Next Article in Journal
Arginine in the FARM and SARM: A Role in Chain-Length Determination for Arginine in the Aspartate-Rich Motifs of Isoprenyl Diphosphate Synthases from Mycobacterium tuberculosis
Previous Article in Journal
Intestinal Saturated Long-Chain Fatty Acid, Glucose and Fructose Transporters and Their Inhibition by Natural Plant Extracts in Caco-2 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Versatile Strategy for the Synthesis of 4,5-Dihydroxy-2,3-Pentanedione (DPD) and Related Compounds as Potential Modulators of Bacterial Quorum Sensing

1
Medicinal Chemistry, Taros Chemicals GmbH & Co. KG, Emil-Figge-Straße 76a, 44227 Dortmund, Germany
2
Department of Drug Sciences, Medicinal Chemistry and Pharmaceutical Technology Section, University of Pavia, Viale Taramelli 12, 27100 Pavia, Italy
3
Drug Research Program, Division of Pharmaceutical Biosciences, Faculty of Pharmacy, University of Helsinki, FI-00014 Helsinki, Finland
*
Author to whom correspondence should be addressed.
Current address: Centurion Biopharma Corporation, Engesserstraße 4, 79108 Freiburg im Breisgau, Germany.
Current address: Selvita S.A., Park Life Science, Bobrzyňskiego 14, 30-348 Krakow, Poland.
§
Current address: DE Shaw Research, 120W 45th Street, New York, NY 10036, USA.
Molecules 2018, 23(10), 2545; https://doi.org/10.3390/molecules23102545
Submission received: 16 September 2018 / Revised: 2 October 2018 / Accepted: 3 October 2018 / Published: 6 October 2018
(This article belongs to the Section Medicinal Chemistry)

Abstract

:
Resistance to antibiotics is an increasingly serious threat to global public health and its management translates to significant health care costs. The validation of new Gram-negative antibacterial targets as sources for potential new antibiotics remains a challenge for all the scientists working in this field. The interference with bacterial Quorum Sensing (QS) mechanisms represents a potentially interesting approach to control bacterial growth and pursue the next generation of antimicrobials. In this context, our research is focused on the discovery of novel compounds structurally related to (S)-4,5-dihydroxy-2,3-pentanedione, commonly known as (S)-DPD, a small signaling molecule able to modulate bacterial QS in both Gram-negative and Gram-positive bacteria. In this study, a practical and versatile synthesis of racemic DPD is presented. Compared to previously reported syntheses, the proposed strategy is short and robust: it requires only one purification step and avoids the use of expensive or hazardous starting materials as well as the use of specific equipment. It is therefore well suited to the synthesis of derivatives for pharmaceutical research, as demonstrated by four series of novel DPD-related compounds described herein.

1. Introduction

Bacterial chemical communication (i.e., quorum sensing, QS) allows bacteria to coordinate their gene expression and act as a population [1,2,3,4,5]. This phenomenon is detrimental for humans as QS regulates pathogenic processes such as the virulence factor production [6,7], susceptibility to antibiotics [8] and biofilm formation [9,10,11]. In recent decades, the modulation of QS has therefore emerged as a potential therapeutic approach to fight bacterial infections [12,13,14,15,16,17].
QS is mediated by production and release of and response to small molecules called autoinducers (AIs). Among these AIs, Autoinducer-2 (AI-2) is responsible for intra- and interspecies bacterial communication and, as a consequence, it has been termed the “universal autoinducer”. The development of small molecules able to modulate the AI-2-mediated signaling would possibly result in broad-spectrum antimicrobial activity. However, targeting the AI-2-based QS remains challenging mostly because of the rapid interconversion of the AI-2 precursor (S)-DPD (Figure 1) to several linear and cyclic forms recognized by different bacteria [18] (Figure 1). In aqueous solutions, (S)-DPD is in equilibrium with its two cyclic stereoisomers (S-DHMF and R-DHMF; Figure 1) [19]. Hydration of the C3 carbonyl group of both the cyclic and linear structures was confirmed by X-ray crystallography. In the presence of boric acid, S-THMF (Figure 1) forms a borate ester (S-THMF-borate; Figure 1) which is recognized by LuxP in V. harveyi (PDB ID: 1JX6) [20]. R-THMF instead (Figure 1) does not coordinate boron and binds to the transporter LsrB which is responsible for its internalization and acts as the active species in S. thyphimurium AI-2-mediated QS (PDB ID: 1TJY) [21]. The hydrated form of linear (S)-DPD (S-THP, Figure 1) is phosphorylated by LsrK, resulting in phospho-DPD (P-DPD, Figure 1) [22] recognized by the transcriptional repressor LsrR (PDB ID: 4L4Z) [23] and responsible for E coli and S. typhimurium AI-2-mediated signaling.
Modulation/inhibition of QS can control several bacterial virulence factors (e.g., biofilm formation) that facilitate human infections and reduce their negative effects, including mortality [24]. Quorum Sensing Inhibitors (QSI) therefore represent interesting tools to use in combination with “conventional” antibiotic therapies against antimicrobial resistance (AMR) [25,26].
In this work, we describe the set-up of a new protocol for the synthesis of racemic DPD and its application to the synthesis of four novel small libraries of DPD-related compounds (Figure 2), designed to target LsrK kinase, a key mediator in AI-2-mediated QS in enteric bacteria. The essential role of the enzyme has been demonstrated by LsrK gene deletion in E. coli, generating a mutant strain unable to activate AI-2-mediated QS [27]. Therefore, we believe that the generation of DPD-related compounds for the inhibition of LsrK may be the starting point for the development of new QSI that will serve as potential tools for overcoming antimicrobial resistance.

2. Results and Discussion

Since 2004, much effort has been devoted to the study of synthetic pathways for the preparation of DPD and analogs in both racemic and enantiomeric forms. Literature analysis revealed that the synthesis of homochiral DPD requires the use of expensive (i.e., (S)-1,4-dioxaspiro[4.5]decane-2-carboxaldehyde) or unstable (i.e., (S)-glyceraldehyde acetonide) chiral starting materials and of further time-consuming purification steps [18,28,29,30,31,32,33]. Conversely, the synthetic procedures published so far to obtain racemic DPD proceed smoothly but suffer from hazardous chemical steps (i.e., reductive ozonolysis or the use of diazomethane) [34,35].
Starting from these considerations and keeping in mind that in the initial phase of the drug discovery process racemic compounds are usually evaluated and only once the most active ones have been identified both enantiomers must be prepared for biological testing [36], herein we studied a novel versatile strategy for the synthesis of racemic DPD suitable for readily supporting practical chemical diversification. The proposed synthetic strategy leading to DPD could be useful for the preparation of C1 DPD-analogs and for the synthesis of DPD structurally related compounds, where the two carbonyl groups of DPD at C2 and C3 are embedded in heteroaromatic rings (Figure 2). To the best of our knowledge, no modification at C2 have been reported and position C3 has been barely explored and no heteroaromatic substituents (except for a furan at C1) were previously described.

2.1. Synthesis of DPD and Ph-DPD

The synthetic strategies originally evaluated are outlined in Scheme 1.
Briefly, the addition of 1-propynylmagnesium bromide to (t-butyldimethylsilyloxy)acetaldehyde [37] (1, Scheme 1), followed by the protection of the resulting secondary alcohol with TBDMSCl or TMSCl afforded compounds 3 or 4, respectively (Scheme 1). The subsequent oxidation of the internal alkyne to yield diketone 5 or 6 was performed under optimized RuO2*H2O/NaIO4-catalyzed conditions (Table 1, entry 5) using CHCl3/ACN/H2O (1:1:1) as the solvent.
The final acidic removal of the two TBDMS groups of compound 5 was performed under different conditions, but resulted in being unsuccessful (Table S1). Particularly, decomposition was observed when H2SO4 (or D2SO4) and TBAF were employed (Table S1). The partial removal of the two protecting groups (up to a maximum of 30% in total) was achieved with the use of acetic acid or Dowex50WX8 (Table S1). When the bulky protecting TBDMS group was replaced with TMS, (Scheme 1), similar results were obtained and a maximum of 40% cleavage was achieved using Dowex50WX8 in ACN-d3.
A different approach was then investigated: compound 2 and the analogous 7 were deprotected in acidic conditions (Dowex50WX8), affording diols 8 and 9, respectively. These intermediates were then protected with a cyclohexyliden group and oxidized under the previously described conditions (Table 1). The oxidation of 10 and 11 was followed by the Dowex50WX8-mediated removal of the protecting group. 1H-NMR analysis of the crude products was consistent with the literature-reported data and revealed the presence of a mixture of structures in equilibrium with each other (see SI for additional details). To further confirm the success of our procedure, the mixtures were treated with o-phenilendiamine to form, respectively, quinoxaline-DPD and quinoxaline-Ph-DPD (Scheme 1), which were isolated and fully characterized.
To sum up, the approach described above allows for the rapid production of racemic DPD in five steps and it does not require the use of dangerous or expensive reagents nor of particular equipment (i.e., ozonolysator); furthermore, only one purification step via column chromatography is necessary. Not less important, this procedure is suitable for the synthesis of C1-DPD analogs (as long as the corresponding Grignard reagent can be purchased or produced) as the synthesis of Ph-DPD demonstrated. Additionally, the ethyne function introduced in the first step is a practical synthetic handle for further chemical derivatization, as demonstrated by the four small series of derivatives described below.

2.2. Synthesis of DPD-Related Compounds

As anticipated, we designed novel DPD-related compounds in which the carbonyl groups at C2 and C3 are embedded in heteroaromatic moiety to obtain compounds stable in solution, thus avoiding the open/closed equilibrium typical of the majority of the DPD-analogs reported so far (Figure 1). As heteroaromatic rings, we selected 1,2,3-triazole and isoxazole, two scaffolds common in medicinal chemistry present in several natural and synthetic drugs including antimicrobial, anticancer, anti-inflammatory and antireumatic drugs [38,39,40,41,42,43].
The newly designed compounds can be obtained starting from the two common intermediates 15 and 16 (Scheme 2) strictly related to 2 and 7 (Scheme 1). In details, as in the case of DPD, the first of the two building blocks necessary to start the synthesis of all the analogs presented in this work was produced by the Grignard addition of ethynylmagnesium bromide to aldehyde 1, followed by acidic removal of the TBDMS protecting group. Further protection of the resulting diol 15 as acetal, using cyclohexanone dimethyl ketal, afforded the second building block compound 16 (Scheme 2).

2.2.1. 1,4- and 1,5-Disubstituted 1,2,3-Triazoles DPD-Derivatives (Series I and II)

1,2,3-triazoles (both 1,4- and 1,5-disubstituted) can be synthesized applying azide-alkyne Huisgen cycloaddition conditions where an azide is reacted with an alkyne in a 1,3-dipolar cycloaddition reaction. At first, we tested three different Copper-Catalyzed Azide-Alkyne Cycloaddition (CuAAC) conditions to select the best procedure for the synthesis of the planned compounds. For this purpose, (2-azidoethyl)benzene (17a) was chosen as the reference azide (Table 2). First, we used CuI (10% mol) and DIPEA (15% mol) in nonaqueous, nonprotic THF to afford the desired product with 58% isolated yield (Table 2, entry 1) [44]. As the addition of AcOH was found to accelerate the protonation of the Cu-C bond [45,46,47] (thus facilitating the formation of the product), a catalytic amount of AcOH was added to the mixture (containing 2% mol CuI and 4% mol DIPEA). This acid-base system jointly promoted CuAAC and resulted in a 14% increase of the isolated yield (Table 2, entry 2) when compared to the previous conditions (Table 2, entry 1) [48]. It is known that the use of ligands is beneficial for the reaction as it prevents Cu(I) oxidation and avoids the use of a base. Therefore, it is not surprising that the in situ generation of Cu(I) by the reduction of CuSO4*5H2O from sodium ascorbate together with the formation of L-ascorbic acid (that acts both as a ligand and as acidic source) raised the yield up to 89% (Table 2, entry 3) [49]. The 1,4-disubstitution was confirmed by the HMBC of compound 18a (see Supporting Information).
The corresponding 1,5-disubstituted 1,2,3-triazole 19a was also synthesized by varying the experimental conditions: the regioselective synthesis was achieved with the use of Ruthenium-catalyzed Azide-Alkyne Cycloaddition (RuAAC) conditions. Azide 17a was reacted with terminal alkyne 15 in the presence of 2% mol pentamethylcyclopentadienylbis (triphenylphosphine)ruthenium(II) chloride (Cp*RuCl(PPh3)2) regioselectively yielding, after stirring overnight the mixture in refluxing 1,4-dioxane, the corresponding 1,5-disubstituted 1,2,3-triazole 19a (Table 2, entry 4). 1H, 13C, TLC, UHPLC, and HMBC unambiguously confirmed the different nature of the two compounds (see Supporting Information) [50].
Once optimal conditions for the regioselective synthesis of 1,4-disubstituted 1,2,3-triazoles were established, we synthesized five azides of different chemical nature including aromatic, heteroaromatic and aliphatic elements (17bf). This was achieved by stirring overnight at room temperature the corresponding bromo compounds with an excess (1.5 eq) of sodium azide. The five azides were reacted with alkyne 15 applying the previously found conditions and products 18bf were isolated in good to excellent yields (60–88%, Table 2, entry 5–9).
As the synthesis of triazoles substituted with short alkyl chains (e.g., methyl, butyl) was unattainable by this route because of safety issues related to the explosive and unstable nature of the required azides, we installed the desired substituents on the triazole scaffold via alkylation. We elected to use a single, small and dangerous azide (i.e., TMSN3) over the use of four different ones. The acetal protected terminal alkyne 16 was carefully reacted with an excess (10.0 eq) of TMSN3 under previously established CuAAC conditions. The resulting unsubstituted triazole (20, Scheme 3) was both deprotected under acidic conditions (18g, Scheme 3) and, to install the desired substituents, alkylated with four different (i.e., methyl, cyclopropylmethyl, butyl, ethoxyethyl) bromides (Scheme 3).
As expected, no regioselectivity was observed and both the 1,4- and the 1,5-disubstituted 1,2,3-triazoles formed. Experimenting with base (i.e, 1.1 eq, 1.3 eq and 1.5 eq of K2CO3) and/or the alkylbromides (i.e., 0.8 eq and 0.9 eq of R1Br) stoichiometry did not consistently changed the ratio of the two regioisomers (data not shown). For each substituent, the two corresponding regioisomers were isolated by preparative HPLC. The resulting eight products (21hk and 22hk, Scheme 3) were lastly deprotected with a catalytic amount of concentrated hydrochloric acid. The ratio of the two regioisomers was determined by crude NMR. For all of the four regioisomeric pairs, the 1,4-dibustituted 1,2,3-triazoles formed in excess when compared to the respective 1,5-regioisomers and, as predictable, the ratio decreased as the sterical hindrance of the R1 substituent increased (Scheme 3). Concentrated HCl was preferred over Dowex 50WX8 for the removal of the acetal protecting group due to the shorter reaction time (1–3 h vs. overnight) and shorter workup (no filtration to remove the acidic resin required).

2.2.2. 3,5-Disubstituted Isoxazoles DPD-Derivatives (Series III and IV)

Compound 15 (Scheme 2) is also the key intermediate for the synthesis of 3,5-disubstituded DPD related compounds 26lr (Scheme 4). Briefly, aldehydes 23lr were converted into their corresponding oximes 24lr using NH2OH·HCl. The resulting crude compounds were directly chlorinated by a reaction with N-chlorosuccinimide (NCS). According to Himo et al. [49], the addition of CuSO4·5H2O, Na ascorbate, and KHCO3 in t-BuOH/H2O (1:1) to the isolated chloro-oximes allowed them to form the nitrile oxide which reacted by 1,3-dipolar cycloaddition with 15. After preparative HPLC purification, the targeted isoxazoles 26lr were, therefore, obtained in good to excellent yields (i.e., 63–89%, Scheme 4).
The same procedure was attempted to obtain 3,5-disubstituted isoxazoles of DPD-analogs bearing an amide moiety at position 3, but starting from the protected precursor 16 instead of 15 due to the cross-reactivity between the 1,3-diol and the reagents necessary in the following steps (e.g., NaOH, DIPEA, Scheme 5). Formation of the nitrile oxide for the cycloaddition was attempted using the dehydration of ethyl nitroacetate with several bases (i.e., DABCO, DMAP, DBU, NMI, Scheme S1, conditions a) and also with a combination of PhNCO/Et3N (Scheme S1, conditions b), commonly used to activate nitro groups. All of the aforementioned methods resulted in a mixture of unreacted starting materials [51,52].
The 1,3-dipole species was then changed to the chloro-oxime of ethyl glyoxalate (50% solution in toluene) but the employment of the same conditions as above (CuSO4·5H2O (5% mol), Na ascorbate (0.5 eq), KHCO3, t-BuOH/H2O (1:1), Scheme 4) did not yield the desired product while the simple use of an equimolar amount of Et3N gave only traces of 28 (Scheme 5) [49,53].
We then change our strategy and employed the oxime of ethyl glyoxalate 27 together with an excess (40.0 eq) of sodium hypochlorite, both as a chlorinating agent and as a base to form the corresponding nitrile oxide, following the procedure already described by Quan et al. [54]. Compound 28 was successfully obtained, even if with a low yield (16%). Different reaction times, as well as ratios of dipolarophile 16 and 1,3-dipole 28, were then tested (Table S2) in order to improve the initially poor yield (i.e., 16%, Table S2). Increasing the concentration of 1,3-dipole 27 enhanced the formation of intermediate 28 up to a maximum of 36% isolated yield (Table S2) with the complete consumption of the dipolarophile 16, followed by removal of the excess of 27 by column chromatography.
Once a solution for the key 1,3-dipolar cycloaddition step was found, the rest of the synthetic pathway proceeded smoothly (Scheme 5). Saponification of the ethylic ester was followed by the amidification of the resulting carboxylic acid moiety using HOBt as the coupling agent and employing both primary and secondary amines (aromatic, heteroaromatic, aliphatic). The final acidic removal of the acetal protecting group afforded six 3,5-disubstituted isoxazoles (with an amide moiety at position 3) 33b, 33sz in moderate to excellent yields (i.e., 37–79%, Scheme 5). Two more products were isolated after the acidic deprotection of intermediates 28 and 30 (i.e., 29 and 31, respectively, Scheme 5).

2.3. Biological Evaluation of Synthetized Compounds

The activity of the synthetized compounds was evaluated with a bioluminescence-based assay against the target enzyme. Our results clearly highlight that racemic DPD prepared using our procedure is efficiently phosphorylated by LsrK (see Supporting Information, Figure S1). In fact, the level of ATP is significantly reduced by the addition of racemic DPD, resulting in a light emission lower than the sample including only LsrK and ATP.
These results confirmed the validity of the approach adopted. Indeed, in this initial phase of the drug discovery process, we prepared racemic DPD and studied a versatile synthesis suitable for readily supporting practical chemical diversification racemic compounds. Only once the most active ones have been identified will both enantiomers be prepared for biological testing. Accordingly, the activity of racemic DPD is essential for demonstrating that our approach has a valid basis. Regarding the DPD-derivatives, unfortunately, they did not show any activity (the data are reported in Supporting Information, Table S3).

3. Experimental

3.1. Chemistry

Chemicals and solvents were obtained from commercial suppliers and were used without further purification. All dry reactions were performed under a nitrogen atmosphere using commercial dry solvents. Flash column chromatography was performed on a silica column using 230-400 mesh silica gel or the Grace Reveleris X2 flash chromatography system using silica gel packed Macherey Nagel Chromabond Flash BT cartridges (60 Å, 45 μm) and Grace Reveleris flash Cartridges (60 Å, 40 μm). Thin layer chromatography was performed on Macherey Nagel precoated TLC aluminum sheets with silica gel 60 UV254 (5–17 μm). TLC visualization was accomplished by irradiation with a UV lamp (254 nm) and/or staining with KMnO4 solutions. 1H-NMR spectra were recorded at room temperature on a Bruker Avance spectrometer operating at 300 MHz (Hamburg, Germany). Chemical shifts are given in ppm (δ) from tetramethylsilane as an internal standard or residual solvent peak. Significant 1H-NMR data are tabulated in the following order: multiplicity (s, singlet; d, doublet; t, triplet; q, quartet; m, multiplet; dd, doublet of doublets; dt, doublet of triplets; td, triplet of doublets; br, broad), coupling constant(s) in hertz, number of protons. Proton decoupled 13C-NMR data were acquired at 100 MHz. 13C chemical shifts are reported in parts per million (δ, ppm). All NMR data were collected at room temperature (25 °C). Analytical, preparative HPLC and Electron Spray Ionization (ESI) mass spectra were performed on an Agilent UHPLC (1290 Infinity, Santa Clara, CA, USA) and an Agilent Prep-HPLC (1260 Infinity), both equipped with a Diode Array Detector and a Quadrupole MS using mixture gradients of formic acid/water/acetonitrile as solvents. High-resolution electrospray ionization mass spectra (ESI-FTMS) were recorded on a Thermo LTQ Orbitrap (Thermo Electron, Dreieich, Germany) coupled to an ‘Accela’ HPLC system supplied with a ‘Hypersil GOLD’ column (Termo Electron).

3.2. Synthesis of DPD and Ph-DPD

Synthesis of 2 and 7: to a stirred solution of (t-butyldimethylsilyloxy)acetaldehyde (1.0 eq) in dry THF, 1-propynylmagnesium bromide was added (over 15 min; 0.5 M in THF, 1.3 eq) (or phenylethynylmagnesium bromide (1.0 M in THF, 1.3 eq) at 0 °C). After the addition, the reaction was allowed to reach room temperature and stirred for 3 h. The solvent was removed under reduced pressure, the residue was poured into a cold saturated solution of NH4Cl and extracted three times with Et2O. The organic layer was washed twice with water and once with brine, dried over MgSO4, filtered and concentrated in vacuo to yield 2 as a yellowish oil (96%) or 7 as a yellow oil (98%).
1-[(t-Butyldimethylsilyl)oxy]pent-3-yn-2-ol (2): yellowish oil, 96%, Rf = 0.20 (CyH/EtOAc 9:1). 1H-NMR (300 MHz, CDCl3) δ 4.36–4.34 (m, 1H), 3.73 (dd, J = 3.6 Hz, J = 10.0 Hz, 1H), 3.59 (dd, J = 7.7 Hz, J = 10.0 Hz,1H), 2.57 (s br, 1H), 1.83 (d, J = 1.9 Hz, 3H), 0.91 (s, 9H), 0.08 (d, J = 1.3 Hz, 6H) ppm; 13C-NMR (100 MHz, CDCl3) δ 81.8, 79.6, 67.3, 66.3, 25.8, 18.3, 3.5, −5.4 ppm [55].
1-[(t-Butyldimethylsilyl)oxy]-4-phenylbut-3-yn-2-ol (7): yellow oil, 98%, Rf = 0.72 (CyH/EtOAc 9:1). 1H-NMR (300 MHz, CDCl3) δ 7.45–7.42 (m, 2H), 7.32–7.29 (m, 3H), 4.65–4.60 (m, 1H), 3.87 (dd, J = 3.8 Hz, J = 10.0 Hz, 1H), 3.75 (dd, J = 6.9 Hz, J = 10.0 Hz, 1H), 2.71 (d, J = 4.9 Hz, 1H), 0.93 (s, 9H), 0.13 (d, J = 3.1 Hz, 6H) ppm; 13C-NMR (100 MHz, CDCl3) δ 131.8, 128.4, 128.2, 122.5, 87.0, 85.3, 67.0, 63.6, 25.9, 18.4, 5.3 ppm [56].
Synthesis of 8 and 9: to a stirred solution of 2 (or 7) (1.0 eq) in MeOH, Dowex50WX8 100–200 mesh (100 mg/1 mL) was added. The reaction was stirred at room temperature overnight. The mixture was filtered through paper and the solvent was evaporated under reduced pressure to yield 8 as an orange oil (98%) or 9 as an orange oil (97%).
Pent-3-yne-1,2-diol (8): orange oil, 98%, Rf = 0.38 (CHCl3/MeOH 9:1). 1H-NMR (300 MHz, CDCl3) δ 4.44–4.39 (m, 1H), 3.70 (dd, J = 3.8 Hz, J = 11.3 Hz, 1H), 3.62 (dd, J = 6.6 Hz, J = 11.3 Hz, 1H), 2.41 (s br, 2H), 1.85 (d, J = 2.1 Hz, 3H) ppm; 13C-NMR (100 MHz, CDCl3) δ 82.8, 79.7, 66.8, 63.4, 3.5 ppm [32].
4-Phenylbut-3-yne-1,2-diol (9): orange oil, 97%, Rf = 0.44 (CHCl3/MeOH 9:1). 1H-NMR (300 MHz, CDCl3) δ 7.46–7.42 (m, 2H), 7.36–7.29 (m, 3H), 4.69 (dd, J = 3.9 Hz, J = 6.5 Hz, 1H), 3.87–3.74 (m, 2H), 2.23 (s br, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 131.8, 128.7, 128.3, 122.0, 86.5, 86.3, 66.6, 63.7 ppm [56].
Synthesis of 10 and 11: to 8 (or 9) (1.0 eq) cyclohexanone dimethyl ketal (3.0 eq) and a catalytic amount of p-TSA was added. The reaction was stirred at room temperature overnight. The solvent was removed under reduced pressure and the crude was re-dissolved in Et2O and washed three times with NaHCO3. The organic layer was dried over MgSO4, filtered and concentrated in vacuo to yield 10 as a yellow oil (64%) or 11 as a yellow oil (72%).
2-(Prop-1-yn-1-yl)-1,4-dioxaspiro[4.5]decane (10): yellow oil, 64%, Rf = 0.50 (CyH/EtOAc 9:1). 1H-NMR (300 MHz, CDCl3) δ 4.70–4.64 (m, 1H), 4.11 (dd, J = 6.2 Hz, J = 7.9 Hz, 1H), 3.81 (t, J = 7.5 Hz, 1H), 1.85 (d, J = 2.1 Hz, 3H), 1.74–1.70 (m, 2H), 1.65–1.57 (m, 6H), 1.43–1.38 (m, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 110.5, 85.8, 82.3, 69.7, 65.5, 35.8, 25.1, 23.9, 3.7 ppm [32].
2-(2-Phenylethynyl)-1,4-dioxaspiro[4.5]decane (11): yellow oil, 72%, Rf = 0.60 (CyH/EtOAc 9:1). 1H-NMR (300 MHz, CDCl3) δ 7.46–7.42 (m, 2H), 7.33–7.28 (m, 3H), 4.95 (t, J = 6.4 Hz, 1H), 4.23 (dd, J = 6.3 Hz, J = 7.9 Hz, 1H), 4.01 (dd, J = 6.5 Hz, J = 7.9 Hz, 1H), 1.81–1.77 (m, 2H), 1.68–1.56 (m, 6H), 1.44–1.41 (m, 2H), ppm; 13C-NMR (100 MHz, CDCl3) δ 131.8, 128.5, 128.2, 122.4, 111.0, 86.6, 85.6, 69.7, 65.7, 35.5, 25.1, 23.9 ppm.
Synthesis of 12 and 13: to a stirred solution of 10 (or 11) (1.0 eq) in a 1:1:1 mixture of CHCl3/ACN/H2O, NaIO4 (4.4 eq) and RuO2·H2O (2.5% mol) were added. The mixture was vigorously stirred at room temperature overnight. The solvent was evaporated under reduced pressure and the crude was re-dissolved in CHCl3 and filtered through a silica pad. The eluate was washed three times with water, dried over MgSO4, filtered and concentrated in vacuo to yield 12 as a yellow oil (54%) or 13 as yellow oil (47%).
1-{1,4-Dioxaspiro[4.5]decan-2-yl}propane-1,2-dione (12): yellow oil, 54%, Rf = 0.42 (CyH/EtOAc 3:1). 1H-NMR (300 MHz, CDCl3) δ 5.14 (dd, J = 5.3 Hz, J = 7.9 Hz, 1H), 4.35 (dd, J = 8.0 Hz, J = 8.9 Hz, 1H), 3.99 (dd, J = 5.3 Hz, J = 8.9 Hz, 1H), 2.39 (s, 3H), 1.66–1.57 (m, 8H), 1.45–1.42 (m, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 197.5, 190.0, 109.2, 75.9, 66.9, 36.4, 35.6, 25.9, 24.8, 24.0 ppm [28].
1-{1,4-Dioxaspiro[4.5]decan-2-yl}-2-phenylethane-1,2-dione (13): yellow oil, 47%, Rf = 0.46 (CyH/EtOAc 9:1). 1H-NMR (300 MHz, CDCl3), δ 7.98 (d, J = 7.1 Hz, 2H), 7.66 (d, J = 7.4 Hz, 1H), 7.51 (t, J = 7.7 Hz, 2H), 5.12 (t, J = 6.2 Hz, 1H), 4.34 (d, J = 6.4 Hz, 2H), 1.65–1.49 (m, 8H), 1.39–1.32 (m, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 200.5, 193.0, 134.9, 132.2, 129.9, 128.9, 112.2, 77.9, 65.9, 35.4, 34.6, 24.9, 23.8, 23.0 ppm.
Synthesis of DPD and Ph-DPD: to a stirred solution of 12 (or 13) (10 mM) in D2O, Dowex 50WX8 resin was added (100 mg/1 mL). The mixture was stirred at room temperature overnight. The mixture was filtered to remove the resin and extracted with CDCl3 to remove the released cyclohexanone.
4,5-Dihydroxy-2,3-pentanedione (DPD): 1H-NMR (300 MHz, D2O) δ 4.41–4.37 (m, 1H), 4.21–4.14 (m, 2H), 4.07 (dd, J = 3.2 Hz, J = 6.0 Hz, 1H), 3.99 (dd, J = 3.8 Hz, J = 7.4 Hz, 1H), 3.86–3.78 (m, 2H), 3.69–3.65 (m, 1H), 3.59 (dd, J = 5.6 Hz, J = 9.4 Hz, 1H), 2.39 (s, 3H), 1.46 (s, 3H), 1.43 (s, 3H) ppm [34]. The NMR shows that some cyclohexanone is left as two multiplets at 1.88–1.86 and 1.75–1.74 ppm.
3,4-Dihydroxy-1-phenylbutane-1,2-dione (Ph-DPD): 1H-NMR (300 MHz, D2O) δ 8.25–8.15 (m, 2H), 8.07–7.92 (m, 2H), 7.73–7.68 (m, 1H), 7.62–7.59 (m, 5H), 7.48–7.46 (m, 5H), 4.49–4.42 (m, 1H), 4.40–4.36 (m, 1H), 4.13 (dd, J = 2.7 Hz, J = 5.6 Hz, 1H), 4.09 (d, J = 2.8 Hz, 1H), 4.06 (d, J = 2.6 Hz, 1H), 3.88 (d, J = 4.0 Hz, 1H), 3.85–3.79 (m, 1H), 3.73–3.66 (m, 1H) ppm [30].
Synthesis of quinoxaline-DPD and quinoxaline-Ph-DPD: to a stirred solution of DPD (or Ph-DPD) in D2O, o-phenylendiamine (2.0 eq) was added. The reaction was stirred at room temperature overnight. The solvent was evaporated under reduced pressure, the crude was re-dissolved in ACN (1 mL), filtered and purified by preparative HPLC.
1-(3-Methylquinoxalin-2-yl)ethane-1,2-diol (Quinoxaline-DPD): orange solid, Rf = 0.52 (CHCl3/MeOH 9:1). 1H-NMR (700 MHz, MeOD) δ 8.09–8.07 (m, 1H), 7.98–7.97 (m, 1H), 7.76 (pd, J = 7.0 Hz, J = 1.6 Hz, 2H), 5.15–5.13 (m, 1H), 4.02 (dd, J = 11.4 Hz, J = 5.4 Hz, 1H), 3.96 (dd, J = 11.4 Hz, J = 6.3 Hz, 1H), 2.84 (s, 3H) ppm; 13C-NMR (176 MHz, MeOD) δ 156.6, 154.7, 142.3, 141. 8, 131.2, 130.5, 129.9, 128.8, 72.9, 66.3, 22.3 ppm; HRMS (ESI-MS) calcd. for C11H12N2O2 [M + H]+ = 205.0899. Found: 205.0972. The NMR was consistent with previously reported data [18]. The NMR was measured with a Bruker DRX700 (700 MHz).
1-(3-Phenylquinoxalin-2-yl)ethane-1,2-diol (Quinoxaline-Ph-DPD): orange solid, Rf = 0.48 (CHCl3/MeOH 9:1). 1H-NMR (700 MHz, CDCl3) δ 8.18 (dd, J = 6.4 Hz, J = 3.3 Hz, 1H), 8.13 (dd, J = 6.1 Hz, J = 3.6 Hz, 1H), 7.82 (dd, J = 6.4 Hz, J = 3.4 Hz, 2H), 7.68 (dd, J = 7.8 Hz, J = 1.3 Hz, 2H), 7.57–7.53 (m, 3H), 5.30 (dd, J = 4.9 Hz, J = 3.6 Hz, 1H), 3.74 (dd, J = 11.7 Hz, J = 3.4 Hz, 1H), 3.54 (dd, J = 11.7 Hz, 5.1 Hz, 1H) ppm; 13C-NMR (176 MHz, CDCl3) δ 153.7, 152.5, 141.8, 139.6, 137.6, 130.5, 130.4, 129.6, 129.4, 129.0, 128.8, 128.4, 70.6, 65.7 ppm. HRMS (ESI-MS) calcd. for C16H14N2O2 [M + H]+ = 267.1055. Found: 267.1129. The NMR was consistent with previously reported data [57]. The NMR was measured with a Bruker DRX700 (700 MHz).

3.3. General Procedures for the Synthesis of 1,4- and 1,5-Disubstituted Triazoles DPD-Derivatives (Series I and II)

Synthesis of 1-[(t-butyldimethylsilyl)oxy]but-3-yn-2-ol (14): to a stirred solution of (t-butyldimethylsilyloxy)acetaldehyde (1.0 eq) in dry THF, ethynylmagnesium bromide (0.50 M in THF, 1.3 eq) was added over 15 minutes at 0 °C. After the addition, the reaction was allowed to reach room temperature and stirred for 3 hours. The solvent was removed under reduced pressure, the residue was poured into a cold saturated solution of NH4Cl and extracted three times with Et2O. The organic layer was washed twice with water and once with brine, dried over MgSO4, filtered and concentrated in vacuo to yield 14 as a yellow oil, 99%, Rf = 0.55 (CyH/EtOAc 3:1). 1H-NMR (300 MHz, CDCl3) δ 4.40–4.37 (m, 1H), 3.79 (dd, J = 3.8 Hz, J = 10.1 Hz, 1H), 3.66 (dd, J = 6.8 Hz, J = 10.0 Hz, 1H), 2.62 (d, J = 5.1 Hz, 1H), 2.42 (d, J = 2.2 Hz, 1H), 0.91 (s, 9H), 0.10 (d, J = 1.5 Hz, 6H) ppm; 13C-NMR (100 MHz, CDCl3) δ 81.9, 73.4, 66.8, 62.9, 25.8, 18.3, −5.4 ppm [58].
Synthesis of but-3-yne-1,2-diol (15): to a stirred solution of 14 in MeOH, a Dowex50WX8 100–200 mesh (100 mg/1 mL) was added. The reaction was stirred at room temperature overnight. The mixture was filtered through paper and the solvent was evaporated under reduced pressure to yield 15 as an orange oil, 99%, Rf = 0.50 (CHCl3/MeOH 9:1). 1H-NMR (300 MHz, CDCl3) δ 4.49–4.45 (m, 1H), 3.80–3.68 (m, 1H), 2.51 (d, J = 2.2 Hz, 1H), 2.31 (s, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 81.5, 74.3, 66.3, 63.0 ppm [59].
Synthesis of 2-ethynyl-1,4-dioxaspiro[4.5]decane (16): to 15 (1.0 eq) cyclohexanone dimethyl ketal (10.0 eq) and a catalytic amount of p-TSA were added. The reaction was stirred at room temperature overnight. The solvent was removed under reduced pressure and the crude was re-dissolved in Et2O and washed three times with NaHCO3. The organic layer was dried over MgSO4, filtered and concentrated in vacuo to yield 16 as a yellow oil, 57%, Rf = 0.42 (CyH/EtOAc 9:1). 1H-NMR (300 MHz, CDCl3) δ 4.71 (dt, J = 2.0 Hz, J = 6.3 Hz, 1H), 4.16 (dd, J = 6.4 Hz, J = 8.0 Hz, 1H), 3.94 (dd, J = 6.3 Hz, J = 8.0 Hz, 1H), 2.48 (d, J = 2.0 Hz, 1H), 1.77–1.72 (m, 2H), 1.65–1.59 (m, 6H), 1.42–1.39 (m, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 111.2, 81.6, 73.7, 69.5, 64.9, 35.6, 25.0, 23.8 ppm [30].
General procedure for the synthesis of 17a–f: to a stirred suspension of NaN3 (1.5 eq) in DMSO (5 mL), the corresponding bromo compound (1.0 eq) was added. The reaction was stirred at room temperature overnight. The mixture was diluted with diethyl ether and extracted five times with water and once with brine, dried over MgSO4, filtered and concentrated in vacuo to yield the desired azide as a colorless/yellowish oil.
General procedure for the synthesis of 18a–f: to a stirred solution of 15 (1.0 eq) in a 1:1 mixture of H2O/t-BuOH, the corresponding azide (1.0 eq), sodium ascorbate (0.5 eq) and CuSO4·5H2O (5% mol) were added. The reaction was stirred at room temperature overnight. The solvent was evaporated under reduced pressure, the crude was redissolved in ACN (1 mL), filtered and purified by preparative HPLC [49].
1-[1-(2-Phenylethyl)-1H-1,2,3-triazol-4-yl]ethane-1,2-diol (18a): orange oil, 89%, Rf = 0.24 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.80, m/z = 234.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.56 (s, 1H), 7.31–7.20 (m, 3H), 7.15 (d, J = 6.7 Hz, 2H), 4.75 (dd, J = 4.3 Hz, J = 6.7 Hz, 1H), 4.57 (t, J = 7.3 Hz, 2H), 3.72 (dd, J = 4.2 Hz, J = 11.2 Hz, 1H), 3.60 (dd, J = 6.9 Hz, J = 11.2 Hz, 1H), 3.17 (t, J = 7.2 Hz, 2H), 2.23 (s br, 1H) ppm; 13C-NMR (100 MHz, CD3CN) δ 149.4, 138.8, 129.7, 129.4, 127.6, 122.8, 68.5, 66.8, 51.9, 36.9 ppm.
1-(1-Benzyl-1H-1,2,3-triazol-4-yl)ethane-1,2-diol (18b): yellowish oil, 60%, Rf = 0.24 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.70, m/z = 220.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.72 (s, 1H), 7.38–7.29 (m, 5H), 5.51 (s, 2H), 4.79 (dd, J = 4.4 Hz, J = 6.3 Hz, 1H), 3.75 (dd, J = 4.2 Hz, J = 11.2 Hz, 1H), 3.64 (dd, J = 6.7 Hz, J = 11.2 Hz, 1H) ppm; 13C-NMR (100 MHz, CD3CN) δ 150.1, 136.9, 129.8, 129.3, 128.9, 123.0, 68.5, 66.7, 54.4 ppm.
1-{1-[.2-(2-Fluorophenyl)ethyl]-1H-1,2,3-triazol-4-yl}ethane-1,2-diol (18c): colorless oil, 62%, Rf = 0.49 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.82, m/z = 252.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.58 (s, 1H), 7.30–7.23 (m, 1H), 7.14–7.04(m, 3H), 4.75(dd, J = 4.3 Hz, J = 6.8 Hz, 1H), 4.58 (t, J = 7.1 Hz, 2H), 3.71 (dd, J = 4.2 Hz, J = 11.2 Hz, 1H), 3.59 (dd, J = 6.9 Hz, J = 11.2 Hz, 1H), 3.21 (t, J = 7.1 Hz, 2H), 2.23 (s br, 1H) ppm; 13C-NMR (100 MHz, CD3CN) δ 162.8 (d, J = 243.8 Hz), 150.2, 132.9 (d, J = 4.6 Hz), 130.6 (d, J = 8.2 Hz), 126.2, 126.0 (d, J = 3.5 Hz), 123.6, 116.8 (d, J = 22.0 Hz), 69.2, 67.5, 51.2, 31.3 (d, J = 2.4 Hz) ppm.
1-{1-[2-(Pyridin-2-yl)ethyl]-1H-1,2,3-triazol-4-yl}ethane-1,2-diol (18d): yellow oil, 88%, Rf = 0.28 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 0.38, m/z = 232.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 8.51 (s, 1H), 7.66–7.59 (m, 2H), 7.21–7.12 (m, 2H), 4.75 (t, J = 7.1 Hz, 3H), 3.71 (dd, J = 4.1 Hz, J = 11.2 Hz, 1H), 3.59 (dd, J = 6.8 Hz, J = 11.1 Hz, 1H), 3.33 (t, J = 7.1 Hz, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 158.5, 150.3, 143.3, 137.5, 124.4, 123.2, 122.8, 68.5, 66.8, 50.0, 38.8 ppm.
6-[4-(1,2-Dihydroxyethyl)-1H-1,2,3-triazol-1-yl]hexanenitrile (18e): orange oil, 72%, Rf = 0.54 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.39, m/z = 225.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.68 (s, 1H), 4.78 (s, 1H), 4.34 (dt, J = 1.9 Hz, J = 7.1 Hz, 2H), 3.76–3.73 (m, 1H), 3.66–3.57 (m, 1H), 3.08 (s br, 1H), 2.37 (dt, J = 1.9 Hz, J = 7.1 Hz, 2H), 1.88–1.83 (m, 2H), 1.68–1.58 (m, 2H), 1.44–1.34 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 148.4, 121.5, 119.8, 67.3, 65.6, 49.2, 28.8, 24.9, 24.2, 16.0 ppm.
1-[1-(2-Cyclohexylethyl)-1H-1,2,3-triazol-4-yl]ethane-1,2-diol (18f): orange oil, 73%, Rf = 0.43 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 2.20, m/z = 240.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.69 (s, 1H), 4.78 (t, J = 5.3 Hz, 1H), 4.35 (t, J = 7.5 Hz, 2H), 3.74 (s, 1H), 3.67–3.62 (m, 1H), 2.21 (s br, 1H), 1.78–1.63 (m, 7H), 1.24–1.17 (m, 4H), 1.0–0.94 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 150.4, 123.5, 69.2, 67.5, 49.4, 39.0, 36.4, 34.2, 27.8, 27.5 ppm.
Synthesis of 1-[1-(2-phenylethyl)-1H-1,2,3-triazol-5-yl]ethane-1,2-diol (19a): to a stirred solution of 15 (1.0 eq) in 1,4-dioxane, (2-azidoethyl)benzene (17a) (1.0 eq) and Cp*RuCl(PPh3)2 (2% mol) were added. The reaction was stirred at reflux overnight. The solvent was evaporated under reduced pressure, the crude was re-dissolved in ACN (1 mL), filtered and purified by preparative HPLC to yield 19a as a yellow solid, 87%, Rf = 0.17 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.75, m/z = 234.2 [M + H] +. 1H-NMR (300 MHz, CD3CN) δ 7.51 (s, 1H), 7.28 (t, J = 7.3 Hz, 2H), 7.23 (t, J = 7.3 Hz, 1H), 7.15 (d, J = 7.1 Hz, 2H), 4.62 (d, J = 6.3 Hz, 1H), 4.61–4.58 (m, 2H), 3.70 (s br, 1H), 3.62 (dd, J = 6.6 Hz, J = 11.3 Hz, 1H), 3.56 (dd, J = 4.9 Hz, J = 11.3 Hz, 1H), 3.21 (t, J = 7.4 Hz, 2H), 2.21 (s, 1H) ppm; 13C-NMR (100 MHz, CD3CN) δ 139.0, 138.5, 132.1, 129.7, 129.4, 127.6, 65.5 (d, J = 7.6 Hz), 50.4, 37.0 ppm.
Synthesis of 4-{1,4-dioxaspiro[4.5]decan-2-yl}-1H-1,2,3-triazole (20): to a stirred solution of 16 (1.0 eq) in a 1:1 mixture of H2O/t-BuOH, trimethylsilyl azide (10.0 eq), sodium ascorbate (0.5 eq) and CuSO4·5H2O (5% mol) were added. The reaction was stirred at room temperature overnight. The solvent was evaporated under reduced pressure, the crude was re-dissolved in EtOAc and extracted three times with water. The organic layer was dried over MgSO4, filtered and concentrated in vacuo. The crude was purified using CyH/TBME (3:1) as an eluent to yield 20 as a yellowish oil, 36%, Rf = 0.61 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 2.13, m/z = 210.2 [M + H] +. 1H-NMR (300 MHz, CD3CN) δ 7.72 (s, 1H), 5.25 (t, J = 6.6 Hz, 1H), 4.31–4.26 (m, 1H), 4.00–3.95 (m, 1H), 2.37 (s br, 1H), 1.64–1.58 (m, 8H), 1.43–1.40 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 147.4, 130.0, 111.1, 70.9, 69.8, 36.8, 35.9, 25.8, 24.7, 24.6 ppm.
General procedure for the synthesis of 21h–k and 22h–k: to a stirred solution of 20 (1.0 eq) in dry THF, K2CO3 (2.0 eq) and the corresponding alkyl halide (bromide or iodide) were added. The reaction was stirred at reflux overnight. The solvent was evaporated under reduced pressure, the crude was re-dissolved in ACN (1 mL), filtered and purified by preparative HPLC.
For each compound, two different fractions were isolated corresponding to the 1,4- and 1,5-disubstitued products. The different substitution was determined by HMBC of two representative samples (21i, 22i).
General procedure for the synthesis of 18g–k and 19h–k: a stirred solution of 20 (or 21hk, or 22hk) in 1,4-dioxane was cooled to 0 °C using an ice bath. A catalytic amount of 12 M HCl was added. The reaction was stirred at room temperature overnight. The solvent was evaporated under reduced pressure; the crude was re-dissolved in Et2O and extracted with water. The aqueous layer was extracted three times with Et2O and dried in vacuo to yield the corresponding products 18gk and 19hk.
1-(1H-1,2,3-Triazol-4-yl)ethane-1,2-diol (18g): yellowish oil, 98%, Rf = 0.13 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 0.32, m/z = 130.3 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 8.38 (s, 1H), 5.05 (t, J = 5.5 Hz, 1H), 3.86–3.74 (m, 2H) ppm; 13C-NMR (100 MHz, MeOD) δ 146.0, 126.4, 67.0, 66.2 ppm.
1-(1-Methyl-1H-1,2,3-triazol-4-yl)ethane-1,2-diol (18h): colorless oil, 66%, Rf = 0.29 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 0.45, m/z = 144.2 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 8.40 (s, 1H), 5.04 (t, J = 5.5 Hz, 1H), 4.35 (s, 3H), 3.94–3.80 (m, 2H) ppm; 13C-NMR (100 MHz, MeOD) δ 151.9, 133.5, 70.0, 68.2, 42.9 ppm.
1-(1-Methyl-1H-1,2,3-triazol-5-yl)ethane-1,2-diol (19h): colorless oil, 65%, Rf = 0.32 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 0.65, m/z = 144.1 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 7.60 (s, 1H), 4.81 (dd, J = 4.8 Hz, J = 6.8 Hz, 1H), 4.13 (s, 3H), 3.81–3.68 (m, 2H) ppm; 13C-NMR (100 MHz, MeOD) δ 150.6, 133.4, 68.7, 67.0, 41.7 ppm.
1-[1-(Cyclopropylmethyl)-1H-1,2,3-triazol-4-yl]ethane-1,2-diol (18i): colorless oil, 73%, Rf = 0.32 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.25, m/z = 184.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 8.29 (d, J = 9.1 Hz, 1H), 6.21 (s br, 2H), 5.02 (dd, J = 3.9 Hz, J = 6.4 Hz, 1H), 4.36 (dd, J = 2.5 Hz, J = 7.4 Hz, 2H), 3.80 (d, J = 5.3 Hz, 2H), 1.40–1.35 (m, 1H), 0.70–0.64 (m, 2H), 0.52–0.47 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 146.8, 125.8, 66.9, 65.7, 57.8, 11.1, 4.5 ppm.
1-[1-(Cyclopropylmethyl)-1H-1,2,3-triazol-5-yl]ethane-1,2-diol (19i): colorless oil, 77%, Rf = 0.35 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.31, m/z = 184.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.87 (s, 1H), 4.91 (t, J = 5.5 Hz, 1H), 4.34 (d, J = 7.3 Hz, 2H), 3.76 (d, J = 5.5 Hz, 2H), 1.43–1.38 (m, 1H), 0.64–0.58 (m, 2H), 0.49–0.46 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 140.3, 130.7, 65.5, 65.4, 55.2, 11.7, 4.6 ppm.
1-(1-Butyl-1H-1,2,3-triazol-4-yl)ethane-1,2-diol (18j): colorless oil, 85%, Rf = 0.37 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.45, m/z = 186.2 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 8.28 (s, 1H), 4.77 (t, J = 5.5 Hz, 1H), 4.39 (t, J = 7.2 Hz, 2H), 3.63–3.49 (m, 2H), 1.81–1.71(m, 2H), 1.23–1.11 (m, 2H), 0.76 (t, J = 7.4 Hz, 3H) ppm; 13C-NMR (100 MHz, MeOD) δ 147.5, 126.7, 67.2, 66.1, 53.8, 32.6, 20.5, 13.7 ppm.
1-(1-Butyl-1H-1,2,3-triazol-5-yl)ethane-1,2-diol (19j): colorless oil, 95%, Rf = 0.39 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.50, m/z = 186.2 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 8.21 (s, 1H), 4.98 (t, J = 5.7 Hz, 1H), 4.62 (dd, J = 6.5 Hz, J = 8.4 Hz, 2H), 3.91–3.78 (m, 2H), 2.04–1.94 (m, 2H), 1.49–1.37 (m, 2H), 1.00 (t, J = 7.3 Hz, 3H) ppm; 13C-NMR (100 MHz, MeOD) δ 142.7, 130.1, 65.9, 65.8, 51.5, 32.8, 20.7, 13.8 ppm.
1-[1-(2-Ethoxyethyl)-1H-1,2,3-triazol-4-yl]ethane-1,2-diol (18k): colorless oil, 91%, Rf = 0.31 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.21, m/z = 202.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.92 (s, 1H), 4.89 (t, J = 5.1 Hz, 1H), 4.54 (t, J = 5.1 Hz, 2H), 3.83–3.68 (m, 4H), 3.47 (q, J = 7.0 Hz, 2H), 1.10 (t, J = 7.0 Hz, 3H) ppm; 13C-NMR (100 MHz, CD3CN) δ 148.3, 124.9, 68.8, 67.7, 66.9, 66.3, 52.1, 15.2 ppm.
1-[1-(2-Ethoxyethyl)-1H-1,2,3-triazol-4-yl]ethane-1,2-diol (19k): colorless oil, 82%, Rf = 0.37 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.23, m/z = 202.2 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 8.10 (s, 1H), 5.00 (t, J = 5.5 Hz, 1H), 4.74 (dd, J = 3.7 Hz, J = 5.3 Hz, 2H), 3.84 (t, J = 5.2 Hz, 2H), 3.77 (t, J = 5.8 Hz, 2H), 3.46–3.38 (m, 2H), 1.05 (t, J = 7.0 Hz, 3H) ppm; 13C-NMR (100 MHz, MeOD) δ 143.4, 130.3, 69.6, 67.7, 66.0, 65.9, 51.7, 15.3 ppm.

3.4. General Procedures for the Synthesis of 3,5-Disubstituted Isoxazoles DPD Derivatives (Series III and IV)

General procedure for the synthesis of 24l–r: to a stirred solution of the corresponding aldehyde (1.0 eq) in EtOH (10 mL), Et3N (1.5 eq) and NH2OH*HCl (1.5 eq) dissolved in water (10 mL) were added. The reaction was stirred at room temperature for 1–3 hours (monitored by TLC). The solvent was evaporated under reduced pressure; the crude was re-dissolved in EtOAc and extracted three times with water. The organic layer was dried over MgSO4, filtered and concentrated in vacuo to yield the corresponding oxime. All the resulting compounds were used in the next step without being purified.
General procedure for the synthesis of 25l–r: to a stirred solution of the corresponding oxime (1.0 eq) in DMF, N-chlorosuccinimide (1.0 eq) was added in two portions. The reaction was stirred at room temperature for 1–2 h (monitored by TLC). The crude was diluted with Et2O and extracted five times with water and once with brine. The organic layer was dried over MgSO4, filtered and concentrated in vacuo to yield the corresponding chloro-oxime. All the resulting compounds were used in the next step without being purified.
General procedure for the synthesis of 26l–r: to a stirred solution of 15 (1.0 eq) in a 1:1 mixture of H2O/t-BuOH, the corresponding chloro-oxime (1.0 eq), sodium ascorbate (0.5 eq), CuSO4·5H2O (5% mol) and KHCO3 (4.3 eq) were added. The mixture was stirred at room temperature overnight. The solvent was evaporated under reduced pressure; the crude was redissolved in ACN (1 mL), filtered and purified by preparative HPLC [49].
1-[3-(4-Methylphenyl)-1,2-oxazol-5-yl]ethane-1,2-diol (26l): white solid, 82%, Rf = 0.38 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 2.15, m/z = 220.1 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.73 (d, J = 8.0 Hz, 2H), 7.31 (d, J = 7.9 Hz, 2H), 6.67 (s, 1H), 4.82 (d, J = 5.2 Hz, 1H), 3.92 (s br, 1H), 3.83–3.71 (m, 2H), 3.14 (s br, 1H), 2.38 (s, 3H) ppm; 13C-NMR (100 MHz, CD3CN) δ 175.3, 163.7, 142.0, 131.3, 128.2, 127.9, 101.2, 69.2, 66.0, 22.0 ppm.
1-[3-(3-Chlorophenyl)-1,2-oxazol-5-yl]ethane-1,2-diol (26m): white solid, 87%, Rf = 0.51 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 2.24, m/z = 240.0 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.87 (s, 1H), 7.77 (dd, J = 5.4 Hz, J = 6.8 Hz, 1H), 7.48 (d, J = 5.8 Hz, 2H), 6.74 (s, 1H), 4.84 (t, J = 5.3 Hz, 1H), 3.78 (dq, J = 5.3 Hz, J = 11.4 Hz, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 175.2, 162.0, 135.4, 132.0, 131.6, 130.9, 127.5, 126.1, 100.8, 68.5, 65.2 ppm.
1-[3-(2,4-Difluorophenyl)-1,2-oxazol-5-yl]ethane-1,2-diol (26n): white solid, 78%, Rf = 0.49 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 2.06, m/z = 242.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 7.97–7.89 (m, 1H), 7.14–7.06 (m, 2H), 6.69 (d, J = 3.3 Hz, 1H), 4.86 (t, J = 5.3 Hz, 1H), 3.85–3.73 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 174.8, 164.9 (dd, J = 8.5 Hz, J = 246.4 Hz), 161.4 (dd, J = 8.5 Hz, J = 249.4 Hz), 157.9, 131.6 (dd, J = 4.6 Hz, J = 10.1 Hz), 114.7 (dd, J = 3.9 Hz, J = 12.6 Hz), 113.2 (dd, J = 3.6 Hz, J = 21.9 Hz), 105.6 (t, J = 26.1 Hz), 102.7 (d, J = 7.4 Hz), 68.5, 65.3 ppm.
1-[3-(Pyridin-3-yl)-1,2-oxazol-5-yl]ethane-1,2-diol (26o): orange solid, 71%, Rf = 0.22 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.14, m/z = 207.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 9.04 (s, 1H), 8.67 (s, 1H), 8.18 (d, J = 7.9 Hz, 1H), 7.46 (dd, J = 5.3 Hz, J = 7.2 Hz, 1H), 6.79 (s, 1H), 4.87 (t, J = 5.3 Hz, 1H), 3.85–3.73 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 175.4, 160.9, 151.9, 148.7, 135.0, 126.2, 124.9, 100.7, 68.5, 65.3 ppm.
1-(3-Cyclopropyl-1,2-oxazol-5-yl)ethane-1,2-diol (26p): yellow oil, 63%, Rf = 0.12 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.91, m/z = 170.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 5.98 (s, 1H), 4.71 (ddd, J = 0.6 Hz, J = 4.6 Hz, J = 6.1 Hz, 1H), 3.74–3.62 (m, 2H), 2.00–1.91 (m, 1H), 1.03–0.97 (m, 2H), 0.78–0.72 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 173.5, 167.3, 99.7, 68.3, 65.2, 8.3, 7.8 ppm.
1-[3-(Oxolan-3-yl)-1,2-oxazol-5-yl]ethane-1,2-diol (26q): yellow oil, 77%, Rf = 0.50 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.53, m/z = 200.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 6.22 (s, 1H), 4.77–4.73 (m, 1H), 4.03–3.98 (m, 1H), 3.94–3.85 (m, 1H), 3.82–3.76 (m, 1H), 3.74–3.65 (m, 2H), 3.55–3.47 (m, 1H), 2.37–2.25 (m, 2H), 2.09–2.00 (m, 1H) ppm; 13C-NMR (100 MHz, CD3CN) δ 174.1, 166.1, 101.2, 72.7, 68.5, 65.3, 63.7, 37.4, 32.5 ppm.
1-(3-Cyclohexyl-1,2-oxazol-5-yl)ethane-1,2-diol (26r): yellow oil, 89%, Rf = 0.31 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 2.13, m/z = 212.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 6.18 (s, 1H), 4.73 (t, J = 5.4 Hz, 1H), 3.71 (dq, J = 5.4 Hz, J = 11.3 Hz, 2H), 2.70 (dt, J = 3.3 Hz, J = 10.7 Hz, 1H), 2.21 (s br, 2H), 1.90–1.69 (m, 5H), 1.50–1.25 (m, 5H) ppm; 13C-NMR (100 MHz, CD3CN) δ 173.9, 169.8, 101.5, 69.2, 66.0, 37.3, 33.4, 27.3, 27.2 ppm.
Synthesis of ethyl (2E)-2-(hydroxyimino)acetate (27): to a stirred solution of ethyl glyoxalate (50% solution in toluene, 1.0 eq) in EtOH, Et3N (1.5 eq) and NH2OH·HCl (1.5 eq) dissolved in water (10 mL) were added. The reaction was stirred at room temperature for 2 hours (monitored by TLC). The solvent was evaporated under reduced pressure; the crude was re-dissolved in Et2O and extracted three times with water. The organic layer was dried over MgSO4, filtered and concentrated in vacuo to yield 27 as a colorless oil, 84%, Rf = 0.64 (CHCl3/MeOH 9:1). 1H-NMR (300 MHz, CDCl3) δ 9.83 (s br, 1H), 7.56 (s, 1H), 4.32 (q, J = 7.1 Hz, 2H), 1.34 (t, J = 7.1 Hz, 3H) ppm; 13C-NMR (100 MHz, CDCl3) δ 162.4, 141.6, 61.8, 13.8 ppm [60].
Synthesis of ethyl 5-{1,4-dioxaspiro[4.5]decan-2-yl}-1,2-oxazole-3-carboxylate (28): to a stirred solution of 16 (1.0 eq) in THF, 27 (2.0 eq) and NaOCl (40.0 eq portion wise over 12 hours) were added. The reaction was stirred at room temperature for 12 hours. The solvent was evaporated under reduced pressure, the crude was re-dissolved in DCM and washed three times with water. The organic layer was dried over MgSO4, filtered and concentrated in vacuo. The crude was re-dissolved in ACN (1 mL), filtered and purified by preparative HPLC to yield 28 as a yellowish oil, 36%, Rf = 0.57 (CyH/EtOAC 3:1), UHPLC-ESI-MS: Rt = 3.04, m/z = 282.2 [M + H]+. 1H-NMR (300 MHz, CDCl3) δ 6.67 (s, 1H), 5.23 (t, J = 6.0 Hz, 1H), 4.43 (q, J = 7.2 Hz, 2H), 4.35 (dd, J = 6.7 Hz, J = 8.6 Hz, 1H), 4.09 (dd, J = 5.4 Hz, J = 8.6 Hz, 1H), 1.71–1.62 (m, 9H), 1.41 (t, J = 7.1 Hz, 4H) ppm; 13C-NMR (100 MHz, CDCl3) δ 173.5, 159.8, 156.3, 111.9, 102.5, 69.8, 68.1, 62.2, 35.8, 34.9, 25.0, 23.9, 23.8, 14.1 ppm.
Synthesis of 5-{1,4-dioxaspiro[4.5]decan-2-yl}-1,2-oxazole-3-carboxylic acid (30): a stirred solution of 28 in THF was cooled to 0 °C using an ice bath. A solution of 10 M NaOH (5.0 eq) was added dropwise and the reaction was stirred at room temperature overnight. The solvent was evaporated under reduced pressure; the crude was re-dissolved in DCM and extracted with water. The aqueous layer was acidified with 1M HCl until pH = 1 and extracted three times with CHCl3/i-PrOH (7:3). The organic layer was dried over MgSO4, filtered and concentrated in vacuo to yield 30 as a white solid, 99%, Rf = 0.17(CHCl3/MeOH 5:1), UHPLC-ESI-MS: Rt = 2.37, m/z = 254.2 [M + H] +. 1H-NMR (300 MHz, CDCl3) δ 6.74 (s, 1H), 5.26 (t, J = 5.9 Hz, 1H), 4.38 (dd, J = 6.6 Hz, J = 8.7 Hz, 1H), 4.12 (dd, J = 5.3 Hz, J = 8.7 Hz, 1H), 1.73–1.60 (m, 8H), 1.47–1.43 (m, 2H) ppm; 13C-NMR (100 MHz, CDCl3) δ 174.2, 162.4, 155.6, 112.1, 102.8, 69.8, 68.1, 35.9, 34.9, 24.9, 23.9, 23.8 ppm.
General procedure for the synthesis of 32b, 32s–z: the reactions were performed in parallel in 15 mL reaction tubes in a 24 position Mettler-Toledo Miniblock® equipped with a heat transfer block and inert gas manifold. Each reaction tube was loaded with a previously prepared solution of 30 mg of 28 (1.0 eq) in 2 mL of DMF, DIPEA (5.0 eq), HOBt (2.0 eq), EDC·HCl (2.5 eq). Then the corresponding amine was added (2.0 eq). The reaction mixtures were stirred at room temperature overnight. The reaction conversion was confirmed through a UHPLC check of some representative samples. The mixtures were evaporated until dryness. The crudes were re-dissolved in 1.0 mL of ACN, filtered and purified with preparative HPLC (gradient acetonitrile/water with 0.1% formic acid, 2–98%).
General procedure for the synthesis of 29, 31, 33b, 33s–z: a stirred solution of 28 (or 30, or 32b, or 32sz) was cooled to 0 °C using an ice bath. A catalytic amount of concentrated HCl was added. The reactions were stirred at room temperature overnight. The solvent was evaporated under reduced pressure, the crudes were re-dissolved in ACN (1 mL), filtered and purified by preparative HPLC.
Ethyl 5-(1,2-dihydroxyethyl)-1,2-oxazole-3-carboxylate (29): colorless oil, 46%, Rf = 0.44 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.51, m/z = 202.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 6.65 (s, 1H), 4.84 (t, J = 5.3 Hz, 1H), 4.37 (q, J = 7.1 Hz, 2H), 3.76 (dd, J = 4.2 Hz, J = 5.2 Hz, 2H), 2.18 (s br, 1H), 1.35 (t, J = 7.1 Hz, 3H) ppm; 13C-NMR (100 MHz, CD3CN) δ 176.2, 160.8, 157.4, 103.0, 68.3, 65.1, 62.9, 14.3 ppm.
5-(1,2-Dihydroxyethyl)-1,2-oxazole-3-carboxylic acid (31): colorless oil, 55%, Rf = 0.11 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 0.42, m/z = 174.2 [M + H]+. 1H-NMR (300 MHz, CD3CN) δ 6.65 (s, 1H), 4.84 (t, J = 5.1 Hz, 1H), 3.81–3.70 (m, 2H) ppm; 13C-NMR (100 MHz, CD3CN) δ 176.2, 161.0, 157.2, 103.2, 68.3, 65.1 ppm.
N-Benzyl-5-(1,2-dihydroxyethyl)-1,2-oxazole-3-carboxamide (33b): white solid, 58%, Rf = 0.25 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.89, m/z = 263.2 [M + H]+. 1H-NMR (300 MHz, Acetone-d6) δ 7.40–7.27 (m, 4H), 7.26–7.22 (m, 1H), 6.68 (s, 1H), 4.91 (t, J = 5.4 Hz, 1H), 4.59 (s, 2H), 3.89–3.78 (m, 2H) ppm; 13C-NMR (100 MHz, Acetone-d6) δ 177.2, 160.6, 141.0, 137.3, 130.2, 129.4, 128.9, 102.9, 69.5, 66.4, 44.4 ppm.
5-(1,2-Dihydroxyethyl)-N-(4-fluorophenyl)-1,2-oxazole-3-carboxamide (33s): white solid, 58%, Rf = 0.25 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.97, m/z = 267.2 [M + H]+. 1H-NMR (300 MHz, Acetone-d6) δ 7.92–7.88 (m, 2H), 7.16 (t, J = 8.8 Hz, 2H), 6.76 (s, 1H), 4.95 (t, J = 5.4 Hz, 1H), 3.92–3.81 (m, 2H) ppm; 13C-NMR (100 MHz, Acetone-d6) δ 177.6, 162.8, 159.8 (d, J = 133.8 Hz), 159.6, 136.3 (d, J = 2.7 Hz), 124.1 (d, J = 7.7 Hz), 117.2 (d, J = 22.6 Hz), 103.1, 69.5, 66.4 ppm.
5-(1,2-Dihydroxyethyl)-N-[(thiophen-2-yl)methyl]-1,2-oxazole-3-carboxamide (33t): white solid, 65%, Rf = 0.34 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.79, m/z = 269.2 [M + H]+. 1H-NMR (300 MHz, Acetone-d6) δ 8.43 (s br, 0.5H), 7.32 (dd, J = 1.3 Hz, J = 5.1 Hz, 1H), 7.06 (dd, J = 1.1 Hz, J = 3.4 Hz, 1H), 6.95 (dd, J = 3.5 Hz, J = 5.1 Hz, 1H), 6.68 (s, 1H), 4.99 (s br, 0.5H), 4.90 (t, J = 5.1 Hz, 1H), 4.77–4.75 (m, 2H), 4.15 (s br, 0.5H), 3.84–3.81 (m, 2H), 2.87 (s br, 0.5H) ppm; 13C-NMR (100 MHz, Acetone-d6) δ 177.3, 160.4, 143.6, 141.4, 128.5, 127.8, 126.8, 102.9, 69.6, 66.5, 39.3 ppm.
5-(1,2-Dihydroxyethyl)-N-[(pyridin-3-yl)methyl]-1,2-oxazole-3-carboxamide (33u): yellow oil, 37%, Rf = 0.33 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 0.38, m/z = 264.2 [M + H]+. 1H-NMR (300 MHz, MeOD) δ 8.56 (s, 1H), 8.44 (s br, 1H), 7.86 (d, J = 7.9 Hz, 1H), 7.43 (dd, J = 4.9 Hz, J = 7.8 Hz, 1H), 6.70 (s, 1H), 4.84 (d, J = 5.8 Hz, 1H), 4.59 (s, 1H), 3.80 (dd, J = 3.3 Hz, J = 5.6 Hz, 2H) ppm; 13C-NMR (100 MHz, MeOD) δ 176.6, 161.6, 159.6, 149.6, 149.0, 137.8, 136.5, 125.3, 102.1, 68.8, 65.6, 41.6 ppm.
5-(1,2-Dihydroxyethyl)-N-(2-methoxyethyl)-1,2-oxazole-3-carboxamide (33v): yellow oil, 78%, Rf = 0.38 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.23, m/z = 231.2 [M + H]+. 1H-NMR (300 MHz, Acetone-d6) δ 6.64 (s, 1H), 4.90 (t, J = 5.4 Hz, 1H), 3.83 (dd, J = 3.6 Hz, J = 5.4 Hz, 2H), 3.58–3.50 (m, 4H), 3.32 (s, 3H), 2.85 (s br, 2H) ppm; 13C-NMR (100 MHz, Acetone-d6) δ 177.3, 160.5, 147.7, 102.8, 72.4, 69.6, 66.5, 59.6, 40.5 ppm.
1-[3-(Pyrrolidine-1-carbonyl)-1,2-oxazol-5-yl]ethane-1,2-diol (33z): yellow oil, 79%, Rf = 0.30 (CHCl3/MeOH 9:1), UHPLC-ESI-MS: Rt = 1.51, m/z = 227.2 [M + H] +. 1H-NMR (300 MHz, Acetone-d6) δ 6.58 (s, 1H), 4.90 (t, J = 5.3 Hz, 1H), 3.83 (t, J = 5.1 Hz, 2H), 3.76 (t, J = 6.6 Hz, 2H), 3.54 (t, J = 6.6 Hz, 2H), 2.84 (s br, 2H), 1.99–1.89 (m, 4H) ppm; 13C-NMR (100 MHz, Acetone-d6) δ 175.8, 161.7, 149.6, 104.2, 69.6, 66.6, 50.0, 48.2, 27.8, 25.5 ppm.

3.5. Biology

All chemicals were purchased from Sigma (Hamburg, Germany) if not otherwise stated. (S)-DPD was purchased from OMM Scientific (Dallas, TX, USA). The ATP Bioluminescence kit CLS II and Kinase Glo Luminescence assay kit were respectively purchased from Roche Scientific (Manheim, Germany) and Promega (Madison, WI, USA).

3.5.1. LsrK Overexpression and Purification

E. coli MET1158 (E. coli, amp resistance, BL21 (DE3) luxS-, with pMET1144 (lsrK-His in pET21b)), kindly donated by Prof. Karina Xavier (Instituto Gulbenkian de Ciência, Portugal) [61], was used for the overexpression of LsrK from S. typhimurium. The bacteria were grown overnight in 2 × YPTG (yeast, tryptone, phosphate buffer and glucose) mediums supplemented with 100 µg/mL ampicillin. At the exponential phase, protein expression was induced by the addition of 0.1 mM isopropyl β-D-1 thiogalactopyranoside for 9 h at 22 °C (250 rpm). Cells were harvested and frozen overnight before proceeding with lysis and purification, according to the literature [62].

3.5.2. DPD Activity Evaluation

Phosphorylation of DPD by LsrK was evaluated with a bioluminescence-based assay, ATP Bioluminescence kit CLSII (Roche) as previously described in Reference [61]. DPD was plated at 200 µM and 400 µM and a reaction mixture containing 200 nM Lsrk and 20 µM ATP in assay buffer (25 mM triethanolamine, pH 7.4, 200 µM MgCl2). Commercially available DPD was tested for comparison at 200 µM. The level of ATP was monitored by the ATP Bioluminescence kit CLSII following the manufacturer’s instructions. The experiment was performed in the kinetic-mode, monitoring the luminescence every 2 min within a time window of 30 min at the Varioskan LUX plate reader (Thermo Fisher Scientific, Vantaa, Finland).

3.5.3. Screening of DPD-Related Compounds

The activity of DPD-related compounds was evaluated in an LsrK inhibition assay. Compounds were plated in a 384 well-plate to a final concentration of 200 µM in triplicate. A 300 nM LsrK and 300 µM DPD diluted in an assay buffer (25 mM triethanolamine, pH 7.4, 200 µM MgCl2, 0.1 mg/mL BSA) were added to the plate followed by 100 µM ATP to start the reaction. After 15 min of reaction, the Kinase Glo Luminescence assay reagent was added according to the manufacturer’s instructions. The experiment was carried on in end-point mode and the luminescence was recorded at the Varioskan LUX plate reader.

4. Conclusions

Resistance to antibiotics poses a continuous threat to public health. In the last few decades, receptors able to modulate QS started to be considered interesting targets for anti-infective therapy and the modulation/inhibition of QS has become an appealing strategy against bacterial resistance. Several studies have already shown that interference with QS affects biofilm formation and biofilm properties (e.g., thickness, mass). Particularly, DPD, the key compound in the biosynthesis of AI-2, is able to modulate QS in both Gram-negative and Gram-positive bacteria. Accordingly, DPD-analogs may have great potential as QSI and, therefore, as antimicrobial drugs. Of note, two different DPD-related compounds (i.e. isobutyl-DPD and phenyl-DPD) in combination with gentamicin have almost completely cleared the pre-existing biofilms in E. coli and P. aeruginosa, respectively [63].
In this work, we successfully developed a new short and robust strategy for the synthesis of DPD which requires only one purification step. Ph-DPD was also synthesized to show the applicability of our protocol to the production of different C1-DPD analogs. The new strategy inspired the synthesis of 30 novel DPD-related compounds: the cycloaddition to two common precursors was employed to produce (in maximum four steps) four different small libraries where the diketo moiety of DPD was embedded in heteroaromatic rings. All the designed compounds were purified and characterized by 1H-NMR, 13C-NMR, and UHPLC-MS (purity > 90%). It is worth noting that in these compounds the open/closed equilibrium (typical of the majority of the DPD-analogs reported so far, Figure 1) is not possible. The so-obtained more stable compounds were easily purify by column chromatography. Moreover, the presence of heteroaromatic groups increases the UV absorbance and MW, rendering the compound detection by the classical analytical method (e.g., LC-MS) easier compared to previously reported analogs (e.g., ethyl-DPD).
Our new synthetic approach allowed us to synthetize a small set of racemic DPD-related compounds in a relatively easy and fast way. We demonstrated that racemic DPD is efficiently phosphorylated by LsrK, corroborating the validity of our approach. On the other hand, all compounds of our library of DPD-related did not show any activity on LsrK. Nevertheless, the synthetic procedure herein proposed might lead to the preparation of a wider compound library, thus, allowing for the discovery of a new class of LsrK inhibitors as potential antivirulence agents. Moreover, we decided to add these products to the library of MuTaLig, an innovative ligand identification platform for the drug-discovery process.

Supplementary Materials

Supplementary materials are available online.

Author Contributions

Conceptualization, S.S., F.G. and S.C.; Methodology, S.S., F.G.; Formal Analysis, F.M., M.P.; Investigation, S.S., V.G., M.P., F.G., S.C., P.T.; Resources, D.T., P.T.; Writing-Original Draft Preparation, S.S., V.G., P.T.; Writing-Review & Editing, S.S., F.G., P.T., S.C.; Supervision, S.C.; Project Administration, A.K., D.T.; Funding Acquisition, F.G., D.T.

Funding

This research was funded by the European Union’s Horizon 2020 research and innovation programme for funding INTEGRATE under the Marie Sklodowska Curie grant agreement N°. 642620 and for the Marie Skłodowska-Curie Actions Individual Fellowship to SS.

Acknowledgments

Martyna Bielska, Xenia Iwanova, Virginia Llemos and Eduard Ackerman are acknowledged for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Surette, M.G.; Miller, M.B.; Bassler, B.L. Quorum sensing in Escherichia coli, Salmonella typhimurium, and Vibrio harveyi: A new family of genes responsible for autoinducer production. Proc. Natl. Acad. Sci. USA 1999, 96, 1639–1644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Waters, C.M.; Bassler, B.L. Quorum sensing: Cell-to-cell communication in bacteria. Annu. Rev. Cell Dev. Biol. 2005, 21, 319–346. [Google Scholar] [CrossRef] [PubMed]
  3. Bassler, B.L.; Losick, R. Bacterially speaking. Cell 2006, 125, 237–246. [Google Scholar] [CrossRef] [PubMed]
  4. Lowery, C.A.; Dickerson, T.J.; Janda, K.D. Interspecies and interkingdom communication mediated by bacterial quorum sensing. Chem. Soc. Rev. 2008, 37, 1337–1346. [Google Scholar] [CrossRef] [PubMed]
  5. Ng, W.-L.; Bassler, B.L. Bacterial Quorum-Sensing Network Architectures. Annu. Rev. Genet. 2009, 43, 197–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Zhu, J.; Miller, M.B.; Vance, R.E.; Dziejman, M.; Bassler, B.L.; Mekalanos, J.J. Quorum-sensing regulators control virulence gene expression in Vibrio cholerae. Proc. Natl. Acad. Sci. USA 2002, 99, 3129–3134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Antunes, L.C.M.; Ferreira, R.B.R.; Buckner, M.M.C.; Finlay, B.B. Quorum sensing in bacterial virulence. Microbiology 2010, 156, 2271–2282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Ahmed, N.A.A.M.; Petersen, F.C.; Scheie, A.A. AI-2 quorum sensing affects antibiotic susceptibility in Streptococcus anginosus. J. Antimicrob. Chemother. 2007, 60, 49–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Davies, D.G.; Parsek, M.R.; Pearson, J.P.; Iglewski, B.H.; Costerton, J.W.; Greenberg, E.P. The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science 1998, 280, 295–298. [Google Scholar] [CrossRef] [PubMed]
  10. Rickard, A.H.; Palmer, R.J.; Blehert, D.S.; Campagna, S.R.; Semmelhack, M.F.; Egland, P.G.; Bassler, B.L.; Kolenbrander, P.E. Autoinducer 2: A concentration-dependent signal for mutualistic bacterial biofilm growth. Mol. Microbiol. 2006, 60, 1446–1456. [Google Scholar] [CrossRef] [PubMed]
  11. Irie, Y.; Parsek, M.R. Quorum sensing and microbial biofilms. Curr. Top. Microbiol. Immunol. 2008, 322, 67–84. [Google Scholar] [PubMed]
  12. Smith, K.M.; Bu, Y.; Suga, H. Library screening for synthetic agonists and antagonists of a Pseudomonas aeruginosa autoinducer. Chem. Biol. 2003, 10, 563–571. [Google Scholar] [CrossRef]
  13. Suga, H.; Smith, K.M. Molecular mechanisms of bacterial quorum sensing as a new drug target. Curr. Opin. Chem. Biol. 2003, 7, 586–591. [Google Scholar] [CrossRef] [PubMed]
  14. Geske, G.D.; Wezeman, R.J.; Siegel, A.P.; Blackwell, H.E. Small Molecule Inhibitors of Bacterial Quorum Sensing and Biofilm Formation. J. Am. Chem. Soc. 2005, 127, 12762–12763. [Google Scholar] [CrossRef] [PubMed]
  15. Rasmussen, T.B.; Givskov, M. Quorum-sensing inhibitors as anti-pathogenic drugs. Int. J. Med. Microbiol. 2006, 296, 149–161. [Google Scholar] [CrossRef] [PubMed]
  16. Geske, G.D.; O’Neill, J.C.; Miller, D.M.; Mattmann, M.E.; Blackwell, H.E. Modulation of Bacterial Quorum Sensing with Synthetic Ligands: Systematic Evaluation of N-Acylated Homoserine Lactones in Multiple Species and New Insights into Their Mechanisms of Action. J. Am. Chem. Soc. 2007, 129, 13613–13625. [Google Scholar] [CrossRef] [PubMed]
  17. Clatworthy, A.E.; Pierson, E.; Hung, D.T. Targeting virulence: A new paradigm for antimicrobial therapy. Nat. Chem. Biol. 2007, 3, 541–548. [Google Scholar] [CrossRef] [PubMed]
  18. Meijler, M.M.; Hom, L.G.; Kaufmann, G.F.; McKenzie, K.M.; Sun, C.; Moss, J.A.; Matsushita, M.; Janda, K.D. Synthesis and Biological Validation of a Ubiquitous Quorum-Sensing Molecule. Angew. Chem. Int. Ed. 2004, 43, 2106–2108. [Google Scholar] [CrossRef] [PubMed]
  19. Globisch, D.; Lowery, C.A.; McCague, K.C.; Janda, K.D. Uncharacterized 4,5-Dihydroxy-2,3-Pentanedione (DPD) Molecules Revealed ThrougH-NMR Spectroscopy: Implications for a Greater Signaling Diversity in Bacterial Species. Angew. Chem. Int. Ed. 2012, 51, 4204–4208. [Google Scholar] [CrossRef] [PubMed]
  20. Chen, X.; Schauder, S.; Potier, N.; Van Dorsselaer, A.; Pelczer, I.; Bassler, B.L.; Hughson, F.M. Structural identification of a bacterial quorum-sensing signal containing boron. Nature 2002, 415, 545–549. [Google Scholar] [CrossRef] [PubMed]
  21. Miller, S.T.; Xavier, K.B.; Campagna, S.R.; Taga, M.E.; Semmelhack, M.F.; Bassler, B.L.; Hughson, F.M. Salmonella typhimurium recognizes a chemically distinct form of the bacterial quorum-sensing signal AI-2. Mol. Cell 2004, 15, 677–687. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, J.; Hixon, M.S.; Globisch, D.; Kaufmann, G.F.; Janda, K.D. Mechanistic Insights into the LsrK Kinase Required for Autoinducer-2 Quorum Sensing Activation. J. Am. Chem. Soc. 2013, 135, 7827–7830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ha, J.-H.; Eo, Y.; Grishaev, A.; Guo, M.; Smith, J.A.I.; Sintim, H.O.; Kim, E.-H.; Cheong, H.-K.; Bentley, W.E.; Ryu, K.-S. Crystal Structures of the LsrR Proteins Complexed with Phospho-AI-2 and Two Signal-Interrupting Analogues Reveal Distinct Mechanisms for Ligand Recognition. J. Am. Chem. Soc. 2013, 135, 15526–15535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Penesyan, A.; Gillings, M.; Paulsen, I.T. Antibiotic discovery: Combatting bacterial resistance in cells and in biofilm communities. Mol. Basel Switz. 2015, 20, 5286–5298. [Google Scholar] [CrossRef] [PubMed]
  25. Kociolek, M. Quorum-Sensing Inhibitors and Biofilms. Anti-Infect. Agents Med. Chem. 2009, 8, 315–326. [Google Scholar] [CrossRef]
  26. Brackman, G.; Coenye, T. Quorum sensing inhibitors as anti-biofilm agents. Curr. Pharm. Des. 2015, 21, 5–11. [Google Scholar] [CrossRef] [PubMed]
  27. Xavier, K.B.; Bassler, B.L. Regulation of uptake and processing of the quorum-sensing autoinducer AI-2 in Escherichia coli. J. Bacteriol. 2005, 187, 238–248. [Google Scholar] [CrossRef] [PubMed]
  28. Semmelhack, M.F.; Campagna, S.R.; Federle, M.J.; Bassler, B.L. An Expeditious Synthesis of DPD and Boron Binding Studies. Org. Lett. 2005, 7, 569–572. [Google Scholar] [CrossRef] [PubMed]
  29. De Keersmaecker, S.C.J.; Varszegi, C.; van Boxel, N.; Habel, L.W.; Metzger, K.; Daniels, R.; Marchal, K.; De Vos, D.; Vanderleyden, J. Chemical synthesis of (S)-4,5-dihydroxy-2,3-pentanedione, a bacterial signal molecule precursor, and validation of its activity in Salmonella typhimurium. J. Biol. Chem. 2005, 280, 19563–19568. [Google Scholar] [CrossRef] [PubMed]
  30. Lowery, C.A.; Park, J.; Kaufmann, G.F.; Janda, K.D. An Unexpected Switch in the Modulation of AI-2-Based Quorum Sensing Discovered through Synthetic 4,5-Dihydroxy-2,3-pentanedione Analogues. J. Am. Chem. Soc. 2008, 130, 9200–9201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Kadirvel, M.; Stimpson, W.T.; Moumene-Afifi, S.; Arsic, B.; Glynn, N.; Halliday, N.; Williams, P.; Gilbert, P.; McBain, A.J.; Freeman, S.; et al. Synthesis and bioluminescence-inducing properties of autoinducer (S)-4,5-dihydroxypentane-2,3-dione and its enantiomer. Bioorg. Med. Chem. Lett. 2010, 20, 2625–2628. [Google Scholar] [CrossRef] [PubMed]
  32. Ascenso, O.S.; Marques, J.C.; Santos, A.R.; Xavier, K.B.; Rita Ventura, M.; Maycock, C.D. An efficient synthesis of the precursor of AI-2, the signalling molecule for inter-species quorum sensing. Bioorg. Med. Chem. 2011, 19, 1236–1241. [Google Scholar] [CrossRef] [PubMed]
  33. Guo, M.; Gamby, S.; Zheng, Y.; Sintim, H. Small Molecule Inhibitors of AI-2 Signaling in Bacteria: State-of-the-Art and Future Perspectives for Anti-Quorum Sensing Agents. Int. J. Mol. Sci. 2013, 14, 17694–17728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Frezza, M.; Soulère, L.; Queneau, Y.; Doutheau, A. A Baylis–Hillman/ozonolysis route towards (±) 4,5-dihydroxy-2,3-pentanedione (DPD) and analogues. Tetrahedron Lett. 2005, 46, 6495–6498. [Google Scholar] [CrossRef]
  35. Smith, J.A.I.; Wang, J.; Nguyen-Mau, S.-M.; Lee, V.; Sintim, H.O. Biological screening of a diverse set of AI-2 analogues in Vibrio harveyi suggests that receptors which are involved in synergistic agonism of AI-2 and analogues are promiscuous. Chem. Commun. 2009, 45, 7033–7035. [Google Scholar] [CrossRef] [PubMed]
  36. Rossi, D.; Tarantino, M.; Rossino, G.; Rui, M.; Juza, M.; Collina, S. Approaches for multi-gram scale isolation of enantiomers for drug discovery. Expert Opin. Drug Discov. 2017, 12, 1253–1269. [Google Scholar] [CrossRef] [PubMed]
  37. Paterson, I.; Delgado, O.; Florence, G.J.; Lyothier, I.; O’Brien, M.; Scott, J.P.; Sereinig, N. A second-generation total synthesis of (+)-discodermolide: The development of a practical route using solely substrate-based stereocontrol. J. Org. Chem. 2005, 70, 150–160. [Google Scholar] [CrossRef] [PubMed]
  38. Peyton, L.R.; Gallagher, S.; Hashemzadeh, M. Triazole antifungals: A review. Drugs Today Barc. Spain 1998 2015, 51, 705–718. [Google Scholar] [CrossRef]
  39. Lal, K.; Yadav, P. Recent Advancements in 1,4-Disubstituted 1H-1,2,3-Triazoles as Potential Anticancer Agents. Anticancer Agents Med. Chem. 2018, 18, 21–37. [Google Scholar] [CrossRef] [PubMed]
  40. Zhou, C.-H.; Wang, Y. Recent Researches in Triazole Compounds as Medicinal Drugs. Curr. Med. Chem. 2012, 19, 239–280. [Google Scholar] [CrossRef] [PubMed]
  41. Haider, S.; Alam, M.S.; Hamid, H. 1,2,3-Triazoles: Scaffold with medicinal significance. Inflamm. Cell Signal. 2014, 1. [Google Scholar] [CrossRef]
  42. Keri, R.S.; Patil, S.A.; Budagumpi, S.; Nagaraja, B.M. Triazole: A Promising Antitubercular Agent. Chem. Biol. Drug Des. 2015, 86, 410–423. [Google Scholar] [CrossRef] [PubMed]
  43. Kharb, R.; Shahar Yar, M.; Sharma, P.C. Recent advances and future perspectives of triazole analogs as promising antiviral agents. Mini Rev. Med. Chem. 2011, 11, 84–96. [Google Scholar] [CrossRef] [PubMed]
  44. MacDonald, J.P.; Badillo, J.J.; Arevalo, G.E.; Silva-García, A.; Franz, A.K. Catalytic Stereoselective Synthesis of Diverse Oxindoles and Spirooxindoles from Isatins. ACS Comb. Sci. 2012, 14, 285–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Nolte, C.; Mayer, P.; Straub, B.F. Isolation of a copper(I) triazolide: A “click” intermediate. Angew. Chem. Int. Ed. 2007, 46, 2101–2103. [Google Scholar] [CrossRef] [PubMed]
  46. Shao, C.; Wang, X.; Xu, J.; Zhao, J.; Zhang, Q.; Hu, Y. Carboxylic Acid-Promoted Copper(I)-Catalyzed Azide−Alkyne Cycloaddition. J. Org. Chem. 2010, 75, 7002–7005. [Google Scholar] [CrossRef] [PubMed]
  47. Shao, C.; Cheng, G.; Su, D.; Xu, J.; Wang, X.; Hu, Y. Copper(I) Acetate: A Structurally Simple but Highly Efficient Dinuclear Catalyst for Copper-Catalyzed Azide-Alkyne Cycloaddition. Adv. Synth. Catal. 2010, 352, 1587–1592. [Google Scholar] [CrossRef]
  48. Shao, C.; Wang, X.; Zhang, Q.; Luo, S.; Zhao, J.; Hu, Y. Acid–Base Jointly Promoted Copper(I)-Catalyzed Azide–Alkyne Cycloaddition. J. Org. Chem. 2011, 76, 6832–6836. [Google Scholar] [CrossRef] [PubMed]
  49. Himo, F.; Lovell, T.; Hilgraf, R.; Rostovtsev, V.V.; Noodleman, L.; Sharpless, K.B.; Fokin, V.V. Copper(I)-Catalyzed Synthesis of Azoles. DFT Study Predicts Unprecedented Reactivity and Intermediates. J. Am. Chem. Soc. 2005, 127, 210–216. [Google Scholar] [CrossRef] [PubMed]
  50. Boren, B.C.; Narayan, S.; Rasmussen, L.K.; Zhang, L.; Zhao, H.; Lin, Z.; Jia, G.; Fokin, V.V. Ruthenium-catalyzed azide-alkyne cycloaddition: Scope and mechanism. J. Am. Chem. Soc. 2008, 130, 8923–8930. [Google Scholar] [CrossRef] [PubMed]
  51. Cecchi, L.; De Sarlo, F.; Machetti, F. 1,4-Diazabicyclo[2.2.2]octane (DABCO) as an Efficient Reagent for the Synthesis of Isoxazole Derivatives from Primary Nitro Compounds and Dipolarophiles: The Role of the Base. Eur. J. Org. Chem. 2006, 2006, 4852–4860. [Google Scholar] [CrossRef]
  52. Chand, P.; Kotian, P.L.; Dehghani, A.; El-Kattan, Y.; Lin, T.-H.; Hutchison, T.L.; Babu, Y.S.; Bantia, S.; Elliott, A.J.; Montgomery, J.A. Systematic Structure-Based Design and Stereoselective Synthesis of Novel Multisubstituted Cyclopentane Derivatives with Potent Antiinfluenza Activity. J. Med. Chem. 2001, 44, 4379–4392. [Google Scholar] [CrossRef] [PubMed]
  53. Jones, P.; Atack, J.R.; Braun, M.P.; Cato, B.P.; Chambers, M.S.; O’Connor, D.; Cook, S.M.; Hobbs, S.C.; Maxey, R.; Szekeres, H.J.; et al. Pharmacokinetics and metabolism studies on (3-tert-butyl-7-(5-methylisoxazol-3-yl)-2-(1-methyl-1H-1,2,4-triazol-5-ylmethoxy) pyrazolo[1,5-d][1,2,4]triazine, a functionally selective GABAA α5 inverse agonist for cognitive dysfunction. Bioorg. Med. Chem. Lett. 2006, 16, 872–875. [Google Scholar] [CrossRef] [PubMed]
  54. Quan, M.L.; Liauw, A.Y.; Ellis, C.D.; Pruitt, J.R.; Carini, D.J.; Bostrom, L.L.; Huang, P.P.; Harrison, K.; Knabb, R.M.; Thoolen, M.J.; et al. Design and Synthesis of Isoxazoline Derivatives as Factor Xa Inhibitors. 1. J. Med. Chem. 1999, 42, 2752–2759. [Google Scholar] [CrossRef] [PubMed]
  55. Mohamed, Y.A.M.; Inagaki, F.; Takahashi, R.; Mukai, C. A new procedure for the preparation of 2-vinylindoles and their [4+2] cycloaddition reaction. Tetrahedron 2011, 67, 5133–5141. [Google Scholar] [CrossRef]
  56. Aponick, A.; Li, C.-Y.; Malinge, J.; Marques, E.F. An Extremely Facile Synthesis of Furans, Pyrroles, and Thiophenes by the Dehydrative Cyclization of Propargyl Alcohols. Org. Lett. 2009, 11, 4624–4627. [Google Scholar] [CrossRef] [PubMed]
  57. Gamby, S.; Roy, V.; Guo, M.; Smith, J.A.I.; Wang, J.; Stewart, J.E.; Wang, X.; Bentley, W.E.; Sintim, H.O. Altering the communication networks of multispecies microbial systems using a diverse toolbox of AI-2 analogues. ACS Chem. Biol. 2012, 7, 1023–1030. [Google Scholar] [CrossRef] [PubMed]
  58. Morriello, G.; Devita, R.; Mills, S.; Moyes, C.; Lin, P. Anti-Hypercholesterolemic Compounds 2010. U.S. Patent 20070078098A1, 5 April 2007. [Google Scholar]
  59. Mames, A.; Stecko, S.; Mikołajczyk, P.; Soluch, M.; Furman, B.; Chmielewski, M. Direct, Catalytic Synthesis of Carbapenams via Cycloaddition/Rearrangement Cascade Reaction: Unexpected Acetylenes’ Structure Effect. J. Org. Chem. 2010, 75, 7580–7587. [Google Scholar] [CrossRef] [PubMed]
  60. Jen, T.; Mendelsohn, B.A.; Ciufolini, M.A. Oxidation of α-Oxo-Oximes to Nitrile Oxides with Hypervalent Iodine Reagents. J. Org. Chem. 2011, 76, 728–731. [Google Scholar] [CrossRef] [PubMed]
  61. Xavier, K.B.; Miller, S.T.; Lu, W.; Kim, J.H.; Rabinowitz, J.; Pelczer, I.; Semmelhack, M.F.; Bassler, B.L. Phosphorylation and Processing of the Quorum-Sensing Molecule Autoinducer-2 in Enteric Bacteria. ACS Chem. Biol. 2007, 2, 128–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Tsuchikama, K.; Zhu, J.; Lowery, C.A.; Kaufmann, G.F.; Janda, K.D. C4-Alkoxy-HPD: A Potent Class of Synthetic Modulators Surpassing Nature in AI-2 Quorum Sensing. J. Am. Chem. Soc. 2012, 134, 13562–13564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Roy, V.; Meyer, M.T.; Smith, J.A.I.; Gamby, S.; Sintim, H.O.; Ghodssi, R.; Bentley, W.E. AI-2 analogs and antibiotics: A synergistic approach to reduce bacterial biofilms. Appl. Microbiol. Biotechnol. 2013, 97, 2627–2638. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds are all available from the authors.
Figure 1. (S)-DPD in an aqueous medium: all species in equilibrium. (2S,4S)-2,4-dihydroxy-2-methyldihydrofuran-3-one (S-DHMF); (S)-4,5-dihydroxy-2,3-pentanedione (DPD); (2R,4S)-2,4-dihydroxy-2-methyldihydrofuran-3-one (R-DHMF); (2S,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (S THMF); (S)-3,3,4,5-tetrahydroxy-2-pentanone (S-THP); (2R,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (R-THMF); (2S,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuranborate (S-THMF-borate); (S)-3,3,4,5-tetrahydroxy-2-pentanone-5-phosphate (P-DPD).
Figure 1. (S)-DPD in an aqueous medium: all species in equilibrium. (2S,4S)-2,4-dihydroxy-2-methyldihydrofuran-3-one (S-DHMF); (S)-4,5-dihydroxy-2,3-pentanedione (DPD); (2R,4S)-2,4-dihydroxy-2-methyldihydrofuran-3-one (R-DHMF); (2S,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (S THMF); (S)-3,3,4,5-tetrahydroxy-2-pentanone (S-THP); (2R,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (R-THMF); (2S,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuranborate (S-THMF-borate); (S)-3,3,4,5-tetrahydroxy-2-pentanone-5-phosphate (P-DPD).
Molecules 23 02545 g001
Figure 2. The DPD-related compounds presented in this work.
Figure 2. The DPD-related compounds presented in this work.
Molecules 23 02545 g002
Scheme 1. The synthesis of racemic DPD, Ph-DPD, quinoxaline-DPD and quinoxaline-Ph-DPD. Reagents and conditions: (a) 1-propynylmagnesium bromide (0.5 M in THF, 1.3 eq), THF, 0 °C to rt, 3 h; (a′) phenylethynylmagnesium bromide (1.0 M in THF, 1.3 eq), THF, 0 °C to rt, 3 h; (b) TBDMSCl (1.2 eq), NaH (2.0 eq), THF, rt, 3 h; (b′) TMSCl (1.2 eq), NaH (2.0 eq), THF, rt, 3 h; (c) NaIO4 (4.4 eq), Ru2O·H2O (2.5% mol), CHCl3/ACN/H2O (1:1:1), rt, 1 h; (d) see Table S1; (e) Dowex50WX8 100–200 mesh, MeOH, rt, overnight; (f) cyclohexanone dimethyl ketal (3.0 eq), p-TSA (cat.), rt, overnight; (g) Dowex50WX8 100–200 mesh, D2O (10 mM), rt, overnight; (h) o-phenylendiamine (2.0 eq), rt, overnight.
Scheme 1. The synthesis of racemic DPD, Ph-DPD, quinoxaline-DPD and quinoxaline-Ph-DPD. Reagents and conditions: (a) 1-propynylmagnesium bromide (0.5 M in THF, 1.3 eq), THF, 0 °C to rt, 3 h; (a′) phenylethynylmagnesium bromide (1.0 M in THF, 1.3 eq), THF, 0 °C to rt, 3 h; (b) TBDMSCl (1.2 eq), NaH (2.0 eq), THF, rt, 3 h; (b′) TMSCl (1.2 eq), NaH (2.0 eq), THF, rt, 3 h; (c) NaIO4 (4.4 eq), Ru2O·H2O (2.5% mol), CHCl3/ACN/H2O (1:1:1), rt, 1 h; (d) see Table S1; (e) Dowex50WX8 100–200 mesh, MeOH, rt, overnight; (f) cyclohexanone dimethyl ketal (3.0 eq), p-TSA (cat.), rt, overnight; (g) Dowex50WX8 100–200 mesh, D2O (10 mM), rt, overnight; (h) o-phenylendiamine (2.0 eq), rt, overnight.
Molecules 23 02545 sch001
Scheme 2. The synthesis of intermediates 15 and 16. Reagents and conditions: (a) ethynylmagnesium bromide (0.5 M in THF, 1.3 eq), THF, 0 °C to rt, 3h; (b) Dowex50WX8 100–200 mesh, MeOH, rt, overnight; (c) cyclohexanone dimethyl ketal (3.0 eq), p-TSA (cat.), rt, overnight.
Scheme 2. The synthesis of intermediates 15 and 16. Reagents and conditions: (a) ethynylmagnesium bromide (0.5 M in THF, 1.3 eq), THF, 0 °C to rt, 3h; (b) Dowex50WX8 100–200 mesh, MeOH, rt, overnight; (c) cyclohexanone dimethyl ketal (3.0 eq), p-TSA (cat.), rt, overnight.
Molecules 23 02545 sch002
Scheme 3. The synthesis of 1,4-disubstituted 1,2,3-triazoles 18gk and 1,5-disubstituted 1,2,3-triazoles 19hk. Reagents and conditions: (a) TMSN3 (10.0 eq), CuSO4·5H2O (5% mol), Na ascorbate (0.5 eq), t-BuOH/H2O (1:1), rt, overnight; (b) 12M HCl (cat.), 1,4-dioxane, 0 °C to rt, 1–3 h; (c) R1Br (1.5 eq), K2CO3 (2.0 eq), THF, 40 °C, overnight; preparative HPLC.
Scheme 3. The synthesis of 1,4-disubstituted 1,2,3-triazoles 18gk and 1,5-disubstituted 1,2,3-triazoles 19hk. Reagents and conditions: (a) TMSN3 (10.0 eq), CuSO4·5H2O (5% mol), Na ascorbate (0.5 eq), t-BuOH/H2O (1:1), rt, overnight; (b) 12M HCl (cat.), 1,4-dioxane, 0 °C to rt, 1–3 h; (c) R1Br (1.5 eq), K2CO3 (2.0 eq), THF, 40 °C, overnight; preparative HPLC.
Molecules 23 02545 sch003
Scheme 4. The synthesis of 3,5-disubstituted isoxazoles 26lr. Reagents and conditions: (a) NH2OH·HCl (3.0 eq), Et3N (1.5 eq), H2O/EtOH (1:1), rt, 1–3 h; (b) NCS (1.0 eq), DMF, rt, 1–2 h; (c) 15 (1.0 eq), CuSO4·5H2O (5% mol), Na ascorbate (0.5 eq), KHCO3, t-BuOH/H2O (1:1), rt, overnight.
Scheme 4. The synthesis of 3,5-disubstituted isoxazoles 26lr. Reagents and conditions: (a) NH2OH·HCl (3.0 eq), Et3N (1.5 eq), H2O/EtOH (1:1), rt, 1–3 h; (b) NCS (1.0 eq), DMF, rt, 1–2 h; (c) 15 (1.0 eq), CuSO4·5H2O (5% mol), Na ascorbate (0.5 eq), KHCO3, t-BuOH/H2O (1:1), rt, overnight.
Molecules 23 02545 sch004
Scheme 5. The synthesis of 3,5-disubstituted isoxazoles (with an amide moiety at position 3) 33b, 33s–z. Reagents and conditions: (a) 27 (2.0 eq), NaOCl (40.0 eq), THF, rt, 12 h; (b) 12M HCl (cat.), 1,4-dioxane, 0 °C to rt, 1–3 h; (c) NaOH 1M (3.0 eq), THF, rt, overnight; (d) amine (2.0 eq), HOBt (2.0 eq), EDC*HCl (2.0 eq), Et3N, DCM, rt, overnight.
Scheme 5. The synthesis of 3,5-disubstituted isoxazoles (with an amide moiety at position 3) 33b, 33s–z. Reagents and conditions: (a) 27 (2.0 eq), NaOCl (40.0 eq), THF, rt, 12 h; (b) 12M HCl (cat.), 1,4-dioxane, 0 °C to rt, 1–3 h; (c) NaOH 1M (3.0 eq), THF, rt, overnight; (d) amine (2.0 eq), HOBt (2.0 eq), EDC*HCl (2.0 eq), Et3N, DCM, rt, overnight.
Molecules 23 02545 sch005
Table 1. The optimization of the conditions for the oxidation of compound 3. All the reactions were performed at room temperature.
Table 1. The optimization of the conditions for the oxidation of compound 3. All the reactions were performed at room temperature.
EntrySolventOxidant and eqTimeYield (%)
1AcetoneKMnO4/NaHCO3/MgSO4 3.8/0.6/2.0OvernightNo reaction
2AcetoneKMnO4/NaHCO3/MgSO4 3.9/0.6/4.2OvernightTraces
3CCl4/ACN (1:1)NaIO4/RuO2·H2O
2.2 eq/2.5% mol
3 hTraces
4CCl4/ACN (1:1)NaIO4/RuO2·H2O
4.4 eq/2.5% mol
3 h23
5CHCl3/ACN/H2O (1:1:1)NaIO4/RuO2·H2O
4.4 eq/2.5% mol
3 h52
Table 2. The reaction conditions to obtain 18af and 19a. All reactions were performed overnight at room temperature except for entry 4 where the mixture was heated at 60 °C.
Table 2. The reaction conditions to obtain 18af and 19a. All reactions were performed overnight at room temperature except for entry 4 where the mixture was heated at 60 °C.
Molecules 23 02545 i001
EntryR1Azide, eqSolventCatalystProductYield (%) aRef.
1(CH2)2-Ph17a, 1.1THFCuI (10% mol)
DIPEA (15% mol)
18a58[44]
2(CH2)2-Ph17a, 1.05DCMCuI (2% mol)
DIPEA (4% mol)
AcOH (cat)
18a72[48]
3(CH2)2-Ph17a, 1.0t-BuOH/H2O
(1:1)
CuSO4·5H2O (5% mol)
Na Ascorbate (0.5 eq)
18a89[49]
4(CH2)2-Ph17a, 1.01,4-dioxane(Cp*RuCl(PPh3)2) (2% mol)19a87[50]
5(CH2)-Ph17b, 1.0t-BuOH/H2O
(1:1)
CuSO4·5H2O (5% mol)
Na Ascorbate (0.5 eq)
18b60[49]
6(CH2)2-o-F-Ph17c, 1.0t-BuOH/H2O
(1:1)
CuSO4·5H2O (5% mol)
Na Ascorbate (0.5 eq)
18c62[49]
7(CH2)2-m-Pyr17d, 1.0t-BuOH/H2O
(1:1)
CuSO4·5H2O (5% mol)
Na Ascorbate (0.5 eq)
18d88[49]
8(CH2)5-CN17e, 1.0t-BuOH/H2O
(1:1)
CuSO4·5H2O (5% mol)
Na Ascorbate (0.5 eq)
18e72[49]
9(CH2)2-CyH17f, 1.0t-BuOH/H2O
(1:1)
CuSO4·5H2O (5% mol)
Na Ascorbate (0.5 eq)
18f73[49]
a Isolated yield.

Share and Cite

MDPI and ACS Style

Stotani, S.; Gatta, V.; Medda, F.; Padmanaban, M.; Karawajczyk, A.; Tammela, P.; Giordanetto, F.; Tzalis, D.; Collina, S. A Versatile Strategy for the Synthesis of 4,5-Dihydroxy-2,3-Pentanedione (DPD) and Related Compounds as Potential Modulators of Bacterial Quorum Sensing. Molecules 2018, 23, 2545. https://doi.org/10.3390/molecules23102545

AMA Style

Stotani S, Gatta V, Medda F, Padmanaban M, Karawajczyk A, Tammela P, Giordanetto F, Tzalis D, Collina S. A Versatile Strategy for the Synthesis of 4,5-Dihydroxy-2,3-Pentanedione (DPD) and Related Compounds as Potential Modulators of Bacterial Quorum Sensing. Molecules. 2018; 23(10):2545. https://doi.org/10.3390/molecules23102545

Chicago/Turabian Style

Stotani, Silvia, Viviana Gatta, Federico Medda, Mohan Padmanaban, Anna Karawajczyk, Päivi Tammela, Fabrizio Giordanetto, Dimitrios Tzalis, and Simona Collina. 2018. "A Versatile Strategy for the Synthesis of 4,5-Dihydroxy-2,3-Pentanedione (DPD) and Related Compounds as Potential Modulators of Bacterial Quorum Sensing" Molecules 23, no. 10: 2545. https://doi.org/10.3390/molecules23102545

Article Metrics

Back to TopTop