Abstract
Bioassay-guided fractionation of the ethyl acetate extract from Teucrium flavum subsp. glaucum, endowed with inhibitory activity towards the HIV-1 reverse transcriptase–associated RNase H function, led to the isolation of salvigenin (1), cirsimaritin (2) and cirsiliol (3) along with the neo-clerodanes teuflavin (4) and teuflavoside (5). Acid hydrolysis of the inactive teuflavoside provided three undescribed neo-clerodanes, flavuglaucins A-C (7-9) and one known neo-clerodane (10). Among all neo-clerodanes, flavuglaucin B showed the highest inhibitory activity towards RNase H function with a IC50 value of 9.1 μM. Molecular modelling and site-directed mutagenesis analysis suggested that flavuglaucin B binds into an allosteric pocket close to RNase H catalytic site. This is the first report of clerodane diterpenoids endowed with anti-reverse transcriptase activity. Neo-clerodanes represent a valid scaffold for the development of a new class of HIV-1 RNase H inhibitors.
Keywords: HIV, RNase H, reverse transcriptase, Teucrium flavum, neo-clerodane diterpenes, flavonoids
Graphical Abstract
1. Introduction
Human Immunodeficiency Virus-1 (HIV-1) is the causative agent of the Acquired Immune Deficiency Syndrome (AIDS). Despite many countries are making progress in lowering AIDS deaths and preventing new infections, in 2019 38 million people were living with HIV1. In fact, albeit the Antiretroviral Therapy (ART) has achieved great success in HIV treatment, there is a sub-optimal treatment coverage of infected people (only the 64% according to UNAIDS 2019). The recent fail of pioneering projects of eradication of the infection2 or immunisation,3 together with raises the number of treatment failures4 due to selection and transmission of drug-resistant variants5–7, enlight the constant need of finding new drugs with innovative mechanisms of action.
Among the different steps of the viral life cycle identified as drug target, one of the most attracting and explored is represented by the HIV-1 reverse transcriptase (RT). The HIV RT is the enzyme responsible for the reverse transcription of the single-stranded RNA genome into a double-stranded DNA that can eventually integrate in the genome of the infected cell8. The RT is a multifunctional enzyme with DNA polymerase (DP) and ribonuclease H (RNase H) activities. All of the currently approved anti-HIV drugs targeting RT inhibit the DP activity. This class can be divided into Nucleoside/Nucleotide RT Inhibitors (NRTIs/NtRTIs) and Non-Nucleoside RT Inhibitors (NNRTIs)9. Although also RNase H function is essential for the reverse transcription process10, inhibitors that target this enzymatic activity have yet to enter clinical development at any stage. HIV RNase H inhibitors (RNHIs) can be divided in two groups: metal-chelating active site and allosteric inhibitors10. The majority of RNHIs are represented by compounds possessing metal coordinating functions such as diketoacids11,12, N-hydroxypyrimidinediones13, 2-hydroxyisoquinoline-1,3-diones14 and 3,4,5-trihydroxybenzoylhydrazones15. Unfortunately, these metal coordinating agents bind many host enzymes catalytic sites leading to toxicity10. By contrast, allosteric RNHIs, binding outside the active site, could be more advantageous to avoid the unspecific off-target enzymes inhibition. Among them, both synthetic16–23 and natural24,25 molecules have been found to inhibit the RNase H function by selectively targeting allosteric sites.
During the continuous search of secondary plant metabolites with antiviral activity24–26, we found that the ethyl acetate (EtOAc) extract from the leaves of Teucrium flavum subsp. glaucum showed a significant inhibitory activity towards this enzyme function, with an IC50 of 28.6 μg/mL (Table 1). In order to identify the molecules responsible of this activity, we decided to carry out a bioguided fractionation of the extract. T. flavum subsp. glaucum is an evergreen perennial shrub that grows in the calcareous mountains of Sardinia and Corsica islands from sea level up to 1000 m. Previous phytochemical studies of the aerial parts of T. flavum subsp. glaucum by Savona et al. revealed the presence of neo-clerodane diterpenes and flavonoids27. However, no biological study on the non-volatile extracts of this plant is reported.
Table 1.
Results of RT RNase H function inhibition by T. flavum subsp. glaucum fractions.
| Fractions | RNase H |
|---|---|
| IC50 (µg/mL)a | |
| EtOAc extract | 28.6 ± 3.0 |
| F1 | 25.6 ± 7.4 |
| F2 | 20.3 ± 4.0 |
| F3 | 9.9 ± 1.5 |
| F4 | >100 (100%)b |
| F5 | >100 (100%)b |
| F6 | >100 (100%)b |
Concentration capable of inhibiting 50% of enzyme activity.
Percentage of residual enzyme activity in the presence of 100 µg/mL extract.
2. Materials and methods
2.1. General experimental procedures
Optical rotations were measured in CHCl3 or MeOH at 25 °C using a Perkin-Elmer 241 polarimeter. UV spectra were recorded on a GBC Cintra 5 spectrophotometer. NMR spectra of all isolated compounds were recorded at 25 °C on Unity Inova 500NB high-resolution spectrometer (Agilent Technologies, CA, USA) operating at 500 MHz for 1H-NMR and 100 MHz for 13 C-NMR, respectively. Spectra were measured in CDCl3 and CD3OD and referenced against residual non-deuterated solvents. HRESIMS were measured on an Agilent 6520 Time of Flight (TOF) MS instrument. Column chromatography was carried out under TLC monitoring using silica gel (40–63 µm, Merck), and Sephadex LH-20 (25–100 µm, Pharmacia). For vacuum-liquid chromatography (VLC), silica gel (40– 63 µm) (Merck) was used. TLC was performed on silica gel 60 F254 or RP-18 F254 (Merck). LiChrolut RP-18 (40–63 μm) 500 mg, 3 mL (Merck) solid phase extraction (SPE) cartridges were also used. Semi-preparative HPLC was conducted by means of a Varian 920 LH instrument fitted with an autosampler module with a 1000 µL loop. The peak purities were monitored using a dual-wavelength UV detector settled at 254 and 360 nm. The columns were a 250 × 10 mm Spherisorb silica, particle size 5 µm (Waters) and a 300 × 7.5 mm Polymeric Reversed Phase (PLRP-S 100 Å), particle size 8 µm (Varian).
2.2. Plant material
The leaves of Teucrium flavum subsp. glaucum were collected in July 2003, at Orgosolo mountains (Sardinia). The plant was identified by Professor Bruno De Martis of the Department of Botanical Sciences of the University of Cagliari. A voucher specimen (No. 0309) was deposited in the Herbarium of the Department of Life and Environmental Science, Drug Sciences Section, University of Cagliari.
2.3. Extraction and isolation
Air-dried and powdered leaves of T. flavum subsp. glaucum (500 g) were ground and extracted with n-hexane (3 L) by percolation at room temperature to give 22 g dried extract. The remaining plant material was then extracted with EtOAc (2.5 L), giving 117 g dried extract. The extracts were subsequently stored at −20 °C. A sample of the EtOAc extract was tested in the RT RNase H inhibition assay in March 2019 and then phytochemically investigated.
An aliquot (17 g) of the EtOAc extract was subjected to Vacuum Liquid Chromatography (VLC) (silica gel, 90 g, 40–63 µm) using a step gradient of n-hexane/ethyl acetate (7.5: 2.5–0: 10, 500 mL each) to yield 45 fractions. Based on the TLC similarities, identical fractions were combined to give a total of six fractions (F1–F6). Fraction F1 (320 mg) was chromatographed by column chromatography (CC) over Sephadex LH-20 (MeOH) to give compound 1 (10.5 mg). Fraction F2 (900 mg) was subjected to CC over silica gel using DCM/MeOH (9.75: 0.25) as eluent, to furnish three subfractions (F2.1–F2.3). F2.1 was further purified over Sephadex LH-20 (MeOH) to give compound 1 (2.4 mg). F2.2 (55 mg) was chromatographed by Sephadex LH-20 (MeOH) to give a mixture of two compounds (12 mg) that was purified by PLRP HPLC using ACN/H2O as eluents (7: 3, flow 2.5 mL/min) to furnish compound 1 (1.5 mg, tR 12.1 min) and compound 2 (1.9 mg, tR 8.9 min) . F2.3 (120 mg) was purified using Sephadex LH-20 (MeOH) to give compound 2 (4.4 mg) and a yellow solid (86.8 mg). The obtained solid was purified further by CC over silica gel using DCM/MeOH (9.75: 0.25) as eluent, to give compound 4 (11.3 mg). An aliquot of fraction F3 (200 mg) was chromatographed by CC over Sephadex LH-20 using MeOH as eluent, to furnish an impure compound (16 mg) that was further purified by RP-18 SPE using ACN/H2O (5: 5) as eluent, to give compound 3 (3.7 mg). F4 (6.5 g) was subjected to VLC (silica gel, 90 g, 40–63 µm) using a step gradient of DCM/MeOH (9.5: 0.5–8: 2, 500 mL each) to yield 4 subfractions (F4.1–F4.4). F4.3 (4.4 g) was subjected to CC over Sephadex LH-20 using MeOH as eluent giving compound 5 (2.3 g).
2.4. Hydrolysis of teuflavoside (5)
Hydrolysis was performed by reacting 800 mg of teuflavoside (5) with 5.6 mL of 2 N H2SO4 in 95 mL of water for 20 min at 95 °C. After cooling, the solution was diluted in water and then extracted with ethyl acetate in a separatory funnel. The organic layer was evaporated by vacuum and the residue (550 mg) was chromatographed over silica gel, using DCM/EtOAc (9.75: 0.25) as eluent giving compound 6 (28.2 mg), 7 (25.1 mg) and two subfractions (FA and FB). FA fraction (28.9 mg) was purified by RP-HPLC, using ACN/H2O (6: 4, flow 2.5 mL/min), to give compound 9 (1.4 mg, tR = 10 min), and compound 6 (2.1 tR = 13 min). FB (10.5 mg) was chromatographed by RP-HPLC, using ACN/H2O (6: 4, flow 2.5 mL/min) giving compound 8 (1.5 mg, tR = 9.5 min), and compound 9 (0.6 mg, tR = 10 min).
Flavuglaucin A (6): amorphous solid; [α]20D + 48.6 (c 0.09, CH2Cl2); 1H (CDCl3, 500 MHz) and 13 C (CDCl3, 100 MHz) NMR, see Table 2; HRTOFESIMS m/z 379.1519 [M + Na]+ (calcd for C21H24O5, 379.1515).
Table 2.
1H NMR and 13 C NMR Spectroscopic Data for Compounds 6–9 (CDCl3, δ in ppm)
| Compound 6 |
Compound 7 |
Compound 8 |
||||
|---|---|---|---|---|---|---|
| position | δC, type | δH (J in Hz) | δC, type | δH (J in Hz) | δC, type | δH (J in Hz) |
| 1a 1b |
24.4, CH2 | 1.53, dddd (13.0, 13.0, 6.0, 5.5) 1.99, m |
24.6, CH2 | 1.52, dddd (13.0, 13.0, 6.0, 5.5) 2.0, m |
23.3, CH2 | 1.89, ma 2.31, ma |
| 2a | 25.9, CH2 | 2.29, ma | 25.8, CH2 | 2.30, ma | 35.5, CH2 | 2.29, ma |
| 2b 3 |
128.2, CH | 5.86, brd | 125.8, CH | 5.86, brd | 197.5, C | 2.54, ma |
| 4 | 131.6, C | 135.9, C | 131.7, C | |||
| 5 | 130.6, C | 130.7, C | 155.0, C | |||
| 6a 6b |
121.6, CH | 5.77, brd | 121.3, CH | 5.91, brd | 30.5, CH2 | 2.05, dd (13.5, 3.5) 2.99, ma |
| 7a 7b |
31.7, CH2 | 2.23, ma 2.50, ma |
31.6, CH2 | 2.26, ma 2.52, ma |
29.1, CH2 | 1.72, dd (14, 3.5) 2.13, m |
| 8 | 34.6, CH | 1.89, m | 34.6, CH | 1.89, m | 38.9, CH | 1.80, m |
| 9 | 50.4, C | 50.6, C | 54.5, C | |||
| 10 | 45.2, CH | 2.35, m | 45.3, CH | 2.34, m | 46.2, CH | 2.49, brd |
| 11 | 41.3, CH2 | 2.51, ma | 41.2, CH2 | 2.51, ma | 40.7, CH2 | 2.51, ma |
| 12 | 71.8, CH | 5.46, t (8.5) | 71.9, CH | 5.46, t (8.5) | 71.8, CH | 5.41, t (8.5) |
| 13 | 125.8, C | 125.8, C | 125.9, C | |||
| 14 | 108.1, CH | 6.40, brd | 108.1, CH | 6.40, brd | 107.8, CH | 6.40, brd |
| 15 | 144.0, CH | 7.43, t (1.5) | 144.0, CH | 7.43, t (1.5) | 144.0, CH | 7.45, t (1.5) |
| 16 | 139.3, CH | 7.45, brd | 139.4, CH | 7.45, brd | 139.3, CH | 7.47, brd |
| 17 | 17.5, CH3 | 1.0, d (6.5) | 17.6, CH3 | 1.0, d (6.5) | 17.0, CH3 | 1.03, d (6.5) |
| 18a 18b |
65.3, CH2 | 4.61, d (12.5) 4.79, d (12.5) |
64.1, CH2 | 4.24, d (12.5) 4.33, d (12.5) |
10.8, CH3 | 1.85, s |
| 19 | 176.4, C | 176.5, C | 177.0, C | |||
| COCH3 | 170.9, C | |||||
| COCH3 | 21.1, CH3 | 2.07, s | ||||
Signals were overlapped.
Flavuglaucin B (7): amorphous solid; [α]20D + 48.9 (c 0.09, CH2Cl2); 1H (CDCl3, 500 MHz) and 13 C (CDCl3, 100 MHz) NMR, see Table 2; HRTOFESIMS m/z 337.1399 [M + Na]+ (calcd for C19H22O4, 337.1392).
Flavuglaucin C (8): amorphous solid; [α]20D + 38 (c 0.05, CH2Cl2); 1H (CDCl3, 500 MHz) and 13 C (CDCl3, 100 MHz) NMR, see Table 2; HRTOFESIMS m/z 315.1593 [M + H]+ (calcd for C19H22O4, 315.1596).
Compound 9: amorphous solid; [α]20D + 260.9 (c 0.06, CH2Cl2); spectroscopic data (MS, NMR) identical to those reported in the literature27.
2.5. Molecular modelling
Flavuglaucin B (7) was docked considering the global minimum energy conformation. The ligand was built within the Maestro platform and the most stable conformation has been determined by molecular mechanics conformational analysis performed with Macromodel software version 9.228. In particular the molecule was submitted to a conformational search of 1000 steps with an energy window for saving structure of 21 kJ/mol (5.02 kcal/mol). The algorithm used was the Monte Carlo method followed by energy minimisation carried out using the MMFFs29, the GB/SA water implicit solvation model30 and the Polak-Ribier Coniugate Gradient (PRCG) method for 5000 iterations, converging on gradient with a threshold of 0.05 kJ/molÅ.
2.5.1. Protein preparation
The coordinates for reverse transcriptase enzyme were taken from the RCSB Protein Data Bank (PDB codes 1RTI)31. The protein was prepared by using the Maestro Protein Preparation Wizard. Original water molecules were removed. Also the mutated enzyme A508V-RT was generated starting from wt protein. Mutated RT was minimised considering OPLS32 force field in GB/SA30 implicit water, setting 10,000 steps interactions analysis with Polak-Ribier Coniugate Gradient (PRCG) method and a convergence criterion of 0.1 kJ/molÅ.
2.5.2. Docking experiments
The docking experiments were performed applying QM-Polarised Ligand Docking (QMPLD)33. In order to better take into account the induced fit phenomena, the most energy favoured generated complexes were fully optimised using OPLS32 united atoms force field in GB/SA implicit water30, setting 10,000 steps interactions analysis with Polak-Ribier Coniugate Gradient (PRCG) method and with a convergence criterion of 0.1 kJ/(molÅ). The resulting complexes were considered for the binding modes graphical analysis with Pymol34 and Maestro35.
2.6. Biochemistry studies
2.6.1. Expression and purification of recombinant HIV-1 RT
HIV-1 RT group M subtype B. Heterodimeric RT was expressed essentially as previously described13. Briefly, E. coli strain M15 containing the p6HRT-prot vector was grown to an optical density at 600 nm of 0.7 and induced with 1.7 mM isopropyl β-D-1-thiogalactopyranoside (IPTG) for 4 h. Protein purification was carried out with a BioLogic LP system (Biorad), using a combination of immobilised metal affinity and ion exchange chromatography. Cell pellets were resuspended in lysis buffer (50 mM sodium phosphate buffer pH 7.8, containing 0.5 mg/mL lysozyme), incubated on ice for 20 min, and after adding NaCl to a final concentration of 0.3 M, were sonicated and centrifuged at 0.30 × g for 1 h. The supernatant was loaded onto a Ni2+-NTA-Sepharose column pre-equilibrated with loading buffer (50 mM sodium phosphate buffer pH 7.8, containing 0.3 M NaCl, 10% glycerol, and 10 mM imidazole) and washed thoroughly with wash buffer (50 mM sodium phosphate buffer pH 6.0, containing 0.3 M NaCl, 10% glycerol, and 80 mM imidazole). RT was eluted with an imidazole gradient in wash buffer (0–0.5 M). Fractions were collected, protein purity was checked by SDS-PAGE and found to be higher than 90%. The 1:1 ration between the p66/p51 subunits was also verified. Enzyme-containing fractions were pooled and diluted 1:1 with 50 mM sodium phosphate buffer pH 7.0, containing 10% glycerol; and then loaded into a Hi-trap heparin HP GE (Healthcare Lifescience) pre-equilibrated with 10 column volumes of loading buffer (50 mM sodium phosphate buffer pH 7.0, containing 10% glycerol and 150 mM NaCl). The column was then washed with loading buffer and the RT was eluted with Elute Buffer 2 (50 mM sodium phosphate pH 7.0, 10% glycerol, 1 M NaCl). Fractions were collected, protein was dialysed and stored in buffer containing 50 mM Tris-HCl pH 7.0, 25 mM NaCl, 1 mM EDTA, and 50% glycerol. Catalytic activities and protein concentrations were determined. Enzyme-containing fractions were pooled and aliquots were stored at −80 °C.
2.6.2. Hiv-1 DNA polymerase-independent RNase H activity determination
HIV RT-associated RNase H activity was measured as described36 using the RNase H inhibitor RDS175911 as a control. In 100 µL reaction volume containing 50 mM Tris-HCl buffer pH 7.8, 6 mM MgCl2, 1 mM dithiothreitol (DTT), 80 mM KCl, 0.25 µM hybrid RNA/DNA 5′-GAUCUGAGCCUGGGAGCU-Fluorescin-3′ (HPLC, dry, QC: Mass Check) (available from Metabion) 5′-Dabcyl-AGCTCCCAGGCTCAGATC-3′ (HPLC, dry, QC: Mass Check), increasing concentrations of inhibitor, whose dilution were made in water, and 20 ng of wt RT according to a linear range of dose-response curve. The reaction mixture was incubated for 1 h at 37 °C, stopped by addition of EDTA and products were measured with a multilabel counter plate reader Victor 3 (Perkin Elmer model 1420–051) equipped with filters for 490/528 nm (excitation/emission wavelength).
2.6.3. Hiv-1 RNA-dependent DNA polymerase activity determination
RNA-dependent DNA polymerase (RDDP) activity was measured as described37 using the NNRTI Efavirenz as a control. In 25 µL volume containing 60 mM Tris-HCl buffer pH 8.1, 8 mM MgCl2, 60 mM KCl, 13 mM DTT, 2.5 µM poly (A)- oligo (dT), 100 µM dTTP, increasing concentrations of inhibitor, whose dilution were made in water, and 6 ng of wt RT according to a linear range of dose-response curve. After enzyme addition, the reaction mixture was incubated for 30 min at 37 °C and the stopped by addition of EDTA. Reaction products were detected by picogreen addition and measured with a multilabel counter plate reader Victor 3 (Perkin Elmer model 1420–051) equipped with filters for 502/523 nm (excitation/emission wavelength).
2.6.4. Site-directed mutagenesis
The QuikChange mutagenesis kit (Agilent Technologies Inc., Santa Clara, CA) was used to introduce amino acid substitutions into the p66 HIV-1 RT subunit coded in a p6HRT- prot plasmid by following the manufacturer’s instructions.
3. Results and discussion
3.1. Isolation and characterisation
The bioactive extract of T. flavum subsp. glaucum was subjected to vacuum liquid chromatography (VLC) with solvent mixtures of increasing polarity. 45 fractions were obtained that were combined into six main fractions (F1–F6), on the basis of their similarity in TLC. The fractions were further tested in the RT RNase H inhibition assay. The screening showed that the anti-RNase activity was concentrated in three fractions (F1–F3) and, in particular, on the third one which was able to inhibit this function with an IC50 of 9.9 µg/mL (Table 1).
Therefore, we decided to purify the three most active fractions using chromatographic techniques such as column chromatography, solid phase extraction (SPE), VLC and semi-preparative HPLC to get three flavones (1–3) and a neo-clerodane, teuflavin (4) (Figure 1). With the aim to find a structure-activity relationship (SAR), we also decided to purify the inactive fraction F4 resulting in the isolation of a further neo-clerodane, teuflavoside (5) (Figure 1). The 1H NMR spectra of F5 and F6 revealed that teuflavoside was the main secondary metabolite of these fractions and, as consequence, they were not purified. The structures of the isolated compounds were deduced from the 1 D and 2 D NMR spectra and confirmed by comparison of 1H- and 13 C NMR data with those reported in the literature27,38–40.
Figure 1.

Structures of the isolated and hydrolysed compounds.
Given the high percentage of teuflavoside in the extract (13% of the extract), we decided to verify whether the hydrolysis of teuflavoside would lead to an aglycone with inhibitory activity on the RT RNase H function. Therefore, this compound was subjected to acid hydrolysis with 2 N H2SO4 at reflux for 20 min. TLC analysis of the crude product revealed different spots and, as consequence, the mixture was separated by column chromatography and semi-preparative HPLC (RP HPLC) to give one known (9) and three new (6–8) clerodane diterpenes (Figure 1).
Compound 6 showed an ion peak at m/z 379.1519 (M + Na) (calcd 379.1515) in the HR-ESIMS (positive mode), accounting for an elemental composition of C21H24O5. The comparison of the 1H and 13 C NMR spectra of compound 6 with those of teuflavoside (5) showed that the oxymethinic proton at position 6 (δ 4.82, 1H, m) and the 2′-O-acetyl-β-D-glucopyranoside moiety of teuflavoside disappeared in the spectrum of 6 (Table 2). Furthermore, in the 1H NMR spectrum of 6 appeared two olefinic protons at 5.86 (1H, s, br) and 5.77 (1H, s, br) ppm that were not present in the spectrum of teuflavoside. These changes suggested that hydrolysis of the glycosidic moiety was accompanied by others structural modifications. From the HSQC spectrum of compound 6 it was possible to assign the respective carbons to each proton. In particular, the two olefinic protons at 5.86 and 5.77 ppm were assigned to the carbons at 128.2 and 121.6 ppm, respectively. The long-range correlations observed in the HMBC spectrum of 6 between the methylene protons at 4.61 (1Ha, d, J = 12.5 Hz) and 4.79 (1Hb, d, J = 12.5 Hz) ppm with the carbons at δ 170.9 (COCH3), 131.6 (C-4), 130.6 (C-5) and 128.2 (C-3), and the olefinic proton at 5.86 with the carbons at 130.6 (C-5), 65.3 (C-18) and 24.2 (C-1) ppm, and the olefinic proton at 5.77 with the carbons at 131.6 (C-4), 45.2 (C-10) 34.6 (C-7) ppm (Figure 2), clearly indicate the presence of two double bonds at position C-3/C-4 and C-5/C-6. ROESY experiments and analysis of scalar (3JH-H) coupling of H-8, H-12 and H-17 confirmed the same stereochemistry of teuflavoside. DQF-COSY, HSQC, HMBC, and ROESY experiments allowed the complete assignment of all signals and the identification of the structure as reported in Figure 1 (Supplementary information, figures S1–S7). It is plausible that the acid hydrolysis of teuflavoside involved the loss of the sugar, dehydration, and subsequent rearrangement of the double bonds (Supplementary information, figure S8). Compound 6 is new to the literature and was trivially named flavuglaucin A.
Figure 2.

Key HMBC correlations of compound 6.
Compound 7 showed an ion peak at m/z 337.1399 (M + Na) (calcd 337.1392) in the HR-ESIMS (positive mode), accounting for an elemental composition of C19H22O4. The 1H NMR spectrum of compound 7 (Table 2) was almost superimposable to that of compound 6, except for the absence of the acetyl group (δ 2.07, s), and the shift of 0.4 ppm to high fields of the oxymethylene protons located at C-18. The absence of the acetyl group in 8 was confirmed by HMBC experiments highlighting no cross-peaks between the oxymethylene protons at C-18 (H-a: 4.24 (d, 1H, J = 12.5 Hz); H-b: 4.33 (d, 1H, J = 12.5 Hz)) and carbonyl groups. As far as we know, compound 7 is new to the literature and was trivially named flavuglaucin B.
The HR-ESIMS of compound 8 showed a molecular ion at m/z 315.1593 (M + H)+ (calcd 315.1596). This molecular mass in combination with 1H and 13 C NMR data allowed the molecular formula to be established as C19H22O4. The analysis of the 1H NMR spectrum of compound 8 showed that the low field region is similar to that of teuflavoside but the olefinic protons at C-3/C-4 and C-5/C-6 of compounds 6 and 7, are absent in 8 (Table 2). Moreover, a methyl group at 1.85 ppm (3H, s) appeared in the spectrum of 8. The HMBC spectrum of 8 revealed that the above mentioned methyl group was correlated with a carbonyl at 197.5 ppm and two unsaturated quaternary carbons at 131.7 and 155.0 ppm (Figure 3). In the same spectrum, the carbonyl group at 197.5 was correlated to the methylene protons at 1.89 (1H-a, m) and 2.31 (1H-b, m). In the HMBC spectrum of 8, further correlations of methyl group at 1.03 (1H, d, J = 6.5 Hz) ppm with carbons at 38.9, 54.5 and 29.1 ppm together with those of the methylene proton at 1.72 (dd, J = 14, 3.5) with the methylene at 2.05 (dd, J = 13.5, 3.5) ppm observed in the COSY spectrum, allowed to identify the structure of the decalinic nucleus (Figure 3). DQF-COSY, HSQC, HMBC, and ROESY experiments allowed the complete assignment of all signals and the identification of the structure as reported in Figure 1. Compound 8 is a previously undescribed molecule and was trivially named flavuglaucin C.
Figure 3.

Selected COSY and HMBC correlations of compound 8.
The structure of compound 9 (Figure 1) was deduced from the study of 1 D and 2 D NMR spectra and MS and confirmed by comparison with the spectral data reported in the literature27.
3.2. Inhibitory effects on HIV-1 RT-associated functions and structure-activity relationships
The compounds isolated from the active fractions of T. flavum extract were evaluated for their anti-RNase H activity (Table 3) using as positive control RDS1759, a diketoacid inhibitor of the RNase H function that binds the catalytic site11. The assays revealed that the anti-RNase H activity of the extract was mainly due to the flavone cirsiliol (3) with an IC50 of 8.2 µM and to a much lower extent (IC50 = 89 µM) to the flavone cirsimaritin (2), while the flavone salvigenin completely inactive up to the concentration of 100 µM. The SAR analysis of the three flavones pointed out the importance of the catechol group to inhibit the RNase H function. Indeed, removing the hydroxyl group from C-3′ position of cirsiliol (3) lead to cirsimaritin (2) with a decrease in the activity of 10 folds. In addition, methoxylation of hydroxy group at C-4′ of cirsimaritin lead to salvigenin (1) that was completely inactive (IC50 >100 µM). Therefore, the maximum activity occurred when both hydroxyl groups are present at C-3′ and C-4′ of the B ring. The importance of the cathecol group for the inhibition of RNase H function was already observed in our previous work, comparing a series of caffeic and ferulic acid derivatives25. The natural neo-clerodanes teuflavin (4) and teuflavoside (5) resulted inactive up to the concentration of 100 µM. The lack of activity of teuflavoside was not surprising because it was purified from an inactive fraction.
Table 3.
Effect of the isolated and hydrolysed compounds on the HIV-1 RT-associated RNase H and RDDP functions
| Compounds | RnaseH IC50 (µM)a | RDDP IC50 (µM)a |
|---|---|---|
| 1 | >100 (80 %)b | NDc |
| 2 | 89 ± 7 | ND |
| 3 | 8.2 ± 0.6 | ND |
| 4 | >100 (80 %) | ND |
| 5 | >100 (74 %) | ND |
| 6 | 20.2 ± 2 | >100 (80 %) |
| 7 | 9.1 ± 0.2 | >100 (80 %) |
| 8 | 52.4 ± 0.4 | >100 (80 %) |
| 9 | 36.4 ± 0.4 | >100 (80 %) |
Compound concentration required to reduce the enzyme activity by 50%.
Percentage of residual enzyme activity in the presence of 100 µM of the compound.
ND, not done.
Among the semi-synthetic neo-clerodanes, flavuglaucin B (7) showed the greatest inhibitory activity on RNase H function with an IC50 of 9.1 µM. Flavuglaucin A (6) was about two folds less active when compared with 7. This data seemed to indicate that the alcohol function was important for interaction with the binding site of the RT-associated RNase H function. Flavuglaucin C (8) and compound 9, containing a methyl or aldehyde group at C-4 position, were about six to four folds less active than flavuglaucin B, respectively. This data confirmed the relevance of the alcoholic function to activity. However, the presence of only one double bond in the decalinic nucleus of compounds 8 and 9 change the molecular planarity and thus could further influence the interaction with RT-associated RNase H function. The neo-clerodanes 6–9 were also evaluated against the RT polymerase function (RDDP) but no inhibitory activity was observed up to the concentration of 100 µM (Table 3).
3.3. Site-directed mutagenesis experiments
Since the neo-clerodane flavuglaucin B was not able to inhibit the RDDP function and apparently it does not contain any functionality able to bind to the RNase H active site coordinating the Mg2+ cofactors, we supposed that this compound might bind an allosteric RT site. In order to verify this hypothesis, it was chosen to perform site-directed mutagenesis, determining the independent impact of several amino acid substitutions on the potency of the compound to inhibit the RNase H function. All the selected aminoacids are localised in the RNase H domain and are potentially crucial for the binding of RNase H function inhibitors. To verify a possible interaction for flavuglaucin B in the allosteric site described by Himmel et al.16, residue V108 was replaced by a phenylalanine in order to reduce the binding available space for the compound. Results showed a slight increase in IC50 when flavuglaucin B was assessed against V108F, compared with the wild type enzyme (Figure 4).
Figure 4.
Effect of flavuglaucin B (7) on the HIV-1 RT associated RNase H function of mutated RTs.
The next mutation involved the residue A502 located in the alpha helix, close to the second identified allosteric binding pocket. This pocket is located in the RNase H domain, between the RNase H active site and the primer grip region, close to the interface of subunits p66 and p51. A502 residue was replaced by a phenylalanine with the aim to provoke a shift of alpha helix that might reduce the space between the two subunits p51 and p66 and therefore hinder the entrance of the compound in the pocket. Also in this case, flavuglaucin B showed a moderate loss in potency (3.5-fold). Conversely, flavuglaucin B showed a significant loss in potency in the case of R448A (5.5-fold), R557A (6.5-fold) and, especially, A508V that totally impaired the RNase H inhibition by flavuglaucin B (IC50 >100 µM). All together these data suggested that flavuglaucin B established strong interactions within the allosteric pocket located between the RNase H active site and the primer grip region, close to the interface of subunits p66 and p51, previously investigated for other allosteric RNase H inhibitors20.
3.4. Docking experiments
To further investigate the mechanism of action of flavuglaucin B (7), we carried out QM polarised ligand (QMPL) docking experiments33. The same docking protocol was applied successfully in previous studies20,41. QMPL docking workflow combines docking with ab initio methods for ligand charges calculation within the protein environment. Subsequently, the best poses were subjected to molecular energy minimisation to consider induced-fit protein conformation change (that takes place after ligand binding) and implicit water solvation.
In agreement with site mutagenesis results, these studies suggested that flavuglaucin B binds into an allosteric pocket close to the RNase H catalytic site interacting with several residues through hydrogen bonds: Gln428, Gln509, Lys431, a cation-π with Lys424 and several hydrophobic interactions (e.g. Leu425, Leu429, Tyr532, Ala508) (Figure 5(a,b)). Hence, when bound to this site, flavuglaucin B might induce the RNase H domain to a position in which the active site might no longer be able to catalyse hydrolysis cleavage of the RNA strand in the of RNA: DNA duplex. The single point mutation of residue Ala508 to Val, in an attempt to reduce the space available for flavuglaucin B accommodation, seems to confirm this mechanism of action. The docking results into the mutated enzyme show that the compound is not able to be accommodated in the same position and it loses several important interactions (Figure 5(c,d)).
Figure 5.
3D representation of the putative binding mode obtained by docking experiments. a) RTwt- flavuglaucin B (7) c) A508V-RT- flavuglaucin B (7) complex and the relative 2D representation of the complexes stabilising interactions with the binding site residues represented with different colour depending on their chemical-physical properties: green, hydrophobic; cyan, polar; violet, positive; red, negative charged residues. Instead, magenta arrows indicate the formation of hydrogen bond between protein and ligand.
4. Conclusions
A bioguided fractionation of the T. flavum subsp. glaucum extract permitted to identify the flavone cirsiliol as the main responsible of the inhibitory activity of the RT-associated RNase H function of the extract. As far as we know, the inhibition of the HIV-1 RT-associated RNase H function by cirsiliol has not been reported in the literature. It is interesting to note that cirsiliol was also able to inhibit the HIV-1 integrase at a concentration of 12 µM42, suggesting cirsiliol as a dual inhibitor of HIV-1.
As regards the products obtained from the hydrolysis of teuflavoside, detailed NMR studies showed that the acid environment did not lead to the expected aglycone, but a series of clerodanes resulting from dehydration from position 6 and subsequent molecular rearrangement. The results seem to be in agreement with those of Savona et al.27 reporting that acid hydrolysis of the 18,2′-bis-deacetylteuflavoside did not lead to the corresponding aglycone. All semi-synthetic compounds (6–9) showed inhibitory activity on the RNase H activity and, in particular, the neo-clerodane flavuglaucin B was the most potent, with an IC50 of 9.1 µM. None of the molecules was able to inhibit the reverse transcriptase RDDP function up to a concentration of 100 µM. To the best of our knowledge, this is the first time that clerodane diterpenes have been identified as inhibitors of HIV-1 RT. Site-directed mutagenesis studies suggested that flavuglaucin B bind to the RT allosteric pocket located between the RNase H active site and the primer grip region, close to the interface of subunits p66 and p51. These results prompt us to undergo further studies to evaluate the activity of the best performing compounds on infected cells and to develop neo-clerodane derivatives with more potent anti-RT activity
Supplementary Material
Acknowledgements
We are grateful to Prof. Bruno De Martis of the University of Cagliari, for the identification of the plant species for this study. This study was financially supported by Regione Autonoma della Sardegna (RAS) (LR 07/2017, annualità 2017) grant no. RASSR17032.
Disclosure statement
No potential conflict of interest was reported by the author(s).
References
- 1.UNAIDS 2018. Global HIV Statistics. Available online : https://www.unaids.org/sites/default/files/media_asset/UNAIDS_FactSheet_en.pdf (accessed on 20 October 2020).
- 2.Henderson LJ, Reoma LB, Kovacs JA, et al. Advances toward Curing HIV-1 Infection in Tissue Reservoirs. J Virol 2019;94:1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Cohen J. Combo of two HIV vaccines fails its big test. Science 2020;367:611–2. [DOI] [PubMed] [Google Scholar]
- 4.Stella-Ascariz N, Arribas JR, Paredes R, et al. The role of HIV-1 drug-resistant minority variants in treatment failure. J Infect Dis 2017;216:S847–S850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Gupta RK, Gregson J, Parkin N, et al. HIV-1 drug resistance before initiation or re-initiation of first-line antiretroviral therapy in low-income and middle-income countries: a systematic review and meta-regression analysis. Lancet Infect Dis 2018;18:346–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Schneider A, Corona A, Spöring I, et al. Biochemical characterization of a multi-drug resistant HIV-1 subtype AG reverse transcriptase: Antagonism of AZT discrimination and excision pathways and sensitivity to RNase H inhibitors. Nucleic Acids Res 2016;44:2310–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Gregson J, Tang M, Ndembi N, et al. Global epidemiology of drug resistance after failure of WHO recommended first-line regimens for adult HIV-1 infection: A multicentre retrospective cohort study. Lancet Infect Dis 2016;16:565–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Adamson CS, Freed EO.. Novel approaches to inhibiting HIV-1 replication. Antiviral Res 2010;85:4838–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.De Clercq E. Fifty years in search of selective antiviral drugs. J Med Chem 2019;62:7322–39. [DOI] [PubMed] [Google Scholar]
- 10.Tramontano E, Corona A, Menéndez-Arias L.. Ribonuclease H, an unexploited target for antiviral intervention against HIV and hepatitis B virus. Antiviral Res 2019;171:104613. [DOI] [PubMed] [Google Scholar]
- 11.Corona A, Di Leva FS, Thierry S, et al. Identification of highly conserved residues involved in inhibition of HIV-1 RNase H function by diketo acid derivatives. Antimicrob Agents Chemother 2014;58:6101–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.(a) Corona A, Schneider A, Schweimer K, et al. Inhibition of foamy virus reverse transcriptase by human immunodeficiency virus type 1 RNase H inhibitors. Antimicrob Agents Chemother 2014;58:4086–93. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Cuzzucoli Crucitti G, Métifiot M, Pescatori L, et al. Structure-activity relationship of pyrrolyl diketo acid derivatives as dual inhibitors of HIV-1 integrase and reverse transcriptase ribonuclease H domain. J Med Chem 2015;58:1915–28. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Costi R, Métifiot M, Chung S, et al. Basic quinolinonyl diketo acid derivatives as inhibitors of HIV Integrase and their activity against RNase H function of reverse transcriptase. J Med Chem 2014;57:3223–34. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Corona A, di Leva FS, Rigogliuso G, et al. New insights into the interaction between pyrrolyl diketoacids and HIV-1 integrase active site and comparison with RNase H. Antiviral Res 2016;134:236–43. [DOI] [PubMed] [Google Scholar]; (e) Pescatori L, Métifiot M, Chung S, et al. N-substituted quinolinonyl diketo acid derivatives as HIV integrase strand transfer inhibitors and their activity against RNase H function of reverse transcriptase. J Med Chem 2015;58:4610–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Wang X, Gao P, Menéndez-Arias L, et al. Update on recent developments in small molecular HIV-1 RNase H inhibitors (2013-2016): opportunities and challenges. Curr Med Chem 2018;25:1682–702. [DOI] [PubMed] [Google Scholar]
- 14.Vernekar SKV, Liu Z, Nagy E, et al. Design, synthesis, biochemical, and antiviral evaluations of C6 benzyl and C6 biarylmethyl substituted 2-hydroxylisoquinoline-1,3-diones: dual inhibition against HIV reverse transcriptase-associated RNase H and polymerase with antiviral activities. J Med Chem 2015;58:651–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Corona A, Ballana E, Distinto S, et al. Targeting HIV-1 RNase H: N’ -(2-Hydroxy-benzylidene)-3,4,5-trihydroxybenzoylhydrazone as selective inhibitor active against NNRTIs-resistant variants. Viruses 2020;12:729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Himmel DM, Sarafianos SG, Dharmasena S, et al. HIV-1 reverse transcriptase structure with RNase H inhibitor dihydroxy benzoyl naphthyl hydrazone bound at a novel site. ACS Chem Biol 2006;1:702–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Christen MT, Menon L, Myshakina NS, et al. Structural basis of the allosteric inhibitor interaction on the HIV-1 reverse transcriptase RNase H domain. Chem Biol Drug Des 2012;80:706–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Distinto S, Esposito F, Kirchmair J, et al. Identification of HIV-1 reverse transcriptase dual inhibitors by a combined shape-, 2D-fingerprint- and pharmacophore-based virtual screening approach. Eur J Med Chem 2012;50:216–29. [DOI] [PubMed] [Google Scholar]
- 19.Distinto S, Maccioni E, Meleddu R, et al. Molecular aspects of the RT/drug interactions. Perspective of dual inhibitors. Curr Pharm Des 2013;19:1850–9. [DOI] [PubMed] [Google Scholar]
- 20.Corona A, Meleddu R, Esposito F, et al. Ribonuclease H/DNA polymerase HIV-1 reverse transcriptase dual inhibitor: Mechanistic studies on the allosteric mode of Action of Isatin-Based Compound RMNC6. PLoS One 2016;11:e0147225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Wendeler M, Lee HF, Bermingham A, et al. Vinylogous ureas as a novel class of inhibitors of reverse transcriptase-associated ribonuclease H activity. ACS Chem Biol 2008;3:635–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Chung S, Miller JT, Johnson BC, et al. Mutagenesis of human immunodeficiency virus reverse transcriptase p51 subunit defines residues contributing to vinylogous urea inhibition of ribonuclease H activity. J Biol Chem 2012;287:4066–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Massari T, Corona A, Distinto S, et al. From cycloheptathiophene-3-carboxamide to oxazinone-based derivatives as allosteric HIV-1 ribonuclease H inhibitors. J Enzyme Inhib Med Chem 2019;34:55–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Esposito F, Mandrone E, Del VC, et al. Multi-target activity of Hemidesmus indicus decoction against innovative HIV-1 drug targets and characterization of lupeol mode of action. Pathog Dis 2017;75:ftx065. [DOI] [PubMed] [Google Scholar]
- 25.Sonar VP, Corona A, Distinto S, et al. Natural product-inspired esters and amides of ferulic and caffeic acid as dual inhibitors of HIV-1 reverse transcriptase. Eur J Med Chem 2017;130:248–60. [DOI] [PubMed] [Google Scholar]
- 26.Fois B, Bianco G, Sonar VP, et al. Phenylpropenoids from Bupleurum fruticosum as anti-human rhinovirus species A selective capsid binders. J Nat Prod 2017;80:2799–806. [DOI] [PubMed] [Google Scholar]
- 27.Savona G, Piozzi F, Servettaz O, et al. A neo-clerodane glucoside and neo-clerodane diterpenoids from Teucrium flavum subsp. glaucum. Phytochemistry 1984;23:843–8. [Google Scholar]
- 28.Mohamadi F, Richards NGJ, Guida W, et al. Macromodel-an integrated software system for modeling organic and bioorganic molecules using molecular mechanics. J Comput Chem 1990;11:440–67. [Google Scholar]
- 29.Halgren T. Merck molecular force field. II. MMFF94 van der Waals and electrostatic parameters for intermolecular interactions. J Comput Chem 1996;17:520–52. [Google Scholar]
- 30.Still WC, Tempczyk A, Hawley RC, T, et al. Semianalytical treatment of solvation for molecular mechanics and dynamics. J Amer Chem Soc 1990;112:6127–9. [Google Scholar]
- 31.Ren JS, Esnouf R, Garman E, et al. High resolution structures of HIV-1 RT from four RT-inhibitor complexes. Nat Struct Biol 1995;2:293–302. [DOI] [PubMed] [Google Scholar]
- 32.Jorgensen WL, Maxwell DS, Tirado-Rives J.. Development and testing of the OPLS all-atom force field on conformational energetics and properties of organic liquids. J Am Chem Soc 1996;118:11225–36. [Google Scholar]
- 33.Schrödinger LLC , QMPolarized protocol, Schrodinger Suite New York, NY, USA.PyMOL, Molecular Graphics System Schrödinger, LLC. [Google Scholar]
- 34.PyMOL , Molecular Graphics System Schrödinger, LLC. [Google Scholar]
- 35.Schrödinger LLC, Maestro GUI, New York, NY, USA, 2013. [Google Scholar]
- 36.Messore A, Corona A, Madia VN, et al. Pyrrolyl pyrazoles as non-diketo acid inhibitors of the HIV-1 ribonuclease H function of reverse transcriptase. ACS Med Chem Lett 2020;11:798–805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Xu L, Grandi N, Del Vecchio C, et al. From the traditional Chinese medicine plant Schisandra chinensis new scaffolds effective on HIV-1 reverse transcriptase resistant to non-nucleoside inhibitors. J Microbiol 2015;53:288–93. [DOI] [PubMed] [Google Scholar]
- 38.Ouyang X, Wei L, Pan Y, et al. Antioxidant properties and chemical constituents of ethanolic extract and its fractions of Ocimum gratissimum. Med Chem Res 2013;22:1124–30. [Google Scholar]
- 39.Awad BM, Habib ES, Ibrahim AK, et al. Cytotoxic activity evaluation and molecular docking study of phenolic derivatives from Achillea fragrantissima (Forssk.) growing in Egypt. Med Chem Res 2017;26:2065–73. [Google Scholar]
- 40.Venditti A, Frezza C, Sciubba F, et al. Volatile components, polar constituents and biological activity of tansy daisy (Tanacetum macrophyllum (. Waldst. et Kit.) Schultz Bip.). Ind Crops Prod 2018;118:225–35. [Google Scholar]
- 41.Meleddu R, Cannas V, Distinto S, et al. Design, synthesis, and biological evaluation of 1,3-diarylpropenones as dual inhibitors of HIV-1 reverse transcriptase. ChemMedChem 2014;9:1869–79. [DOI] [PubMed] [Google Scholar]
- 42.Tewtrakul S, Miyashiro H, Hatrori M, et al. Inhibitory effects of flavonoids on human immunodeficiency virus type-1 integrase. Yakugaku Zasshi 2001;18:229–38. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



