Journal of
Medicinal Plants Research

  • Abbreviation: J. Med. Plants Res.
  • Language: English
  • ISSN: 1996-0875
  • DOI: 10.5897/JMPR
  • Start Year: 2007
  • Published Articles: 3834

Full Length Research Paper

Morelloflavone and its semisynthetic derivatives as potential novel inhibitors of cysteine and serine proteases

Vanessa Silva Gontijo
  • Vanessa Silva Gontijo
  • Department of Exact Science, Laboratory of Phytochemistry and Medicinal Chemistry, Federal University of Alfenas, MG, Brazil.
  • Google Scholar
Jaqueline Pereira Januario
  • Jaqueline Pereira Januario
  • Department of Exact Science, Laboratory of Phytochemistry and Medicinal Chemistry, Federal University of Alfenas, MG, Brazil.
  • Google Scholar
Wagner Alves de Souza Judice
  • Wagner Alves de Souza Judice
  • Interdisciplinary Center of Biochemical Investigation, Mogi das Cruzes University, Mogi das Cruzes, SP, Brazil.
  • Google Scholar
Alyne Alexandrino Antunes
  • Alyne Alexandrino Antunes
  • Interdisciplinary Center of Biochemical Investigation, Mogi das Cruzes University, Mogi das Cruzes, SP, Brazil.
  • Google Scholar
Ingridy Ribeiro Cabral
  • Ingridy Ribeiro Cabral
  • Department of Exact Science, Laboratory of Phytochemistry and Medicinal Chemistry, Federal University of Alfenas, MG, Brazil.
  • Google Scholar
Diego Magno Assis
  • Diego Magno Assis
  • Department of Biophysics, Federal University of Sao Paulo, SP, Brazil.
  • Google Scholar
Maria Aparecida Juliano
  • Maria Aparecida Juliano
  • Department of Biophysics, Federal University of Sao Paulo, SP, Brazil.
  • Google Scholar
Ihosvany Camps
  • Ihosvany Camps
  • Department of Biological Sciences, Laboratory of Molecular Biology, Federal University of Alfenas, MG, Brazil.
  • Google Scholar
Marcos Jose Marques
  • Marcos Jose Marques
  • Department of Biological Sciences, Laboratory of Molecular Biology, Federal University of Alfenas, MG, Brazil.
  • Google Scholar
Claudio Viegas Junior
  • Claudio Viegas Junior
  • Department of Exact Science, Laboratory of Phytochemistry and Medicinal Chemistry, Federal University of Alfenas, MG, Brazil.
  • Google Scholar
Marcelo Henrique dos Santos*
  • Marcelo Henrique dos Santos*
  • Department of Exact Science, Laboratory of Phytochemistry and Medicinal Chemistry, Federal University of Alfenas, MG, Brazil.
  • Google Scholar


  •  Received: 09 October 2014
  •  Accepted: 11 March 2015
  •  Published: 03 April 2015

 ABSTRACT

This article reports the three biflavonoids isolated from the fruit pericarp of Garcinia brasiliensis Mart. (Clusiaceae): Morelloflavone-4```-glycoside (compound 1), (±)-Fukugiside (compound 2), and Morelloflavone (compound 3). Structural modifications by acylation and alkylation reactions were performed on the natural biflavonoid (±) morelloflavone to obtain three semisynthetic compounds: Morelloflavone-7,4`,7``,3```,4```-penta–O–acetyl (compound 4), Morelloflavone-7,4`,7``,3```,4```-penta-O-methyl (compound 5), and Morelloflavone-7,4`,7``,3```,4```-penta-O-butanoyl (compound 6). The inhibitory effects of these naturally isolated biflavonoid-type compounds and three semisynthetic derivatives on the activity of the cysteine proteases papain and cruzain, and on the serine protease trypsin were investigated. The potential inhibitory IC50 of natural bioflavonoids compounds 1, 2, and 3 were 11.0 ± 3.0, 23.0 ± 4.0, and 10.5 ± 0.3 µM, respectively, for papain;  0.86 ± 0.12, 106 ± 7, and 3.8 ± 0.1, 50 ± 2, 119.5 ± 5,  and 9.6 ± 1.0 µM, respectively, for cruzain. On the other hand, the semisynthetic biflavonoids compounds 4, 5, and 6 were more efficient in the inhibition of enzyme activity with IC50 values 0.60 ± 0.02 µM (papain) for biflavonoids compound 4, 1.64 ± 0.11 µM (trypsin) for biflavonoids compound 5, and 8.1 ± 0.6 µM (cruzain) for biflavonoids compound 6. Compound 4 is more active owing to the carbonyl group in the structure; perhaps, this modification could favor a higher nucleophilic attack by serine and cysteine proteases. However, the semisynthetic compound 5 (IC50 = 15.4 ± 0.7 µM for papain), which has no carbonyl group in structure, was less active in the inhibition. Interestingly, structure–activity relationships (SARs) were confirmed by flexible docking simulations. Likewise, the potential usefulness of natural compound 1 as an antioxidant compound was strengthened by our results concerning the antiproteolytic activity.

 

Key words: Garcinia brasiliensis, biflavonoids, proteases, antiproteolytic activity.

Abbreviation: r-CPB2.8, Recombinant cysteine protease type B; EOAc, ethyl acetate extract; MeOH, methanol; Z-Phe-Arg-MCA, carbobenzoxicarbonil-Phe-Arg-7-amino-4-methylcoumarin; DTT, dithiothreitol; SARs, structure–activity relationships; FT-NMR, Fourier transform nuclear magnetic resonance; Me4Si, trimethylsilane; TLCK, N-alpha-tosyl-L-lysinyl-chloromethylketone; E-64, 1-[[(Ltrans-epoxysuccinyl)-L-leucyl]amino]-4-guanidinobutane.

 INTRODUCTION

The Guttiferae family, also known as Clusiaceae, belongs to the angiosperm phylogeny group and is characterized by the conspicuous presence of latex in most of their species. This family consists of 47 genera such as Vismia, Garcinia, Clusia, Cratoxylum, Harungana, Mesua, Hypericum and Kielmeyera, and more than 1000 species that are grouped in six subfamilies widely spread all over Brazil (Piccinelli et al., 2005; Derogis et al., 2008). The Garcinia brasiliensis species is native to the Amazon and is cultivated throughout Brazilian territory where it is known as “bacupari.” It is a tree of medium size that blooms from August to September. It has yellow fruit with a white and edible mucilaginous pulp used in folk medicine as a wound healing agent and for peptic ulcer, urinary, and tumor disease treatments (Corrêa, 1978; Santa-Cecília et al., 2011).

According to the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology, proteases are classified in subgroup 4 of group 3 (hydrolases). In this system, enzymes are given an Enzyme Commission number (EC). However, proteases do not comply easily with the general system of enzyme nomenclature due to their broad structural and functional diversity. Currently, proteases are classified on the basis of three major criteria: (i) type of reaction catalyzed, (ii) chemical nature of the catalytic site, and (iii) evolutionary relationship with reference to structure (Rawlings and Barett, 1995). Proteases or peptidases are enzymes that catalyze reactions on peptide chains, hydrolyzing them into short fragments. Endopeptidases split the peptide bond between amino acid residues placed within the ribbon, while exopeptidases split the residues at the polymeric backbone end. These enzymes can be further joined according to the reactant groups that are present in the catalytic site, as serine (EC 3.4.21), cysteine (EC 3.4.22), aspartic acid (EC 3.4.23), metalo (EC 3.4.24) and threonine (EC 3.4.25) proteases. Papain (EC 3.4.22.2), a plant cysteine protease isolated from Carica papaya latex, preferentially cleaves peptide chains on either Arg and Lys residues or hydrophobic Phe residues (Brocklehurst, 1987; Martins et al., 2009). Trypsin (E.C. 3.4.21.4) has specificity for peptides containing Arg and Lys residues, and its catalytic triad is composed of the amino acids serine, histidine and aspartate, which is similar to other serine proteases (Martins et al., 2009; Beynon, 1989). Each amino acid of the triad has a spe-cific role in the peptide bond cleavage of the substrates. The carboxylate group of the aspartic acid forms a hy-drogen bond to the amine nitrogen atom of the histidine residue, contributing to increase the electronegativity of the   imine  nitrogen  within  the  same  heterocyclic  side chain. In this way, the free electron pair of this last imidazolyl nitrogen atom is devoted to accepting the hydrogen from the serine hydroxyl group, thereby enhancing the nucleophilic attack by this serine residue on the carbonyl carbon atom of the peptide bond that is properly oriented into the enzymatic active site.

Cruzain is a member of a large family of closely-related isoforms found in the parasite Trypanosoma cruzi. It is involved in intracellular replication and differentiation, and is essential at all stages of the parasite’s life cycle (Campetella et al., 1992; Lima et al., 1994). Since currently available therapeutics are both ineffective against the acute phase of the disease and are too toxic prolonged administration, efforts have been made to identify novel therapeutic targets, among which the proteolytic apparatus of T. cruzi is a possible candidate (McKerrow et al., 2009).

In this work, the bioactivity of three natural and three semisynthetic biflavonoid compounds were reported as part of our continuous study of G. brasiliensis (Gontijo et al., 2012a, b). The potential enzymatic inhibitory effects of the natural and semisynthetic compounds were eva-luated against papain, trypsin, and cruzain. In addition, flexible molecular docking simulations with papain, trypsin, and cruzain were performed.


 MATERIALS AND METHODS

Plant materials

 

Fruit pericarps of G. brasiliensis Mart. were collected at the campus of the Federal University of Viçosa-MG, Brazil, in February (summer) 2009. Botanical identification was performed in the Horto Botânico of the Federal University of Viçosa by Dr. João Augusto Alves Meira Neto. A voucher specimen (number VIC2604) was deposited at the Herbarium of the Federal University of Viçosa.

 

 

General considerations

 

Reagents were used as received from commercial suppliers or dried by standard methods and recondensed or distilled before use. The reactions were extracted with CH2Cl2. The organic phase was separated and dried over Na2SO4. If necessary the residue was purified by column chromatography on silica, eluting with CH2Cl2 (100%). NMR spectra were measured on a multinuclear FT-NMR spectrometer (Bruker AVANCE DRX-400 or Bruker AVANCE DPX-200 MHz and DRX-500 MHz). 1H and 13C chemical shifts are δ values and given in ppm relative to Me4Si. Coupling constants refer to H–H (1H NMR) or (13C NMR) unless stated otherwise. Molecular masses where determined by MALDI-TOF mass spectrometry (LCMS-2020EV) equipped with an ESI probe (Shimadzu), which was connected to the circuit after the UFLC SPD-20A UV detector.

Analyses were detected at 220 and 272 nm. Melting  points  were determined in sealed capillaries and uncorrected. Elemental analysis was performed using a TruSPEC (CHN-S) Analyzer by Leco Instruments, as described by Gontijo et al. (2012b). Protein Preparation Wizard protocol, Schrödinger Suite (Schrödinger Suite, 2009a).

The natural compounds were obtained from the ethyl acetate (EtOAc) extract of the fruit pericarp of G. brasiliensis; they were purified either by column chromatography on silica, eluting with hexane/ethyl acetate (60:40) or by Sephadex LH-20 eluting with MeOH. The isolated biflavonoids were identified as morelloflavone- 4```-glycoside (compound 1) (Gontijo et al., 2012a), (±)-Fukugiside (compound 2) (Elfita et al., 2009) and morelloflavone (compound 3) (Li et al., 2002).

The three semisynthetic derivatives of morelloflavone (compound 3) (Figure 1) (compounds 4, 5 and 6) were obtained from acylation and alkylation reactions.

 

 

Enzymes

 

The enzymes cruzain and rCPB2.8 (recombinant cysteine protease type B) were generously supplied by Dr. Luiz Juliano (Department of Biophysics, Federal University of São Paulo, Brazil), trypsin and papain were from Merck (Darmstadt, Germany), TLCK (N-alpha-tosyl-L-lysinyl-chloromethylketone), E-64 (1-[[(Ltrans-epoxysuccinyl)-L-leucyl]amino]-4-guanidinobutane), and fluorogenic substrate Z-FR-MCA (carbobenzyloxy-Phe-Arg-(7-amino-4-methylcoumarin) were obtained from Sigma-Aldrich Sigma (St. Louis, USA). Substrate hydrolyses were monitored in a spectrofluorometer F2500 Hitachi, and the enzymatic molar concentrations were estimated by titration according to kinetic parameters (Martins et al., 2009).

 

 

Inhibition assays

 

Compounds 1 to 6 were tested for their inhibitory potential on papain, cruzain, and trypsin by spectrofluorometric method. Enzyme inhibition was expressed as the compound concentrations causing a 50% decrease in enzyme activity (IC50 values). IC50 values were calculated by time-course, dose-response curves using inhibitors at different concentrations, and the data were analyzed in Grafit 5.0 software using Equation 1.

 

 

In Equation 1, Range is the fitted uninhibited value, y is the enzyme activity, x is the inhibitor concentration, and s is a slope factor. The equation assumes that y reduces with increasing x.

Inhibition reversibility assays for enzymes were carried out in 100 mM sodium acetate buffer, pH 5.5, 20% glycerol (to stabilize the enzymes), 5 mM EDTA, 5 mM dithiothreitol (DTT) (as cysteine protease papain and cruzain activator), and the enzyme solutions pre-activated for 5 min, at 37°C. The trypsin assays was carried out in 100 mM Trizma base HCl buffer, pH 7.5, 10% glycerol (to stabilize the enzyme). The substrate Z-FR-MCA was used as fluorogenic probe, and the hydrolysis of the substrate was monitored in a spectrofluorometer Hitachi F2500 at wavelength λEx = 360 nm (excitation) and λEm = 480 nm (emission); the reference inhibitors used were E-64 for papain and cruzain, and TLCK for trypsin. The Z-FR-MCA concentrations were 18.5 µM at cruzain (KM = 1.8 µM), 7.4 µM at papain (KM = 0.7 µM)  and  10.9 µM  at  trypsin (KM = 1.1 µM). The substrate concentrations were almost 10-fold over the KM for all enzymes in the study.

 

 

Computational docking

 

In the docking studies, the 1.8 Å resolution structure files for trypsin, papain and cruzain were downloaded from the Brookhaven Protein Data Bank (PDB code 2RA3, 1PE6 and 1F2C, respectively) (Assis et al., 2012). This structure was prepared using the Protein Preparation Wizard protocol implemented in the Schrödinger Suite (2009a). After the protein preparation step, the docking procedure was carried out using the Induced Fit Protocol (IFD) Suite (Schrödinger Suite, 2009b).

The top 20 poses for each test ligand (with regard to the GlideScore) from the initial softened-potential docking step, were retained to sample the protein plasticity using the prime program from the Schrödinger, LLC suite. In stage 2, a conformational search and minimization were carried out on residues having at least one atom within 5.0 Å of distance to the ligand for each of the 20 ligand-protein complexes. In stage 3, the 20 induced conforma-tions created for each inhibitor were ranked by the total prime energy (molecular mechanics plus solvation) of the complexes. Those within 30 kcal mol-1 of the minimum energy structure were subjected to re-docking with the ligands and scored using glide. Glide XP (extra precision) was used for all re-docking calculations. The binding affinity of each complex was established through the GlideScore, which approximates the Gibbs free energy. The more negative the value of the GlideScore, the more favorable the binding (Jacobson et al., 2004). Maestro 9.0 was used as the graphical user interface (GUI) for all of Schrödinger’s computational programs (Maestro, 2009).

 

 

Data analysis

 

Statistical analysis was performed using Grafit 5.0 software. Statistical significance was determined using one-way analysis of variance (ANOVA) followed by Tukey’s post hoc test. Values of p < 0.05 were considered statistically significant. Data are expressed as mean ± standard deviation (SD) unless otherwise specified.


 RESULTS AND DISCUSSION

The natural biflavonoid compounds 1 to 3, semisynthetic compounds 4 to 6 and the reference compounds TLCK and E64 were assayed against the cysteine proteases papain and cruzain and the serine protease trypsin to determine the potential inhibitory effects as quantified in terms of IC50 values (Table 1). The results show that all compounds tested inhibit the assayed proteases with different degrees of selectivity. The derivative compounds 4 and 5 were the most potent inhibitors to papain (IC50 = 0.60 ± 0.02 µM) and trypsin (IC50 = 1.64 ± 0.11 µM), respectively.

From the results (Table 1), compound 4 has a 35-fold higher affinity to papain than to trypsin. On the other hand, compound 5 has a 9-fold higher affinity to trypsin compared to papain. The results therefore showed that only compounds 4 and 5 exhibit a substantial degree of selectivity between the cysteine and serine proteases. However, these same compounds (4 and 5) did not efficiently inhibit the cysteine protease cruzain, where  we found the IC50 values of 537 ± 93 and 235 ± 38 µM, respectively. Comparing the inhibitory effects of both natural and semisynthetic biflavonoid compounds against the trypanossomatid cysteine proteases rCPB2.8 and rCPB3 from L. mexicana (Gontijo et al., 2012b) and cruzain from T. cruzi, it is clear that the biflavonoids stu-died here also showed species selectivity. The derivative compounds 4 and 5 are 801- and 335-fold more potent inhibitors for rCPB2.8 and rCPB3, respectively, than for cruzain (Data not shown, Table 1) (Gontijo et al., 2012b).

 

 

As shown in Table 1, the inhibitory potential (IC50) of compound 1 was approximately 123 and 4.4-fold higher than compounds 2 and 3, respectively, against trypsin. Compound 1 also displayed 2.1 and 2.4-fold higher inhibition against papain and cruzain than compound 2, respectively. Compared to compound 1, compound 3 showed a similar inhibitory effect against papain, but was more active against cruzain (5.2-fold). Excluding the natural glycoside compound 2, compounds 1 and 3 and the semisynthetic compounds 4, 5 and 6 showed IC50 values ranging from 0.6 to 15.7 µM against papain, and from 0.86 to 21.1 µM against trypsin.

As observed in Table 1, cruzain was poorly inhibited by both the natural and semisynthetic compounds. However, the biflavonoid compounds 3 (natural) and 6 (semi-synthetic) were significantly effective in the inhibition of cruzain with IC50 values of 9.6 ± 1.0 and 8.1 ± 0.6 µM, respectively. The derivative compounds 4 and 5 showed different degrees of inhibition of proteases, although the compound 4 bound preferentially to papain (IC50 = 0.60 ± 0.02 µM) making it a potential antiproteolytic. It is im-portant to highlight that the classical reference irre-versible inhibitor of serine proteases, TLCK, was 7-fold less effective at trypsin inhibition (IC50 = 11.8 ± 1.3 µM) than compound 5. On the other hand, the cysteine protease irreversible inhibitor E64 is 4614-fold and 78970-fold more effective in the inhibition of papain and cruzain, respectively. All three tested enzymes are known as trypsin-like hydrolases due to S1 specificity for both Arg and Lys basic residues, cleaving the C–N bond of substrate Z-FR-MCA between the Arg amino acid and MCA group (Assis et al., 2012).

At this point, the following interesting structure–activity relationship could be proposed: the capacity to inhibit the enzymatic activity of papain decreases according to the number of carbons in the R1=R2=R3=R4=R5 ligands (acetyl, methyl or butanoyl ligands) attached at the biflavonoid core; the absence of carbonyl group in compound 5 increases the potential inhibitory effect over trypsin by almost 13-fold.

The docking simulations of compounds with enzymes (Figure 2) showed very similar modes of binding. The amino acid sequences of papain and cruzain are 38% homologous, and their active sites are nearly identical (Assis et al., 2012; Melo et al., 2001; PDB ID, 1996). Papain and trypsin were selected because they are commercially available in large quantities, and their structures are known. The hydrophobic interactions and hydrogen bonds for compound 3 in papain, trypsin and cruzain were related previously (Martins et al., 2007, 2009; Assis et al., 2012; Sasaki et al., 1986, 1990).

According to the molecular docking simulations, papain establishes hydrophobic interactions with ring A and with rings A/B by the residues of the S2 subsite with com-pound 1 (Figure 2A). One the other hand, hydrophobic interactions are established with ring C and ring A with residues of the S2 subsite with compound 2 (Figure 2B).

Compounds 4, 5 and 6 are semisynthetic molecules derived from natural compound 3 (Figures 2C, 3C and 4C) with modifications in the five hydroxyl groups as shown in Figure 1. The acetyl, methyl and butanoyl groups   of    ring  C  from  compounds  4  (Figure 2D),  5 (Figure 2E) and 6 (Figure 2F), respectively, are establishing hydrophobic interactions with the S2 subsite of papain.

 

 

The hydroxyl groups from ring A of compound 1 are responsible for forming three hydrogen bounds (Table 2) with papain, and the OH- from the glycosyl group are forming another two H-bonds, and ring C is establishing a π-π stacking interaction between Trp69 from the S2 subsite (Figure 2A). Compound 2 are forming H-bonds between OH- from ring A with Glu66, and π-π stacking interaction between S2 subsite with rings A`/B` (Figure 2A). Hydrogen bonds (Table 2) are established between compound 4 (Figure 2D) with Trp69, Ala160, Lys156, and Gly23; compound 5 (Figure 2E) with Tyr67 and Ala160; and compound 6 (Figure 2F) with Lys156 and Trp69.

Observing the molecular modeling of trypsin (Figure 3), we verified hydrophobic interactions between compound 1 (Figure 3A) at rings A`/B` with residues of S1`, ring C` with residues of the S2 subsite, and the glycosyl group with residues of amino acids that stabilize the S1 subsite; on the other hand, compound 2 (Figure 3B) established hydrophobic interactions of the glycosyl group with amino acids of the S2 subsite and Gly193/Ser195, which are responsible for forming the oxyanion hole of trypsin.

For interaction of compounds 4, 5 and 6 (derivatives from compound 3) with the serine protease trypsin, we verified hydrophobic inter-actions of acetyl groups of compound 4 (Figure 3D) at residues of S1, S1´, and S3/S4 subsites; a methyl group of compound 5 (Figure3E) with residues of S1 subsite; and butanoyl groups of compound 6 (Figure 3F) with residues of S1 and S3/S4subsites.

 

 

 

It is important to note that some amino acid residues of trypsin can occupy two subsites, as can be observed for Gln192 (member of S1 and S1` subsites) and for residues of S3 and S4 subsites like Leu99, Trp215, Val213, Ser217.

Moreover, residues of trypsin established hydrogen bonds (Table 2) between compound 1 (Figure 3A) and residues Arg96, Trp151, Asp189, Trp215, and Gly216. On the other hand, compound 2 is establishing OH-bonds (Figure 3B) with  residues  Lys97,  Ser190,  Cys191,  Gln192, and Trp215. Furthermore, the substitution of hydroxyl groups of compound 3 at the acetyl (Figure 3Dà compound 4), the methyl (Figure 3Eà compound 5) or the butanoyl (Figure 3Fà compound 6) promotes a decrease in the number of hydrogen bonds with trypsin when compared with compounds 1, 2, and 3. Compound 6 also performs a π-π stacking interaction between His57 and ring C.

 

 

In the molecular docking of cruzain with compounds 1 (Figure 4A) and 2 (Figure 4B), the glycosyl groups are establishing hydrophobic interactions with residues of the S2 subsite (Trp26, Leu67, Met68, Ala133, and Gly160 ); and compound 4 (Figure 4D) at the acetyl group with residues of the S2 subsite; and compound 5 (Figure 4E) at the methyl group of C4` of ring C with Cys22, Gly23, Cys25, Ala136, Ser139, Trp141, and Asp158. Furthermore, it performs a π-π stacking interaction between ring A with Trp177. Compound 6 (Figure 4F) performs hydrophobic interactions at the butanoyl group of C3``` of ring C` with residues of the S2 subsite.

The Trp177 residue is widely conserved throughout the papain superfamily (McGrath 1999), and the quality of the aromatic–aromatic interactions is, in part, dependent upon the length of the inhibitor. The indole ring of this residue promotes the correct orientation of catalytic asparagine to interact with catalytic histidine by a π-NH interaction facilitating the formation of a thiolate-imidazolium ion pair during the catalysis (Brinen et al., 2000). Thus, blocking the formation of this ion pair prevents catalysis.

Cruzain protease is establishing hydrogen bonds (Table 2) with bioflavonoid compounds. Compound 1 (Figure 4A) performs H-bonds with Gln19, Gly20, Gln21, Gly66, Leu157, and Asp158; compound 2 (Figure 4B), with residues Gly66, Leu157, Asp158, and Glu205; compound 4 (Figure 4D), with Gly23 and Gln156A; compound 5 (Figure 4E), one H-bond with Gln19; and compound 6 (Figure 4F), H-bonds with Gln19 and Ala136. The increase of hydrophobicity of the compounds causes a reduction to the number of hydrogen bonds due to replacement of OH-groups by carbon chains.

 

 

 


 CONCLUSION

Three semisynthetic biflavonoid derivatives have shown potential inhibitory effects on papain and trypsin proteases; the structural modifications of compounds  are important for inhibitory activity of enzymes. The concentration of morelloflavone-7,4`,7``,3```,4```-penta-O-butanoyl (compound 6) to selectively inhibit trypsin is only 1.6-fold to that of the peptide-based classical inhibitor of this enzyme, TLCK. TLCK can potentiate such derivative as antiproteolytic drug for treatment of diseases in which this protease is involved. Therefore, the structural modifications were significantly enhanced inhibition of protease enzymes by improving the activity of natural compound 3, as observed by compounds 4 and 5 over papain and trypsin, respectively. The molecular docking models showed that the compounds mainly interact with the cysteine proteases papain and cruzain at the S2 subsite; on the other hand, the interactions occur mainly at the S1 subsite of the serine protease trypsin. However, parts of amino acid residues can participate in more than one subsite, making characterization of residue interaction with a specific subsite difficult. The structural characteristics and the hydroxyl group substitutions in the natural compound 3 increased the potency of some compounds in vitro. These semisynthetic biflavonoid derivatives are therefore potential candidates to therapeutically inhibit proteases involved in parasitic infections caused by T. cruzi.


 ACKNOWLEDGEMENTS

This work was supported by grants from Universal Project (CNPq, Brazil, #477876/2011-0) and Universal Projects (FAPEMIG, Brazil, #CEX-APQ-01072-08 and #CDS-BPD-00301/12). The authors are also grateful for the fellowships granted by CAPES, CNPq and FAMEMIG.


 ABBREVIATIONS

r-CPB2.8, Recombinant cysteine protease type B; EOAc, ethyl acetate extract; MeOH, methanol; Z-Phe-Arg-MCA, carbobenzoxicarbonil-Phe-Arg-7-amino-4-methylcoumarin; DTT, dithiothreitol; SARs, structure–activity relationships; FT-NMR, Fourier transform nuclear magnetic resonance; Me4Si, trimethylsilane; TLCK, N-alpha-tosyl-L-lysinyl-chloromethylketone; E-64, 1-[[(Ltrans-epoxysuccinyl)-L-leucyl]amino]-4-guanidinobutane.


 CONFLICT OF INTEREST

Authors have not declared any conflict of interest.



 REFERENCES

Assis MD, Gontijo VS, Pereira IO, Santos JAN, Camps I, Nagem TJ, Ellena J, Izidoro MA, Tersariol ILS, Barros NMT, Doriguetto AC, Santos MH, Juliano MA (2012). Inhibition of cysteine proteases by a natural biflavone: behavioral evaluation of fukugetin as papain and cruzain inhibitor. J. Enzyme Inhib. Med. Chem. 28(4):661-670.
Crossref
 
Beynon RJ, Bond JS (1989). Proteolytic Enzymes: A Practical Approach, Oxford University, Liverpool, UK.
 
Brinen LS, Hansell E, Cheng J, Roush WR, McKerrow JH, Fletterick RJ (2000). A target within the target: probing cruzain's P1' site to define structural determinants for the Chagas' disease protease. Structure 8(8):831-840.
Crossref
 
Brocklehurst K, Willenbrock F, Salih E (1987). In New Comprehensive Biochemistry; Neuberger A, Brocklehurst K, Eds: Elsevier: Amsterdam. 16:39-158.
 
Campetella O, Henriksson J, Aslund L, Frasch AC, Pettersson U, Cazzulo JJ (1992). The major cysteine proteinase (cruzipain) from Trypanosoma cruzi is encoded by multiple polymorphic tandemly organized genes located on different chromosomes. Mol. Biochem. Parasitol. 50:225-234.
Crossref
 
Corrêa MP (1926-1978). Dicionário das plantas úteis do Brasil e das plantas exóticas cultivadas. Imprensa Nacional: Rio de Janeiro.
 
Derogis PB, Martins FT, Souza TC, Moreira EC, Souza Filho JD, Doriguetto AC, Kamila RDS, Veloso MP, Santos MH (2008). Complete assignment of the 1H and 13C NMR spectra of garciniaphenone and keto-enol equilibrium statements for prenylated benzophenones. Mag. Res. Chem 46:278-282.
Crossref
 
Elfita E, Muharni M, Latie M, Darwati D, Widiyantoro A, Supriyatna S, Bahti HH, Dachriyanus D, Cos P, Maes L, Foubert K, Apers S, Pieters L (2009). Antiplasmodial and other constituents from four Indonesian Garcinia spp. Phytochemistry 70:907-912.
Crossref
 
Gontijo V S, Judice W AS, Codonho B, Pereira IO, Assis DM, Januário JP, Caroselli E E, Juliano MA, Dosatti ADC, Marques M J, Viegas CV, Santos MH (2012b). Leishmanicidal, antiproteolytic and antioxidant evaluation of natural biflavonoids isolated from Garcinia brasiliensis and their semisynthetic derivatives. Eur. J. Med. Chem. 58:613-623.
Crossref
 
Gontijo VS, Souza TC, Rosa IA, Soares MG, Silva MA, Vilegas W, Junior CV, Santos MH (2012a). Isolation and evaluation of the antioxidant activity of phenolic constituents of the Garcinia brasiliensis epicarp. Food Chem. 132:1230-1235.
Crossref
 
Jacobson MP, Pincus DL, Rapp CS, Day TJF, Honig B, Shaw DE, Friesner RA (2004). A hierarchical approach to all-atom loop prediction. Proteins 55:351-367.
Crossref
 
Li X, Joshi AS, Tan B, Elsohly HN, Walker LA, Zjawiony JK, Ferreira D (2002). Absolute configuration, conformation, and chiral properties of flavanone-(3→8″)- flavone biflavonoids from Rheedia acuminata.
 
Lima AP, Tessier DC, Thomas DY, Scharfstein J, Storer AC, Vernet T (1994). Identification of new cysteine protease gene isoforms in Trypanosoma cruzi. Mol. Biochem. Parasitol. 67:333-338.
Crossref
 
Maestro version 9.0 (2009). Schrödinger, LLC, New York, NY.
 
Martins FT, Assis DM, Santos MH, Camps, I, Veloso MP, Juliano MA, Alves LC, Doriguetto AC (2009). Natural polyprenylated benzophenones inhibiting cysteine and serine proteases. Eur. J. Med. Chem. 44:1230-1239.
Crossref
 
Martins FT, Cruz JWJ, Derogis PBMC, dos Santos MH, Veloso MP, Ellena J, Doriguetto AC (2007). Natural polyprenylated benzophenones inhibiting cysteine and serine proteases. J. Braz. Chem. Soc. 18:1515-1523.
Crossref
 
McGrath M, James ET, Bromme PD, Somoza JR (1998). Crystal structure of human cathepsin S. Protein Sci. 7:1294-1302.
Crossref
 
McKerrow JH, Doyle PS, Engel JC, Podust LM, Robertson SA,Ferreira R(2009). Two approaches to discovering and developing new drugs for Chagas disease. Mem. Inst. Oswaldo Cruz 104(Suppl 1):263-269.
Crossref
 
Melo RL, Alves LC, Nery ED, Juliano L, Juliano MA (2001). Synthesis and hydrolysis by cysteine and serine proteases of short internally quenched fluorogenic peptides. Anal. Bioch 293:71-77.
Crossref
 
Piccinelli AL, Cuesta-Rubio O, Chica MB, Mahmood N, Pagano B, Pavone M, Barone V, Rastrelli L (2005). Structural Revision of Clusianone (I) and 7-epi-Clusianone (II) and anti-HIV Activity of Polyisoprenylated Benzophenones. Tetrahedron 61(34):8206-8211.
Crossref
 
Rawlings ND, Barett AJ, (1995). Evolutionary families of metallopeptidases. Methods Enzymol. 248:183-228.
Crossref
 
Santa-Cecília FV, Vilela FC, Rocha CQ, Moreira EC, Dias DF, Giusti-Paiva A, Santos MH (2011). Antinociceptive and anti-inflammatory properties of 7-epiclusianone, a prenylated benzophenone from Garcinia brasiliensis. J. Ethnopharmacol. 133:467-473.
Crossref
 
Sasaki T, Kikuchi T, Fukui I, Murachi T (1986). Inativation of calpain I and calpain II by specificity-oriented tripeptidyl chloromethyl ketones. J. Biochem. 99:173-179.
Pubmed
 
Sasaki T, Kishi M, Saito M, Tanaka T, Higuchi N, Kominami E, Katunuma N, Murachi T (1990). Inhibitory effect of di- and tripeptidyl aldehydes on calpains and cathepsins. J. Enzymol. Inhib. Med. Chem. 3:195-201.
Crossref
 
Schrödinger suite (2009a). Protein preparation wizard; EPIK version 2.0, Schrödinger, LLC, New York, NY.
 
Schrödinger suite (2009b). Induced fit docking protocol; GLIDE version 5.5, schrödinger, LLC, New York.
 
 

 




          */?>