Inhibition of Human Monoamine Oxidase: Biological and Molecular

Nov 9, 2016 - Naturally occurring flavonoids display a plethora of different biological activities, but emerging evidence suggests that this class of ...
1 downloads 14 Views 683KB Size
Subscriber access provided by University of Otago Library

Article

Inhibition of Human Monoamine Oxidase: Biological and Molecular Modelling Studies on Selected Natural Flavonoids Simone de Carradori, Maria Concetta Gidaro, Anel Petzer, Giosuè Costa, Paolo Guglielmi, Paola Chimenti, Stefano Alcaro, and Jacobus Petrus Petzer J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b03529 • Publication Date (Web): 09 Nov 2016 Downloaded from http://pubs.acs.org on November 14, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 26

Journal of Agricultural and Food Chemistry

1

Inhibition of Human Monoamine Oxidase: Biological and

2

Molecular Modelling Studies on Selected Natural Flavonoids

3 4

Simone Carradori†,*, Maria Concetta Gidaro‡, Anél Petzer§, Giosuè Costa‡, Paolo Guglielmi⊥,

5

Paola Chimenti⊥, Stefano Alcaro‡, Jacobus P. Petzer§

6 7 8 9 10



Department of Pharmacy, “G. d’Annunzio” University of Chieti-Pescara, Via dei Vestini 31,

66100 Chieti, Italy ‡

Dipartimento di Scienze della Salute, “Magna Graecia” University of Catanzaro, Campus

11

Universitario “S. Venuta”, Viale Europa Loc. Germaneto, 88100 Catanzaro, Italy

12

§

13

South Africa

14



15

00185 Rome, Italy

Centre of Excellence for Pharmaceutical Sciences, North-West University, Potchefstroom 2531,

Department of Drug Chemistry and Technologies, Sapienza University of Rome, P.le A.Moro 5,

16 17 18

Corresponding author

19

*

(S. C.) Tel./Fax: +39 0871 3554583. E-mail: [email protected].

ACS Paragon Plus Environment

1

Journal of Agricultural and Food Chemistry

20

Abstract

21

Naturally occurring flavonoids display a plethora of different biological activities, but emerging

22

evidence suggests that this class of compounds may also act as antidepressant agents endowed with

23

multiple mechanisms of action in the central nervous system; increasing central neurotransmission,

24

limiting the reabsorption of bioamines by synaptosomes and modulating the neuroendocrine and

25

GABAA systems. Due to their presence in foods, food-derived products and nutraceuticals, we

26

established their role and structure-activity relationships as reversible and competitive human

27

monoamine oxidase inhibitors. In addition, molecular modelling studies, which evaluated their

28

modes of MAO inhibition, are presented. These findings could provide pivotal implications in the

29

quest of novel drug-like compounds and for the establishment of harmful drug-dietary supplement

30

interactions commonly reported in the therapy with antidepressant agents.

31 32 33 34 35

Keywords: monoamine oxidase inhibitor, apigenin, taxifolin, diosmetin, naringin, eriocitrin,

36

isorhamnetin, hesperidin, quercetin-3-O-glucoside.

37

ACS Paragon Plus Environment

Page 2 of 26

Page 3 of 26

Journal of Agricultural and Food Chemistry

38

Introduction

39

Human monoamine oxidases (hMAO-A and hMAO-B; EC 1.4.3.4), are mitochondrial enzymes

40

which oxidatively deaminate monoaminergic neurotransmitters and (potentially harmful) dietary

41

monoamines. In the brain, their main role is the regulation of important monoamines such as

42

noradrenaline, dopamine, serotonin, and adrenaline. The hMAO-A isoform is present in

43

catecholaminergic neurons and degrades sterically hindered amines like serotonin, adrenaline, and

44

noradrenaline; hMAO-B is mainly abundant in the periventricular region of the hypothalamus and

45

the substantia nigra, where it metabolizes small monoamines like β-phenethylamine and

46

benzylamine. Both isoforms use dopamine and p-tyramine as common substrates.1

47

Due to their crucial physiological roles, the hMAO-A isoform is an established pharmacological

48

target for the therapy of (resistant and atypical) depression, anxiety disorders, dysthymia and major

49

depressive disorder, whereas the hMAO-B isoform is a target for the treatment of

50

neurodegenerative disorders including Parkinson’s disease and Alzheimer’s disease. Moreover, the

51

selective hMAO-B inhibitors have been shown to protect neurons against degeneration by means of

52

anti-apoptotic mechanisms, the reduction of reactive oxygen species and the stabilization of the

53

mitochondria.2

54

A plethora of molecules of natural origin have been shown to inhibit the hMAOs and, in addition,

55

different synthetic scaffolds have often been inspired by natural products such as flavonoids,

56

coumarins, purine derivatives and alkaloids.3-6 Secondary metabolites from plants are frequently

57

used by medicinal chemists in an attempt to design simple small molecules that mimic the

58

heterogeneity of scaffolds found in natural products. Moreover, there has been an interest in

59

traditional flavonoid-enriched herbal medicines because of their application for the therapy of

60

specific disorders.7-11 The exact knowledge of the molecular mechanism of action of several natural

61

products, which could be ingested daily through the diet in gram amounts, could be important to

62

avoid or limit harmful food–food and drug–food interactions.

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 26

63

Flavonoids could be classified as plant polyphenolic metabolites responsible of specific effects in

64

promoting human health and modulating physiological functions, which depend on the substitution

65

pattern and related chemical-physical properties.12-14 Indeed, these compounds are characterized by

66

a highly reactive hydroxyl moiety, which acts as a free radical scavenger by means of oxidation to

67

less reactive molecule, thus preventing ailments associated to oxidative stress. These compounds

68

also trigger important enzymes involved in mitochondrial respiration and chelate divalent metal

69

ions. Emerging evidence suggests that flavonoids may possess antidepressant activity by acting via

70

multiple mechanisms to increase monoamine neurotransmission in the brain. These mechanisms

71

include inhibition of the reabsorption of bioamines by synaptosomes and modulation of the

72

neuroendocrine and GABAA systems.15, 16

73

Recently, we analyzed the structure-activity relationships of this class of natural compounds as

74

MAO inhibitors and investigated their antidepressant-like effects in animal models.17, 18 Generally,

75

the increasing number of -OH groups on the aromatic ring of this scaffold decreased the MAO

76

inhibitory activity. Moreover, the introduction of a sugar or glucuronic acid at the C7 lowered this

77

biological activity. MAO inhibition is strongly dependent on the presence of a (p-OH-

78

substituted)phenyl at the C2, unsaturation at the C2-C3 positions of the structure, the possibility of

79

establishing hydrophobic interactions, and ring planarity. These polar plant metabolites can cross

80

the blood-brain barrier due to their cLogP values ranging from 1.5 to 3.5 and since they were shown

81

to induce biological effects directly in the CNS.17

82

Among natural flavonoid compounds, we selected the most abundant and representative derivatives

83

and evaluated them as potential hMAO inhibitors in an attempt to improve and corroborate SARs

84

for their target interaction (Figure 1). Apigenin (1) and diosmetin (2) are 2,5,7-trisubstituted

85

flavones which are known inhibitors of human and rat liver MAO.8,

86

isorhamnetin (4) possess an additional OH moiety on C3. Moreover, due to numerous reports of the

87

MAO inhibition properties of quercetin, we also studied its 3-O-glucosidic derivative, isoquercitrin

88

(5), which displayed an IC50 of 11.64 µM for the inhibition of bovine brain MAO-B.17 Lastly, 4 ACS Paragon Plus Environment

17

Taxifolin (3) and

Page 5 of 26

Journal of Agricultural and Food Chemistry

89

naringin (6), eriocitrin (7), and hesperidin (8) are flavanones characterized by different bulky di-

90

glycosidic substituents on C7. After the determination of their MAO inhibitory activity and kinetics

91

in vitro, we also performed docking simulations and molecular dynamics studies in order to

92

investigate the mechanism of action and ligand-target interactions for these widely distributed

93

flavonoids.

94

Materials and Methods

95

All commercial samples of selected flavonoids (apigenin ≥99%, diosmetin ≥98%, taxifolin ≥98%,

96

isorhamnetin ≥99%, isoquercitrin ≥98%, naringin ≥95%, eriocitrin ≥98%, hesperidin ≥97%) and

97

reference drugs (harmine 98% and safinamide mesylate salt ≥98%) were purchased by Sigma-

98

Aldrich (Milan, Italy).

99

Microsomes from insect cells containing recombinant hMAOs (5 mg/mL) were used as sources of

100

the MAO enzymes. These isoenzymes, kynuramine, (R)-deprenyl and pargyline were supplied by

101

Sigma-Aldrich (St. Louis, MO, USA). The Prism 5 software package (GraphPad Software, La Jolla,

102

CA, USA) was chosen for data analyses and for the construction of graphs. Fluorescence

103

spectrophotometry was carried out with a Varian Cary Eclipse fluorescence spectrophotometer

104

(Agilent Technologies, Santa Clara, CA, USA).

105

For the computational studies the following commercial softwares were used: Maestro, version 9.7

106

(Schrödinger, LLC, New York, NY); LigPrep, version 2.9 (Schrödinger, LLC, New York, NY);

107

Glide, version 6.2 (Schrödinger, LLC, New York, NY); Desmond Molecular Dynamics System,

108

version 3.7 (D. E. Shaw Research, New York, NY); Maestro-Desmond Interoperability Tools,

109

version 3.7 (Schrödinger, New York, NY); The PyMOL Molecular Graphics System, version

110

1.7.0.0 (Schrödinger, LLC, New York, NY).

111

Biochemistry

112

The measurement of IC50 and Ki values

113

IC50 values for the inhibition of hMAO-A and hMAO-B were determined according to the

114

published protocol.19 The experimental protocol was performed in 96-well microtiter plates (white) 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

115

to a final volume of 200 µL. The reactions contained kynuramine (50 µM), different concentrations

116

of the test inhibitors (0.003–100 µM) and DMSO (4%) as co-solvent. Potassium phosphate buffer

117

(100 mM, pH 7.4, made isotonic with KCl) was used as reaction solvent. The enzymatic reactions

118

were initiated with the addition of hMAO-A (0.0075 mg protein/mL) or hMAO-B (0.015 mg

119

protein/mL) and, after 20 min incubation at 37 ºC, were terminated with the addition of sodium

120

hydroxide (80 µL of 2 N). The MAO-generated metabolite, 4-hydroxyquinoline, was subsequently

121

quantitated by fluorescence spectrophotometry (λex = 310; λem = 400 nm). For this purpose, linear

122

calibration curves constructed with 4-hydroxyquinoline (0.047–1.56 µM) were employed. After

123

fitting the inhibition data to the one site competition model incorporated into the Prism 5 software

124

package, the IC50 values were calculated. These are expressed as the mean ± standard deviation

125

(SD) of triplicate determinations (three replicates of one concentration range). To calculate Ki

126

values for the inhibition of hMAOs, Lineweaver-Burk plots were accomplished. For each inhibitor,

127

a set of six plots were constructed using the following inhibitor concentrations: 0, 0.25, 0.5, 0.75,

128

1.0, and 1.25 x IC50, respectively. For each plot, eight different concentrations of kynuramine were

129

employed (15–250 µM). The enzymatic reactions and fluorometric measurements were conducted

130

as previously described with the exception that the concentration of both hMAO-A and hMAO-B

131

was 0.015 mg protein/mL. Moreover, these experiments were performed in a final volume of 500

132

µL and, after termination with NaOH (400 µL of 2 N), 1 mL water was added and the resulting

133

samples were quantitated by fluorescence spectrophotometry (using a 3.5 mL quartz cuvette). The

134

Ki values were estimated from replots of the slopes of the Lineweaver-Burk plots versus inhibitor

135

concentration (–Ki = x-axis intercept) as well as by global fitting of the inhibition data to the

136

Michaelis-Menten equation using the Prism 5 software package.

137

Dialysis studies

138

Dialysis of mixtures containing the hMAOs and test inhibitors were carried out using Slide-A-Lyzer

139

dialysis cassettes (Thermo Scientific, Waltham, MA, USA) with a molecular weight cut-off of 10

140

000 and a sample volume capacity of 0.5–3 mL. This protocol has been reported previously.20 The 6 ACS Paragon Plus Environment

Page 6 of 26

Page 7 of 26

Journal of Agricultural and Food Chemistry

141

hMAOs (0.03 mg protein/mL) were incubated with the test inhibitors, apigenin (1) and diosmetin

142

(2), as well as with the reference irreversible inhibitors, pargyline (IC50 = 13 µM) and (R)-deprenyl

143

(IC50 = 0.079 µM), for 15 min at 37 °C. For these experiments, the inhibitor concentrations were 4

144

× IC50 and the final volume, 0.8 mL. Control incubations were also evaluated in the absence of

145

inhibitor. The incubation mixtures were dialyzed using potassium phosphate buffer (100 mM, pH

146

7.4, containing 5% sucrose) as dialysis buffer (80 mL). The dialysis buffer was replaced with fresh

147

buffer at 3 h and 7 h after the start of dialysis. Following dialysis, the mixtures were diluted twofold

148

with the addition of kynuramine to obtain a substrate concentration equal to 50 µM and an inhibitor

149

concentration of 2 × IC50. The final volume of these reactions was 500 µL and the residual hMAO

150

activities were measured as described for the measurement of Ki values above. For comparison,

151

non-dialyzed incubation mixtures containing the hMAOs and 1 or 2 were maintained at 4 °C for 24

152

h, diluted twofold and the residual enzyme activity measured. The hMAO activities are reported as

153

the mean ± SD of triplicate determinations. The Kruskal-Wallis test with Dunn’s post hoc test was

154

used to determine the statistical differences among the means of the residual enzyme rates. A p

155

value < 0.05 is judged as being statistical significantly different. These analyses were performed by

156

the Prism 5 software package.

157

Molecular modelling

158

The X-ray structures of complexes of hMAO-A/harmine and hMAO-B/safinamide were

159

downloaded from the Protein Data Bank (PDB)21 with the respective accession codes 2Z5X22 and

160

2V5Z23. For each protein model, the “Protein Preparation Wizard” function in Maestro ver. 9.7

161

(Schrödinger LLC, New York, NY) of Suite 2014 was applied in order to add hydrogens, assign

162

partial charges, build side chains and loops with missing atoms. The co-crystallized inhibitors,

163

harmine and safinamide, were used to generate the docking grid box and were then removed prior

164

to grid generation in the next step. For each flavonoid (1-8), the 3D SDF file was imported from the

165

PubChem24 site into the Maestro GUI 9.7 of the Scrodinger Suite 2014. Both tautomeric and

166

protonated forms at pH 7.4 of each flavonoid were considered by the “LigPrep” module, ver. 2.9 7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

167

(Schrödinger). The lowest energy conformations, obtained using the OPLS-2005 force field, were

168

used as starting points for the docking simulations carried out by Glide ver. 6.2 (Schrödinger). The

169

Desmond Molecular Dynamics System, ver. 3.7 (Schrödinger) with OPLS-2005 force field, was

170

used to perform the MD simulations of 100 ns with the NPT ensemble, temperature of 10 K and

171

pressure of 1 atm. The Simulation Interactions Diagram utility of Desmond 3.7 was used to display

172

the ligand-target interactions during the MD simulations. Finally, the PyMOL molecular graphics

173

system, ver. 1.7.0.0 (Schrödinger) was used to visualize the molecules and generate all figures.

174

Results and Discussion

175

The hMAO inhibitory activities of the naturally occurring flavonoids 1-8 were evaluated using

176

recombinant hMAO-A and hMAO-B (Table 1).25 To measure MAO activity the non-specific MAO-

177

A and B substrate, kynuramine, was used. In a well-established protocol, the MAO-catalyzed

178

oxidation product of kynuramine, 4-hydroxyquinoline, was quantitated by fluorescence

179

spectrophotometry.26 By thus measuring MAO activity in the presence of several concentrations of

180

each inhibitor, sigmoidal plots (residual MAO activity vs. logarithm of inhibitor concentration)

181

were constructed from which IC50 values were extrapolated. All inhibition data were generated with

182

the substrate concentration at ~1 × Km. The IC50 values for the inhibition of hMAO-A and hMAO-B

183

by the selected flavonoids are collected in Table 1.

184

Apigenin (1), the simplest flavonoid within our series, displayed promising hMAO inhibitory

185

activity with IC50 values in the low micromolar range for the inhibition of both isoforms. However,

186

it possessed slight selectivity for hMAO-A compared to the B isoform. The 3'-hydroxy-4'-methoxy

187

derivative, diosmetin (2), inhibited the hMAOs with similar potencies compared to 1. Interestingly,

188

the isoform selectivity of 2 is reversed compared to that of 1. Hydroxylation on C3 of the flavonoid

189

moiety (4) significantly reduced hMAO-A and hMAO-B inhibition. This reduction of inhibition

190

potency may also be attributed to the 4'-hydroxy-3'-methoxy substitution pattern, which differs from

191

those of 1 and 2. hMAO inhibitory activity is completely abolished in 3, most likely due to the loss

192

of planarity of the structure (absence of the C2-C3 double bond). Flavonoid 5 was inactive (>100 8 ACS Paragon Plus Environment

Page 8 of 26

Page 9 of 26

Journal of Agricultural and Food Chemistry

193

µM) against both isoforms, whereas the di-glycosidic derivatives (6-8) exhibited relatively low

194

hMAO inhibitory activity compared to the most potent flavonoids of the series, 1 and 2.

195

Collectively, all selected natural flavonoids were less potent than the reference drugs, harmine and

196

safinamide. As evident from the IC50 values, both safinamide (MAO-B inhibitor) and harmine

197

(MAO-A inhibitor) are high potency inhibitors of the respective MAO isoforms. Also evident is the

198

fact that these reference inhibitors are highly specific, with safinamide acting as a MAO-B inhibitor

199

and harmine as a MAO-A inhibitor.

200

To better characterize the inhibition mode of the two most potent flavonoid hMAO inhibitors, 1 and

201

2 were incubated with hMAOs at concentrations equal to 4 × IC50 and subsequently dialyzed for 24

202

h.27 The irreversible MAO-A and MAO-B inhibitors, pargyline and (R)-deprenyl, were similarly

203

incubated and dialyzed, and served as positive controls. As negative control, the hMAOs were

204

dialyzed in the absence of inhibitor. After dialysis, the mixtures were diluted twofold to yield

205

inhibitor concentrations equal to 2 × IC50, and the residual MAO activities were measured. These

206

activities were reported as percentage of the residual activity of the negative control (100%). The

207

results of the dialysis experiments showed that 1 and 2 are both reversible hMAO-A and -B

208

inhibitors (Figure 2) since dialysis restored the enzyme activity. In this respect, the enzyme

209

activities were recovered to 104–130% of the negative control. Conversely, inhibition persists in

210

enzyme-inhibitor mixtures that were not dialyzed, with the residual activities at 24–50% of the

211

negative control, and similarly, after irreversible inhibition with pargyline and (R)-deprenyl, both

212

hMAO-A and -B enzymatic activities were not recovered after dialysis and their residual activities

213

remain at 3.2–3.9% of the negative control.

214

We also focused our attention on the most active flavonoids of the current study in order to analyze

215

their inhibition modes. 1 and 2 were shown to act as competitive hMAO-A and -B inhibitors, thus

216

providing additional evidence for their reversible modes of inhibition. For this purpose,

217

Lineweaver-Burk plots were constructed for each inhibitor by measuring hMAO-A and hMAO-B

218

activities in the absence and presence of 1 and 2. As shown in Figure 3, the sets of Lineweaver9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 26

219

Burk plots were, in all instances, typical of competitive inhibition since the linear lines of each set

220

intersect on the y-axis. Ki values for the inhibition of hMAOs by 1 and 2 were estimated from the

221

Lineweaver-Burk plots as well as by global fitting of the inhibition data to the Michaelis-Menten

222

equation. These values are given in Table 2.

223

Natural and synthetic coumarin derivatives, which are structurally related to flavonoids, are known

224

to inhibit the hMAOs28-31 and, in particular the MAO-B isoform. Two X-ray crystal structures of

225

hMAO-B co-crystallized with coumarin derivatives are available in the Protein Data Bank21

226

(accession codes 2V60 and 2V61). However, in order to perform the docking simulations we

227

selected the X-ray crystal structures of the hMAOs in complex with the reference compounds used

228

in this study. Therefore, hMAO-A/harmine and hMAO-B/safinamide complexes were obtained

229

(accession codes 2Z5X22 and 2V5Z,23 respectively). The results of the modelling studies show that

230

1 and 2 bind competitively with the best poses docked into the catalytic site of the enzymes (Figures

231

4 and 5). The theoretical results confirmed that 1 is better accommodated in the hMAO-A active

232

site than that of hMAO-B, with Gscore values of -9.09 and -8.73 Kcal/mol, respectively. Moreover,

233

2 displayed a binding affinity for hMAO-B of -9.21 Kcal/mol, which is higher than the binding

234

affinity to hMAO-A (-8.48 Kcal/mol). The reversal of isoform selectivity of 2 compared to that of 1

235

may be explained by taking into account the different mechanism of recognition observed during

236

the in silico studies. As often remarked, the amino acid residues in the area opposite to the FAD

237

cofactor (hMAO-A, I180, N181, F208, S209 and I335 with the corresponding hMAO-B residues,

238

L171, C172, I199, S200 and Y326 respectively) define the shapes of the active sites with the

239

hMAO-A active binding site being smaller and broader than the elongated and narrower one of

240

hMAO-B.32 Therefore, the docking study shows that in hMAO-A, both the flavonoids bind with the

241

flavonoid portion oriented toward the FAD cofactor, while the phenyl substituents extend towards

242

the active site entrance (Figure 4). In hMAO-B, 2 exhibits a similar positioning, while 1 binds in a

243

reversed orientation (Figure 5). The different binding orientations in hMAO-B may explain the

244

different isoform selectivity of 1 and 2. 10 ACS Paragon Plus Environment

Page 11 of 26

Journal of Agricultural and Food Chemistry

245

As reported in our previous work, bulky natural compounds such as crocin may inhibit the hMAOs

246

non-competitively by possibly interacting with an allosteric site of the enzyme surface.33 In this

247

study, we observed that the two glycosidic derivatives, 6 and 7, which are modest hMAO inhibitors,

248

also bind to a site that is distant from the substrate binding cavity (Table 1 and Figure 6). Therefore,

249

molecular dynamics studies were carried out for the complexes of hMAO-A/B with 1 and 2, in

250

order to further investigate ligand-target interactions established with residues defining the binding

251

pocket, and with 6 and 7 to explore target stabilization by a binding site that is distal from the

252

catalytic site.

253

The molecular dynamics results suggest that the glycosidic flavonoids do not fit into the binding

254

pockets of the hMAOs due to the steric hindrance. Stabilization of complexes is mainly attributed to

255

several H-bond contacts involving -OH groups of the sugar portions and the protein residues. In

256

contrast, within the hMAOs catalytic sites which are defined by mostly hydrophobic amino acid

257

residues, 1 and 2 establish a great number of hydrophobic contacts with key residues as well as

258

several H-bond contacts. As a possible explanation for the higher selectivity of 1 for the hMAO-A

259

compared to 2, the molecular dynamics results revealed that 1 establishes a great number of H-

260

bond/hydrophobic contacts, whereas for 2, which is di-substituted on the phenyl, ligand-target

261

interactions are limited due to intramolecular bonding between the ortho -OH and -OCH3 groups

262

(Table 1).

263

In conclusion, identification of (selective) hMAO inhibitors is of much interest for the therapy of

264

central nervous system affecting diseases and to avoid possible interactions with other serotonergic

265

drugs (serotonin syndrome) and foods rich in dietary-monoamines (cheese effect). Dietary

266

restrictions are strictly mandatory for patients under therapy with MAO inhibitors. For this reason,

267

we assessed the hMAO inhibitory properties of the most important and abundant flavone (1-5) and

268

flavanone (6-8) derivatives and investigated the binding modes of selected derivatives to the

269

hMAOs by molecular modelling studies for the first time. Flavonoids that are active MAO

270

inhibitors are abundantly present in many foods. For example, 1 occurs in many common 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 26

271

vegetables and fruits such as parsley, grapefruit, orange, and onions, and is also present in plant-

272

derived beverages, such as chamomile tea.34 Due to their role as hMAO inhibitors, these natural

273

compounds may be useful for the treatment of motor symptoms in the early stage of

274

neurodegenerative diseases.35

275

Abbreviations

276

CNS, central nervous system; hMAO, human monoamine oxidase; SARs, structure-activity

277

relationships; PDB, Protein Data Bank; SID, Simulation Interactions Diagram; Gscore, Glide

278

scoring function.

279

Acknowledgments

280

This research was funded by “Progetto di Ateneo Ricerca 2013” (P. Chimenti) and the Interregional

281

Research Center for Food Safety and Health at the Magna Græcia University of Catanzaro (MIUR

282

PON a3_00359).

283

Supporting Information

284

Table 1S reports the theoretical affinities of the selected flavonoids for the hMAOs. Figures 1S-4S

285

provides summaries of the MD simulations for the complexes of hMAO-A/B with apigenin (1),

286

diosmetin (2), naringin (6) and eriocitrin (7). This material is available free of charge via the

287

Internet at http://pubs.acs.org.”

288

Conflict of interest

289

The authors declare no conflicts of interest and that have received no payment for the preparation of

290

this manuscript.

291

References

292

1. Tipton, K. F.; Boyce, S.; O’Sullivan, J.; Davey, G. P.; Healey, J. Monoamine oxidases:

293 294 295

certainties and uncertainties. Curr. Med. Chem. 2004, 11, 1965-1982. 2. Youdim, M. B.; Edmondson, D.; Tipton, K. F. The therapeutic potential of monoamine oxidase inhibitors. Nat. Rev. Neurosci. 2006, 7, 295-309.

12 ACS Paragon Plus Environment

Page 13 of 26

296 297 298 299 300 301 302 303

Journal of Agricultural and Food Chemistry

3. Carradori, S.; Silvestri, R. New frontiers in selective human MAO-B inhibitors. J. Med. Chem. 2015, 58, 6717-6732. 4. Carradori, S., Petzer, J. P. Novel monoamine oxidase inhibitors: a patent review (2012-2014). Expert Opin. Ther. Pat. 2015, 25, 91-110. 5. Petzer, A.; Pienaar, A.; Petzer, J. P. The interactions of caffeine with monoamine oxidase. Life Sci. 2013, 93, 283-287. 6. Petzer, J. P.; Petzer, A. Caffeine as a lead compound for the design of therapeutic agents for the treatment of Parkinson’s disease. Curr. Med. Chem. 2015, 22, 975-988.

304

7. Gidaro, M. C.; Astorino, C.; Petzer, A.; Carradori, S.; Alcaro, F.; Costa, G.; Artese, A.; Rafele,

305

G.; Russo, F. M.; Petzer, J. P.; Alcaro, S. Kaempferol as selective human MAO-A inhibitor:

306

analytical detection in Calabrian red wines, biological and molecular modelling studies. J. Agric.

307

Food Chem. 2016, 64, 1394-1400.

308

8. Chaurasiya, N. D.; Ibrahim, M. A.; Muhammad, I.; Walker, L. A.; Tekwani, B. L. Monoamine

309

oxidase inhibitory constituents of propolis: kinetics and mechanism of inhibition of recombinant

310

human MAO-A and MAO-B. Molecules 2014, 19, 18936-18952.

311 312

9. Slimestad, R.; Fossen, T.; Vågen, I. M. Onions: a source of unique dietary flavonoids. J. Agric. Food Chem. 2007, 55, 10067-10080.

313

10. Barreca, D.; Bellocco, E.; Caristi, C.; Leuzzi, U.; Gattuso, G. Flavonoid composition and

314

antioxidant activity of juices from chinotto (Citrus x myrtifolia Raf.) fruits at different ripening

315

stages. J. Agric. Food Chem. 2010, 58, 3031-3036.

316

11. Zhang, L.; Ravipati, A. S.; Koyyalamudi, S. R.; Jeong, S. C.; Reddy, N.; Smith, P. T.; Bartlett,

317

J.; Shanmugam, K.; Münch, G.; Wu, M. J. Antioxidant and anti-inflammatory activities of

318

selected medicinal plants containing phenolic and flavonoid compounds. J. Agric. Food Chem.

319

2011, 59, 12361-12367.

320 321

12. Hwang, S. L.; Shih, P. H.; Yen, G. C. Neuroprotective effects of citrus flavonoids. J. Agric. Food Chem. 2012, 60, 877-885. 13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

322 323 324 325 326 327 328 329 330 331

Page 14 of 26

13. Matos, M. J.; Vazquez-Rodriguez, S.; Uriarte, E.; Santana, L. Potential pharmacological uses of chalcones: a patent review (from June 2011-2014). Expert Opin. Ther. Pat. 2015, 25, 351-366. 14. Kumar, S.; Pandey, A. K. Chemistry and biological activities of flavonoids: an overview. ScientificWorldJournal. 2013, 162750. 15. Gong, J.; Huang, J.; Ge, Q.; Chen, F.; Zhang, Y. Advanced research on the antidepressant effect of flavonoids. Curr. Opin. Complement. Alternat. Med. 2014, 1, e00011. 16. Wasowski, C.; Marder, M. Flavonoids as GABAA receptor ligands: the whole story? J. Exp. Pharmacol. 2012, 4, 9-24. 17. Carradori, S.; D’Ascenzio, M.; Chimenti, P.; Secci, D.; Bolasco, A. Selective MAO-B inhibitors: a lesson from natural products. Mol. Divers. 2014, 18, 219-243.

332

18. Chimenti, F.; Cottiglia, F.; Bonsignore, L.; Casu, L.; Casu, M.; Floris, C.; Secci, D.; Bolasco,

333

A.; Chimenti, P.; Granese, A.; Befani O.; Turini P.; Alcaro S.; Ortuso F.; Trombetta G.; Loizzo

334

A.; Guarino, I. Quercetin as the active principle of Hypericum hircinum exerts a selective

335

inhibitory activity against MAO-A: Extraction, biological analysis, and computational study. J.

336

Nat. Prod. 2006, 69, 945–949.

337 338

19. Strydom, B.; Bergh, J. J.; Petzer, J. P. The inhibition of monoamine oxidase by 8-(2phenoxyethoxy)caffeine analogues. Arzneimittelforschung. 2012, 62, 513-518.

339

20. Petzer, A.; Harvey, B. H.; Wegener, G.; Petzer, J. P. Azure B, a metabolite of methylene blue, is

340

a high-potency, reversible inhibitor of monoamine oxidase. Toxicol. Appl. Pharmacol. 2012,

341

258, 403-409.

342 343

21. The Research Collaboratory for Structural Bioinformatics (RCSB) Protein Data Bank (PDB), http://www.rcsb.org (Accessed: 30th October 2016).

344

22. Son, S. Y.; Ma, J.; Kondou, Y.; Yoshimura, M.; Yamashita, E.; Tsukihara, T. Structure of

345

human monoamine oxidase A at 2.2-A resolution: the control of opening the entry for

346

substrates/inhibitors. Proc. Natl. Acad. Sci. USA. 2008, 105, 5739-5744.

14 ACS Paragon Plus Environment

Page 15 of 26

Journal of Agricultural and Food Chemistry

347

23. Binda, C.; Wang, J.; Pisani, L.; Caccia, C.; Carotti, A.; Salvati, P.; Edmondson, D. E; Mattevi,

348

A. Structures of human monoamine oxidase B complexes with selective noncovalent inhibitors:

349

safinamide and coumarin analogs. J. Med. Chem. 2007, 50, 5848-5852.

350

24. PubChem, https://pubchem.ncbi.nlm.nih.gov (Accessed: 30th October 2016).

351

25. Novaroli, L.; Reist, M.; Favre, E.; Carotti, A.; Catto, M.; Carrupt, P. A. Human recombinant

352

monoamine oxidase B as reliable and efficient enzyme source for inhibitor screening. Bioorg.

353

Med. Chem. 2005, 13, 6212-6217.

354

26. Chimenti, P.; Petzer, A.; Carradori, S.; D'Ascenzio, M.; Silvestri, R.; Alcaro, S.; Ortuso, F.;

355

Petzer, J. P.; Secci, D. Exploring 4-substituted-2-thiazolylhydrazones from 2-, 3-, and 4-

356

acetylpyridine as selective and reversible hMAO-B inhibitors. Eur. J. Med. Chem. 2013, 66,

357

221-227.

358 359

27. Mostert, S.; Petzer, A.; Petzer, J. P. Indanones as high-potency reversible inhibitors of monoamine oxidase. ChemMedChem. 2015, 10, 862-873.

360

28. Guo, B.; Zheng, C.; Cai, W.; Cheng, J.; Wang, H.; Li, H.; Sun, Y.; Cui, W.; Wang, Y.; Han, Y.,

361

Lee, S. M.; Zhang, Z. Multifunction of chrysin in Parkinson's model: anti-neuronal apoptosis,

362

neuroprotection via activation of MEF2D, and inhibition of monoamine oxidase-B. J. Agric.

363

Food Chem. 2016, 64, 5324-5333.

364

29. Bandaruk, Y.; Mukai, R.; Kawamura, T.; Nemoto, H.; Terao, J. Evaluation of the inhibitory

365

effects of quercetin-related flavonoids and tea catechins on the monoamine oxidase-A reaction

366

in mouse brain mitochondria. J. Agric. Food Chem. 2012, 60, 10270-10277.

367

30. Lee, M. H.; Lin, R. D.; Shen, L. Y.; Yang, L. L.; Yen, K. Y.; Hou, W. C. Monoamine oxidase B

368

and free radical scavenging activities of natural flavonoids in Melastoma candidum D. Don. J.

369

Agric. Food Chem. 2001, 49, 5551-5555.

370

31. Chimenti, F.; Secci, D.; Bolasco, A.; Chimenti, P.; Granese, A.; Carradori, S.; Befani, O.; Turini,

371

P.; Alcaro, S.; Ortuso, F. Synthesis, molecular modeling studies, and selective inhibitory

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 26

372

activity against monoamine oxidase of N,N’-bis[2-oxo-2H-benzopyran]-3-carboxamides.

373

Bioorg. Med. Chem. Lett. 2006, 16, 4135-4140.

374

32. D’Ascenzio, M.; Carradori, S.; Secci, D.; Mannina, L.; Sobolev, A. P.; De Monte, C.; Cirilli, R.;

375

Yáñez, M.; Alcaro, S.; Ortuso, F. Identification of the stereochemical requirements in the 4-

376

aryl-2-cycloalkylidenhydrazinylthiazole scaffold for the design of selective human monoamine

377

oxidase B inhibitors. Bioorg. Med. Chem. 2014, 22, 2887-2895.

378

33. De Monte, C.; Carradori, S.; Chimenti, P.; Secci, D.; Mannina, L.; Alcaro, F.; Petzer, A.; N'Da,

379

C. I., Gidaro, M. C.; Costa, G.; Alcaro, S.; Petzer, J. P. New insights into the biological

380

properties of Crocus sativus L.: chemical modifications, human monoamine oxidases inhibition

381

and molecular modeling studies. Eur. J. Med. Chem. 2014, 82, 164-171.

382 383 384 385

34. Shukla, S.; Gupta, S. Apigenin: a promising molecule for cancer prevention. Pharm. Res. 2010, 27, 962-978. 35. Solanki, I.; Parihar, P.; Mansuri, M.L.; Pariha, M.S. Flavonoid-based therapies in the early management of neurodegenerative diseases. Adv. Nutr. 2015, 6, 64-72.

386

16 ACS Paragon Plus Environment

Page 17 of 26

Journal of Agricultural and Food Chemistry

387

Figure captions

388

Figure 1. Structures of the selected flavonoids (1-8) assayed as hMAO inhibitors.

389

Figure 2. Dialysis restores the activities of the hMAOs following inhibition by (A) apigenin (1) and

390

(B) diosmetin (2). Dialysis, however, does not restore the hMAO activities following inhibition by

391

the irreversible inhibitors, pargyline (parg) and (R)-deprenyl (depr). NI, no inhibitor. *Statistical

392

significantly different from the mean of inhibitor–dialyzed. The values are given as mean ± SD of

393

triplicate determinations.

394

Figure 3. Lineweaver-Burk plots for the inhibition of hMAO-A and hMAO-B by (A) apigenin (1)

395

and (B) diosmetin (2). Also shown as insets are replots of the slopes of the Lineweaver-Burk plots

396

versus inhibitor concentration.

397

Figure 4. Best orientations into the hMAO-A catalytic site of (A) apigenin (1) and (B) diosmetin

398

(2), displayed as orange and violet sticks, respectively. The FAD cofactor is shown grey sticks

399

while the amino acid residues that are involved in ligand-target interactions are shown as grey lines.

400

The hMAO proteins are represented as grey surfaces and cartoons. All non-carbon atoms are

401

colored according to atom types.

402

Figure 5. Best orientations into the hMAO-B catalytic site of (A) apigenin (1) and (B) diosmetin

403

(2), displayed as orange and violet sticks, respectively. The FAD cofactor is shown as grey sticks

404

while the amino acid residues that are involved in ligand-target interactions are shown as grey lines.

405

The hMAO proteins are represented as grey surfaces and cartoons. All non-carbon atoms are

406

colored according to atom types.

407

Figure 6. Molecular recognition of (A) naringin (6) and (B) eriocitrin (7), shown as blue and yellow

408

sticks, respectively, by a binding site distal from the catalytic site of the hMAOs. The hMAO

409

proteins are displayed as transparent surfaces and the FAD cofactors are shown as grey sticks in

410

hMAO-A and cyan sticks in hMAO-B, respectively. All non-carbon atoms of the ligands are

411

colored according to atom types.

412 17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 26

413

Table 1. IC50 Values for the Inhibition of hMAO-A and hMAO-B by Selected Flavonoids 1-8 and

414

Reference Drugs. Compound 1 2 3 4 5 6 7 8 Harmine Safinamide

hMAO-A inhibition (µM)* 1.55 ± 0.147 5.74 ± 0.571 > 100 64.2 ± 7.69 > 100 33.3 ± 7.05 86.5 ± 32.4 > 100 0.0029 ± 0.00042 112 ± 5.23

hMAO-B inhibition (µM)* 5.16 ± 0.410 1.58 ± 0.887 > 100 21.2 ± 4.99 > 100 µM 44.6 ± 11.2 164 ± 39.2 > 100 >100 0.048 ± 0.0047

*The values are given as mean ± SD of triplicate determinations.

18 ACS Paragon Plus Environment

Page 19 of 26

Journal of Agricultural and Food Chemistry

Table 2. Ki Values for the Inhibition of hMAOs by Apigenin (1) and Diosmetin (2).

1 2

Ki determined from Lineweaver-Burk plots (µM) hMAO-A hMAO-B 1.15 4.91 4.60 1.43

Ki determined by global fitting to MichaelisMenten equation (µM) hMAO-A hMAO-B 0.901 ± 0.040 (r2 = 0.99) 3.48 ± 0.266 (r2 = 0.99) 4.69 ± 0.259 (r2 = 0.99) 0.860 ± 0.056 (r2 = 0.99)

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 26

OH

OH OH HO

OH

OCH3

O

O

HO

O

HO

OH OH O

OH O

OH O Diosmetin (2)

Apigenin (1)

Taxifolin (3)

OCH3 OH HO

OH OH HO

O

O OH O O HO

OH OH O

OH O

OH OH

Quercetin-3-O-ß-D-glucoside (isoquercitrin) (5)

Isorhamnetin (4)

OH OH OH

CH3 HO

O

HO

O

HO HO HO HO

O

OH O O

O

OH

O

O OH

O

HO

OH

O

O

H3C

OH O

OH OH

Naringin (6)

Eriocitrin (7) OH

HO HO H3C

O

OH

O HO HO

OCH3 O

O

OH OH O Hesperidin (8)

415

Figure 1 20 ACS Paragon Plus Environment

Page 21 of 26

Journal of Agricultural and Food Chemistry

A

MAO-B

MAO-A 150 125 100

100

Rate (%)

Rate (%)

125

75 50

50 25

25

*

0

NI

1

parg

dialysed

B

75

*

0

NI

1

1

depr

dialysed

undialysed

MAO-A

1 undialysed

MAO-B 125

100

Rate (%)

Rate (%)

100

75

50

25

75 50 25

*

0

NI

2

parg

dialysed

*

0

2

NI

undialysed

2

depr

dialysed

2 undialysed

416 Figure 2

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

A

1200

800

100

800 400

400

75 −2

−1

0 [I], µ M

1

50

25

0

0

0.00

0.02 0.04 1/[S]

−0.02

0.06

0 2 [I], µ M

4

6

50

25

−0.02

−4 −2

2

1/V (%)

1/V (%)

75

Slope

Slope

1200

100

Page 22 of 26

0.00

0.02 0.04 1/[S]

0.06

1 (MAO-B)

1 (MAO-A)

B

1/V (%)

75

800

100

400 −4

−1

2 [I], µ M

5

75

8

50

25

0

0 0.00

0.02 0.04 1/[S]

0.06

800 400

−2

−1

0 1 [I], µ M

2

50

25

−0.02

Slope

1200

1/V (%)

100

Slope

1200

−0.02

0.00

2 (MAO-A)

0.02 0.04 1/[S]

0.06

2 (MAO-B)

417 Figure 3 418

22 ACS Paragon Plus Environment

Page 23 of 26

Journal of Agricultural and Food Chemistry

419

Figure 4

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 26

420

Figure 5

24 ACS Paragon Plus Environment

Page 25 of 26

Journal of Agricultural and Food Chemistry

Figure 6

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 26

For Table of Contents Only SAR regarding MAO inhibition of the selected flavonoids

Substitution with sugars decreased or abolished hMAO inhibitory activity

Glycoside O

R1 R2

Mono-substitution led to hMAO-A selectivity; disubstitution led to hMAO-B selectivity

O OH OH O

Insaturation is important; introduction of -OH group decreased hMAO inhibitory activity

26 ACS Paragon Plus Environment