increased for molecules in clinical progression and that drugs have an average Fsp3 of 0.47.85 This places the tetramate scaffold in an ideal region of chemical space for development. Moreover, increased complexity appears to reduce toxicity of drug candidates,86 and this was consistent with our observations: for the antibacterially active tetramates for which cytotoxicity against HaCaT mammalian cell line was determined, the mean Fsp3 calculated for tetramates with low toxicity (Fsp3 = 0.41), defined as those that possess ≥10-fold antibacterial activity (against at least one bacterial strain) over cytotoxicity, was higher than those with higher toxicity levels (Fsp3 = 0.38).86 No reaction of tetramate 9f with benzylamine and cysteine under ambient conditions was observed, consistent with the lack of toxicity observed in these systems.

Figure 8.

Figure 8.

Correlation of physicochemical properties (MW, clogD7.4, and rel. PSA) with antibacterial activity.

Experimental

General procedure: Esterification of L-serine and DL-cysteine.12

To a suspension of the amino acid (1.0 eq) in MeOH (2 mL/mmol) at 0 °C , SOCl2 (1.2 eq) was added drop-wise under continuous stirring and warmed to rt, then heated at reflux for 1–3 h. The reaction mixture was concentrated in vacuo to obtain the respective amino ester.

General procedure: Synthesis of N-acylated thiazolidines 6,7.12

Step 1: To ʟ-cysteine methyl ester hydrochloride (1.0 eq) in petrol (25 mL/1 g), Et3N (1.2 eq) and aldehyde (1.2 eq) were added. The mixture was heated at reflux, with continuous removal of water using a Dean-Stark apparatus, for 19 h. It was then filtered and washed with Et2O. The combined filtrates were concentrated in vacuo and residue was purified by silica gel flash column chromatography (eluent: EtOAc/petrol) to give the required thiazolidine.

Step 2: A solution of ethyl hydrogen malonate (1.2 eq) in CH2Cl2 (2.5 mL/mmol) was added to a stirred solution of thiazolidine (1.0 eq), DCC (1.2 eq) and DMAP (0.1 eq) in CH2Cl2 (5 mL/mmol) at 0°C. The mixture was stirred at 0 °C for 15 min and then at rt for 15 h. The reaction mixture was filtered to remove dicyclohexylurea and the residue was washed with CH2Cl2. The combined filtrates were concentrated in vacuo and purified by silica gel flash column chromatography (eluent: EtOAc/petrol) to give N-acylated thiazolidines 6,7.

General procedure: Synthesis of bicyclic tetramates 8.12

KOtBu (1.2 eq) was added to a solution of the N-acylated thiazolidine in THF and heated at reflux for 3 h. It was then cooled to rt, concentrated in vacuo and partitioned between Et2O and water. The aqueous layer was extracted and acidified with 2M HCl (to pH 1) and extracted with EtOAc. The combined EtOAc extracts were washed with brine, dried over Na2SO4, filtered and concentrated in vacuo. The residue was purified by flash column chromatography (with 1% Et3N) to give the desired product. The product was then dissolved in CH2Cl2 and washed with 5% citric acid. The organic fractions were dried over Na2SO4, filtered and concentrated in vacuo to yield the desired bicyclic tetramates 8a-g.

General procedure: Synthesis of carboxamide tetramates 9a-g’

To tetramic acid (1.0 eq) dissolved in THF/toluene or DMSO/toluene (1:9, 10 mL/mmol), amine (1.5 eq) was added. The solution was heated at reflux for 16 h, cooled to rt and concentrated in vacuo. The residue was purified by silica gel flash column chromatography (eluent: EtOAc/MeOH/1 % Et3N). The product was dissolved in CH2Cl2 and washed with 5% citric acid. The organic fractions were dried over Na2SO4, filtered and concentrated in vacuo to yield the bicyclic carboxamide tetramate.

Conclusion

Thiazolidine derived tetramate esters were readily elaborated to carboxamido tetramates by direct uncatalyzed ester to amide exchange, and conditions were identified which permit skeletal elaboration by Suzuki-Miyaura coupling despite the presence of a cyclic thioether function. While tetramate esters were inactive against both Gram-positive and Gram-negative bacterial strains, the presence of a carboxamide pendant group at C-7 of the tetramate core conferred antibacterial activity against Gram-positive strains; however, the introduction of more polar substituents led to reduced or no activity as observed for 17 and 29b. The antibacterial activity spectrum observed for various C-2 functionalized tetramates demonstrated the ability to accommodate a range of substituents while retaining antibacterial activity, providing scope for physicochemical property optimization, and an analysis of physicochemical properties indicates that bioactive tetramates generally occupy physicochemical space with MW of 300–600, clogD7.4 of −2.5 to 4 and rel. PSA of 11–22%. With regards to cytotoxicity, compounds with low toxicity had a higher Fsp3 value compared to those with high toxicity. Evaluation of antibacterial activity of tetramates in the presence of HSA suggests possible binding of tetramates to albumin may occur. Although the major tautomeric pair of these species has been identified in solution, it should not be assumed that this is also the binding tautomer at its enzyme target site, since a number of tautomeric forms may be in free equilibrium.

More generally, the identification of novel, non-planar, synthetically accessible heterocyclic systems has become of interest in drug discovery.85, 87, 88 The work described here shows that highly functionalised heterocyclic skeletons with several points of diversity, modelled on bioactive natural products, are available in a short synthetic sequence, and that further functional group elaboration may give structures which exhibit similar bioactivity profiles to that of the parent natural product, in this case antibacterial activity. Importantly, these compounds provide well-defined 3D templates which minimise aromatic ring count,88 but maintain acceptable starting values of MW, PSA, numbers of rotatable bonds, H-bond acceptors and H-bond donors, while leaving ample scope for lead optimisation in the drug discovery process. They therefore offer suitable skeletons for application in fragment-based drug design89, 90 which allow escape “from flatland”.85

Supplementary Material

sup1
sup2

Acknowledgements

This work was supported by MRC Confidence in Concept grant to M.G.M. and National Institutes of Health Grants GM041376 and AI109713–1 to R.H.E.

Footnotes

Conflicts of interest

There are no conflicts of interest to declare.

References

  • 1.Stefanucci A, Novellino E, Costante R and Mollica A, Heterocycles, 2014, 89, 1801–1825. [Google Scholar]
  • 2.Schobert R, Naturwissenschaften, 2007, 94, 1–11. [DOI] [PubMed] [Google Scholar]
  • 3.Royles BJL, Chem. Rev, 1995, 95, 1981–2001. [Google Scholar]
  • 4.Henning HG and Gelbin A, in Adv. Heterocycl. Chem, ed. Katritzky AR, 1993, 57, 139–185. [Google Scholar]
  • 5.Mo XH, Li QL and Ju JH, RSC Advances, 2014, 4, 50566–50593. [Google Scholar]
  • 6.Schobert R and Schlenk A, Bioorg. Med. Chem, 2008, 16, 4203–4221. [DOI] [PubMed] [Google Scholar]
  • 7.Jeong Y-C, Anwar M, Moloney MG, Bikadi Z and Hazai E, Chem. Sci, 2013, 4, 1008–1015. [Google Scholar]
  • 8.Anwar M and Moloney MG, Tetrahedron Lett., 2007, 48, 7259–7262. [Google Scholar]
  • 9.Josa-Cullere L, Moloney MG and Thompson AL, Synlett, 2016, 27, 1677–1681. [Google Scholar]
  • 10.Matiadis D, Catalysts, 2019, 9. [Google Scholar]
  • 11.Jeong Y-C, Anwar M, Nguyen TM, Tan BSW, Chai CLL and Moloney MG, Org. Biomol. Chem, 2011, 9, 6663–6669. [DOI] [PubMed] [Google Scholar]
  • 12.Panduwawala TD, Iqbal S, Tirfoin R and Moloney MG, Org. Biomol. Chem, 2016, 14, 4464–4478. [DOI] [PubMed] [Google Scholar]
  • 13.Petermichl M and Schobert R, Synlett, 2017, 28, 654–663. [Google Scholar]
  • 14.Zaghouani M and Nay B, Natural Product Reports, 2016, 33, 540–548. [DOI] [PubMed] [Google Scholar]
  • 15.Jeong Y-C and Moloney MG, J. Org. Chem, 2011, 76, 1342–1354. [DOI] [PubMed] [Google Scholar]
  • 16.Jeong Y-C and Moloney MG, Beilstein J. Org. Chem, 2013, 9, 1899–1906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Low temperature,91 single crystal X-ray diffraction data were collected on a representative crystal of 9f using a (Rigaku) Oxford Diffraction Supernova A diffractometer (λ = 1.54180 Å). Raw frame data were reduced using CrysAlisPro and solved using Superflip92 prior to refinement with CRYSTALS.93 Full details can be found in the ESI (CIF); Crystallographic data have also been deposited with the Cambridge Crystallographic Data Centre (CCDC 1904326) and copies of these data can be obtained free of charge via. http://www.ccdc.cam.ac.uk/data_request/cif.
  • 18.Nolte MJ, Steyn PS and Wessels PL, J. Chem. Soc., Perkin Trans 1, 1980, 1057–1065. [Google Scholar]
  • 19.Coleman MD, Br. J. Dermatol, 1993, 129, 507–513. [DOI] [PubMed] [Google Scholar]
  • 20.Phillips J, Goetz M, Smith S, Zink D, Polishook J, Onishi R, Wiltsie SS,J, Allocco J, Sigmund J, Dorso K, Lee S, Skwish S, de la Cruz M, Martín J, Vicente F, Genilloud O, Lu J, Painter R, Young K, Overbye K, Donald R and Singh S, Chem Biol., 2011, 18, 955–965. [DOI] [PubMed] [Google Scholar]
  • 21.Sawa R, Takahashi Y, Hashizume H, Sasaki K, Ishizaki Y, Umekita M, Hatano M, Abe H, Watanabe T, Kinoshita N, Homma Y, Hayashi C, Inoue K, Ohba S, Masuda T, Arakawa M, Kobayashi Y, Hamada M, Igarashi M, Adachi H, Nishimura Y and Akamatsu Y, Chem. - Eur. J, 2012, 18, 15772–15781. [DOI] [PubMed] [Google Scholar]
  • 22.Spicer CD and Davis BG, Chem. Commun , 2013, 49, 2747–2749. [DOI] [PubMed] [Google Scholar]
  • 23.Smits E, Engberts JBFN, Kellogg RM and van Doren HA, J. Chem. Soc., Perkin Trans 1 1996, 2873–2877. [Google Scholar]
  • 24.Athanasellis G, Igglessi-Markopoulou O and Markopoulos J, Bioinorg. Chem. Appl, 2010, 315056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Cherian PT, Deshpande A, Cheramie MN, Bruhn DF, Hurdle JG and Lee RE, J. Antibiot, 2017, 70, 65–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Shang Z, Li L, Espósito BP, Salim AA, Khalil ZG, Quezada M, Bernhardt PV and Capon RJ, Org. Biomol. Chem, 2015, 13, 7795–7802. [DOI] [PubMed] [Google Scholar]
  • 27.Biersack B, Diestel R, Jagusch C, Sasse F and Schobert R, J. Inorg. Biochem, 2009, 103, 72–76. [DOI] [PubMed] [Google Scholar]
  • 28.Chang P-K, Ehrlich KC and Fujii I, Toxins, 2009, 1, 74–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Steyn PS and Rabie CJ, Phytochemistry, 1976, 15, 1977–1979. [Google Scholar]
  • 30.Dandawate P, Padhye S, Schobert R and Biersack B, Expert Opin. Drug Discovery, 2019, 14, 563–576. [DOI] [PubMed] [Google Scholar]
  • 31.Gavrielatos E, Athanasellis G, Heaton BT, Steiner A, Bickley JF, Igglessi-Markopoulou O and Markopoulos J, Inorg. Chim. Acta, 2003, 351, 21–26. [Google Scholar]
  • 32.Gavrielatos E, Mitsos C, Athanasellis G, Heaton BT, Steiner A, Bickley JF, Igglessi-Markopoulou O and Markopoulos J, J. Chem. Soc., Dalton Trans, 2001, 639–644. [Google Scholar]
  • 33.Heaton BT, Jacob C, Markopoulos J, Markopoulou O, Nahring J, Skylaris C-K and Smith AK, J. Chem. Soc., Dalton Trans, 1996, 1701–1706. [Google Scholar]
  • 34.Tietze O, Reck G, Schulz B and Zschunke A, J. Prakt. Chem./Chem.-Ztg, 1996, 338, 642–646. [Google Scholar]
  • 35.Barkley JV, Markopoulos J and Markopoulou O, J. Chem. Soc., Perkin Trans 2, 1994, 1271–1275. [Google Scholar]
  • 36.Lebrun M-H, Duvert P, Gaudemer F, Gaudemer A, Deballon C and Boucly P, J. Inorg. Biochem, 1985, 24, 167–181. [DOI] [PubMed] [Google Scholar]
  • 37.Josa-Culleré L, Towers C, Willenbrock F, Thompson AL and Moloney MG, Eur. J. Org. Chem, 2018, 7055–7059. [Google Scholar]
  • 38.Suzuki A, Angew. Chem. Int. Ed. Engl, 2011, 50, 6722–6737. [DOI] [PubMed] [Google Scholar]
  • 39.Fairhead M, Shen D, Chan LKM, Lowe ED, Donohoe TJ and Howarth M, Bioorg. Med. Chem, 2014, 22, 5476–5486. [DOI] [PubMed] [Google Scholar]
  • 40.Martin R and Buchwald SL, Acc. Chem. Res, 2008, 41, 1461–1473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Swamy KCK, Kumar NNB, Balaraman E and Kumar KVPP, Chem. Rev, 2009, 109, 2551–2651. [DOI] [PubMed] [Google Scholar]
  • 42.Iwai Y, Gligorich KM and Sigman MS, Angew. Chem. Int. Ed. Engl, 2008, 47, 3219–3222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Figueroa M, Raja H, Falkinham JO, Adcock AF, Kroll DJ, Wani MC, Pearce CJ and Oberlies NH, J. Nat. Prod , 2013, 76, 1007–1015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Singh SB, Bioorg. Med. Chem, 2016, 24, 6291–6297. [DOI] [PubMed] [Google Scholar]
  • 45.Sawa R, Takahashi Y, Hashizume H, Sasaki K, Ishizaki Y, Umekita M, Hatano M, Abe H, Watanabe T, Kinoshita N, Homma Y, Hayashi C, Inoue K, Ohba S, Masuda T, Arakawa M, Kobayashi Y, Hamada M, Igarashi M, Adachi H, Nishimura Y and Akamatsu Y, Chem. - Eur. J, 2012, 18, 15772–15781. [DOI] [PubMed] [Google Scholar]
  • 46.Reinscheid F and Reinscheid UM, J. Mol. Struct, 2017, 1147, 96–102. [Google Scholar]
  • 47.Petermichl M, Loscher S and Schobert R, Angew. Chem. Int. Ed, 2016, 55, 10122–10125. [DOI] [PubMed] [Google Scholar]
  • 48.Pronin SV, Martinez A, Kuznedelov K, Severinov K, Shuman HA and Kozmin SA, J. Am. Chem. Soc, 2011, 133, 12172–12184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Křen V and Řezanka T, FEMS Microbiol Rev, 2008, 32, 858–889. [DOI] [PubMed] [Google Scholar]
  • 50.Montero J-L, Winum J-Y, Leydet A, Kamal M, Pavia AA and Roque J-P, Carbohydr. Res, 1997, 297, 175–180. [Google Scholar]
  • 51.Hongu M, Saito K and Tsujihara K, Synth. Commun, 1999, 29, 2775–2781. [Google Scholar]
  • 52.Loim NM and Kelbyscheva ES, Russ. Chem. Bull, 2004, 53, 2080–2085. [Google Scholar]
  • 53.Clark JH, Chem. Rev, 1980, 80, 429–452. [Google Scholar]
  • 54.Ishihara J and Hatakeyama S, The Chemical Record, 2014, 14, 663–677. [DOI] [PubMed] [Google Scholar]
  • 55.Moloney MG, Trippier PC, Yaqoob M and Wang Z, Curr. Drug Discovery Technol, 2004, 1, 181–199. [DOI] [PubMed] [Google Scholar]
  • 56.Schwartz RE, Helms GL, Bolessa EA, Wilson KE, Giacobbe RA, Tkacz JS, Bills GF, Liesch JM, Zink DL, Curotto JE, Pramanik B and Onishi JC, Tetrahedron, 1994, 50, 1675–1686. [Google Scholar]
  • 57.Dyer J, Keeling S, King A and Moloney MG, J. Chem. Soc., Perkin Trans 1, 2000, 2793–2804. [Google Scholar]
  • 58.Bailey JH, Cherry DT, Crapnell KM, Moloney MG, Shim SB, Bamford M and Lamont RB, Tetrahedron, 1997, 53, 11731–11744. [Google Scholar]
  • 59.Silverman RB and Levy MA, J. Org. Chem, 1980, 45, 815–818. [Google Scholar]
  • 60.Beard MJ, Bailey JH, Cherry DT, Moloney MG, Shim SB, Statham K, Bamford M and Lamont RB, Tetrahedron, 1996, 52, 3719–3740 and corrigenda 1997, 3753, 1177. [Google Scholar]
  • 61.Thottathil JK, Moniot JM, Mueller RH, Wong MKY and Kissick TP, J. Org. Chem, 1986, 51, 3140–3143. [Google Scholar]
  • 62.Cowley AR, Hill TJ, Kocis P, Moloney MG, Stevenson RD and Thompson AL, Org. Biomol. Chem, 2011, 9, 7042–7056. [DOI] [PubMed] [Google Scholar]
  • 63.Postma TM and Albericio F, Org. Lett , 2014, 16, 1772–1775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Pattenden G, Thom SM and Jones MF, Tetrahedron, 1993, 49, 2131–2138. [Google Scholar]
  • 65.Onoda T, Shirai R, Koiso Y and Iwasaki S, Tetrahedron, 1996, 52, 14543–14562. [Google Scholar]
  • 66.Corey EJ and Reichard GA, J. Am. Chem. Soc, 1992, 114, 10677–10678. [Google Scholar]
  • 67.Singh SB, Goetz MA, Smith SK, Zink DL, Polishook J, Onishi R, Salowe S, Wiltsie J, Allocco J, Sigmund J, Dorso K, Cruz M. d. l., Martín J, Vicente F, Genilloud O, Donald RGK and Phillips JW, Biorg. Med. Chem. Lett, 2012, 22, 7127–7130. [DOI] [PubMed] [Google Scholar]
  • 68.Tuske S, Sarafianos SG, Wang X, Hudson B, Sineva E, Mukhopadhyay J, Birktoft JJ, Leroy O, Ismail S, Clark AD, Dharia C, Napoli A, Laptenko O, Lee J, Borukhov S, Ebright RH and Arnold E, Cell, 2005, 122, 541–552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Reusser F, Antimicrob Agents Chemother., 1976, 10, 618–622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Alt S, Mitchenall LA, Maxwell A and Heide L, Antimicrob Agents Chemother., 2011, 66, 2061–2069. [DOI] [PubMed] [Google Scholar]
  • 71.Aubry A, Pan X-S, Fisher LM, Jarlier V and Cambau E, Antimicrob Agents Chemother., 2004, 48, 1281–1288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Mukhopadhyay J, Das K, Ismail S, Koppstein D, Jang M, Hudson B, Sarafianos S, Tuske S, Patel J, Jansen R, Irschik H, Arnold E and Ebright RH, Cell, 2008, 135, 295–307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Maffioli SI, Zhang Y, Degen D, Carzaniga T, Del Gatto G, Serina S, Monciardini P, Mazzetti C, Guglierame P, Candiani G, Chiriac AI, Facchetti G, Kaltofen P, Sahl H-G, Dehò G, Donadio S and Ebright RH, Cell, 2017, 169, 1240–1248.e1223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Zhang Y, Degen D, Ho MX, Sineva E, Ebright KY, Ebright YW, Mekler V, Vahedian-Movahed H, Feng Y, Yin R, Tuske S, Irschik H, Jansen R, Maffioli S, Donadio S, Arnold E and Ebright RH, eLife, 2014, 3, e02450–e02450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Liu C-Y. 2007. “The use of single-molecule nanomanipulation to study transcription kinetics”, Ph.D. thesis, Rutgers University. [Google Scholar]
  • 76.Baldwin JE, Pratt AJ and Moloney MG, Tetrahedron, 1987, 43, 2565–2575. [Google Scholar]
  • 77.Grigoryan KR, J. Phys. Chem. A, 2009, 83, 2368–2370. [Google Scholar]
  • 78.Arakawa T, Kita Y and Timasheff SN, Biophys. Chem, 2007, 131, 62–70. [DOI] [PubMed] [Google Scholar]
  • 79.O’Neill AJ and Chopra I, Expert Opin. Drug Discov, 2004, 13, 1045–1063. [DOI] [PubMed] [Google Scholar]
  • 80.O’Shea R and Moser HE, J. Med. Chem, 2008, 51, 2871–2878. [DOI] [PubMed] [Google Scholar]
  • 81.Lipinski CA, Lombardo F, Dominy BW and Feeney PJ, Adv. Drug Delivery Rev, 2001, 46, 3–26. [DOI] [PubMed] [Google Scholar]
  • 82.Mugumbate G and Overington JP, Bioorg. Med. Chem, 2015, 23, 5218–5224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Gleeson MP, J. Med. Chem, 2008, 51, 817–834. [DOI] [PubMed] [Google Scholar]
  • 84.Fan Y, Unwalla R, Denny RA, Di L, Kerns EH, Diller DJ and Humblet C, J. Chem. Inf. Model, 2010, 50, 1123–1133. [DOI] [PubMed] [Google Scholar]
  • 85.Lovering F, Bikker J and Humblet C, J. Med. Chem, 2009, 52, 6752–6756. [DOI] [PubMed] [Google Scholar]
  • 86.Lovering F, MedChemComm, 2013, 4, 515–519. [Google Scholar]
  • 87.Pitt WR, Parry DM, Perry BG and Groom CR, J. Med. Chem, 2009, 52, 2952–2963. [DOI] [PubMed] [Google Scholar]
  • 88.Ritchie TJ and MacDonald SJF, Drug Discov. Today, 2009, 14, 1011–1020. [DOI] [PubMed] [Google Scholar]
  • 89.Congreve M, Chessari G, Tisi D and Woodhead AJ, J. Med. Chem, 2008, 51, 3661–3680. [DOI] [PubMed] [Google Scholar]
  • 90.Hajduk PJ, J. Med. Chem, 2006, 49, 6972–6976. [DOI] [PubMed] [Google Scholar]
  • 91.Cosier J and Glazer AM, J. Appl. Cryst, 1986, 19, 105–107. [Google Scholar]
  • 92.Palatinus L and Chapuis G, J. Appl. Cryst , 2007, 40, 786–790. [Google Scholar]
  • 93.Betteridge PW, Carruthers JR, Cooper RI, Prout K and Watkin DJ, J. Appl. Cryst, 2003, 36, 1487–1487. [Google Scholar]

Associated Data

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

Supplementary Materials

sup1
sup2

RESOURCES