Original article
Fungal transformation of cedryl acetate and α-glucosidase inhibition assay, quantum mechanical calculations and molecular docking studies of its metabolites

https://doi.org/10.1016/j.ejmech.2013.01.036Get rights and content

Abstract

The fungal transformation of cedryl acetate (1) was investigated for the first time by using Cunninghamella elegans. The metabolites obtained include, 10β-hydroxycedryl acetate (3), 2α, 10β-dihydroxycedryl acetate (4), 2α-hydroxy-10-oxocedryl acetate (5), 3α,10β-dihydroxycedryl acetate (6), 3α,10α-dihydroxycedryl acetate (7), 10β,14α-dihydroxy cedryl acetate (8), 3β,10β-cedr-8(15)-ene-3,10-diol (9), and 3α,8β,10β -dihydroxycedrol (10). Compounds 1, 2, and 4 showed α-glucosidase inhibitory activity, whereby 1 was more potent than the standard inhibitor, acarbose, against yeast α-glucosidase. Detailed docking studies were performed on all experimentally active compounds to study the molecular interaction and binding mode in the active site of the modeled yeast α-glucosidase and human intestinal maltase glucoamylase. All active ligands were found to have greater binding affinity with the yeast α-glucosidase as compared to that of human homolog, the intestinal maltase, by an average value of approximately −1.4  kcal/mol, however, no significant difference was observed in the case of pancreatic amylase.

Graphical abstract

This paper describes fungal transformation of cedryl acetate (1) and α-glucosidase inhibitory activity of the transformed products. Detailed docking studies of all metabolites were performed using Autodock.

  1. Download : Download high-res image (95KB)
  2. Download : Download full-size image

Highlights

► Reporting seven new bio-transformed products by Cunninghamella elegans. ► Introducing new class of organic compounds as potent yeast a-glucosidase inhibitors. ► Rationalizing inhibitors’ activity at molecular level. ► Inhibitors optimization at QM level to eliminate possible error in docking study.

Introduction

It is well known that odoriferous essential oils, which are of high value in the perfume trade, contain some minor constituents which can be regarded as the real odor carriers of these products. This has been demonstrated by the studies carried out on cedrol (2), which is a major constituent of essential oils of Juniperus species [1]. Cedrol (2) possesses a rigid tricyclic sesquiterpene structure and is known to have sedative effects [2]. It is abundantly available as a fragrance or as a synthon for synthesis of related compounds. The microbiological hydroxylation of cedrol (2) by Beauveria species [3], and Cephalosporium aphidicola [4], has been shown to take place predominantly at C(3). Hydroxylations of compound 2 with Curvularia lunata [5], Rhizopus stolonifer [6], and Streptomyces bikiniensis [6] have been less regiospecific, and generally take place at C(2), C(3), C(4), C(9), C(10), and C(12).

The zygomycete fungi of the genus Cuninghamella have the ability to metabolize drugs in a manner similar to that in mammals. In the course of our work on microbial transformation of bioactive natural products [7], [8], [9], [10], [11], [12], [13], we have studied the transformation of various sesquiterpenes such as (−) ambrox, (+) sclareolide, (+) isolongifolene and (+) isolongifolol [14], [15], [16]. In the present study, we have chosen cedryl acetate (1), a tricyclic sesquiterpene isolated from the plant Psidium caudatum [17], as a model compound and investigated its metabolism by Cunninghamella elegans to obtain interesting metabolites with unique structures, as well as to assess the directing role of the acetate group at C(8).

α-Glucosidase inhibitors have been used for the control of blood glucose levels via control of the degradation of dietary disaccharides and starch. Glucosidase inhibitors are valuable in a number of applications such as their use in the study of protein glycosylation and control of viral infections through interference with the, normal glycosylation of viral coat protein [18]. In an ongoing project aimed to develop new α-glucosidase inhibitors, we screened a large number of compounds including cedryl acetate (1) and cedrol (2). Both exhibited potent α-glucosidase inhibitory activity. Based on this we carried out the microbial transformation of the most potent compound 1, and screened the transformed products for the same activity. We have reported several α-glucosidase inhibitors more potent than acarbose [19]. However, this is the first report describing the α-glucosidase inhibitory activity of cedrol (2), cedryl acetate (1), and some of the transformed products of cedryl acetate. The structures have been also optimized computationally at Hartree–Fock (HF) level of theory using valence triple-zeta plus diffuse and polarization functions (6-311++G*) basis sets for H, C, N, and O atoms to get insight into the 3D structure of these metabolites. GAMESS package [20] have been used for all quantum chemical calculations. Molecular docking studies have been also performed to delineate the ligand–protein interactions at molecular level using autodock vina program [21]. Avogadro [22], Gabedit [23], VMD [24] and Chimera [25] have been used for the structure building, analysis and visualization for our calculations.

Section snippets

Structural elucidations

Cedryl acetate (1), C17H28O2, was incubated with C. elegans for 6 days to obtain a series of minor metabolites. These metabolites can be divided into two categories. Those retaining the C(8) acetyl group (metabolites 38) and those in which deacetylation and then hydration and/or dehydration had taken place (metabolites 2, 9 and 10) (Fig. 1).

Compound 3 showed the M+ peak at m/z 280.1566 (C17H28O3) by HR-EI-MS. The 1H NMR spectrum of 3 showed a new downfield methine signal at δ(H) 3.98, while 13

General

Cedryl acetate (1) was purchased from Fluka. Column chromatography (CC): silica gel (70–230 mesh). TLC: silica gel protected PF254 plates (0.25 mm; Merck); detection by spraying with vanillin soln. M.P.: Büchi-535 apparatus; uncorrected. Optical rotations: JASCO DIP-360 digital polarimeter; in CHCl3. UV spectra: Hitachi U-3200 spectrophotometer; λmax in nm. IR spectra: Shimadzu FT-IR-8900 spectrophotmeter; in cm−1. 1H NMR and 13C NMR spectra: Bruker Avance-400 apparatus, at 500 (1H) and 125 MHz (

Acknowledgments

We are grateful to Higher Education Commission, Islamabad, Pakistan, 600-RMI/ERGS 5/3 (4/2012) and Dana Kecemerlangan 600-RMI/DANA 5/3 RIF (39/2012) (research excellence fund UiTM Malaysia) for financial support.

References (27)

  • S. Dayawansa et al.

    Auton. Nerosci.

    (2003)
    G. Maatooq et al.

    J. Nat. Prod.

    (1993)
  • J.R. Hanson et al.

    Phytochemistry

    (1993)
  • D.O. Collins et al.

    Phytochemistry

    (2001)
  • W.R. Abraham et al.

    Z. Naturforsch.

    (1987)
  • M.I. Choudhary et al.

    Helv. Chim. Acta

    (2004)
  • F. Shaheen et al.

    Eur. J. Org. Chem.

    (2006)
  • E.F. Pettersen et al.

    J. Comput. Chem.

    (2004)
  • T. Matsui et al.

    Biosci. Biotechnol. Biochem.

    (1996)
  • K.C. Wang et al.

    J. Chin. Biochem. Soc.

    (1972)
  • V. Lamare et al.

    Tetrahedron Lett.

    (1987)
  • Atta-ur-Rahman et al.

    Phytochemistry

    (1998)
  • M.I. Choudhary et al.

    Nat. Prod. Res.

    (2003)
  • M.I. Choudhary et al.

    Steroids

    (2005)
  • Cited by (0)

    View full text