Oleanolic, Ursolic, and Betulinic Acids as Food Supplements or

(1) Type 2 diabetes is characterized by hyperglycemia resulting from insulin resistance, .... Oleanolic, ursolic, and betulinic acids (IC50 = 14.9, 9,...
0 downloads 0 Views 1MB Size
Subscriber access provided by MAHIDOL UNIVERSITY (UniNet)

Review

Oleanolic, Ursolic and Betulinic Acids as Food Supplements or Pharmaceutical agents for Type 2 Diabetes – Promise or Illusion? Filomena S.G. Silva, Paulo J. Oliveira, and Maria Fatima Duarte J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.5b06021 • Publication Date (Web): 25 Mar 2016 Downloaded from http://pubs.acs.org on March 29, 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 69

Journal of Agricultural and Food Chemistry

1

Oleanolic, Ursolic and Betulinic Acids as Food Supplements or Pharmaceutical agents for

2

Type 2 Diabetes – Promise or Illusion?

3

Filomena S.G. Silva†, Paulo J. Oliveira‡, Maria F. Duarte†*

4 5 6 7 8

Centro de Biotecnologia Agrícola e Agro-Alimentar do Alentejo (CEBAL)/Instituto Politécnico de Beja (IPBeja),

9

ABSTRACT

10

Oleanolic (OA), ursolic (UA) and betulinic (BA) acids are three triterpenic acids (TAs) with

11

potential effects for treatment of type 2 diabetes. Mechanistic studies showed that these TAs act

12

as hypoglycemic and anti-obesity agents mainly through (i) reducing the absorption of glucose;

13

(ii) decreasing endogenous glucose production; (iii) increasing insulin sensitivity; (iv) improving

14

lipid homeostasis; and (v) promoting body weight regulation. Besides these promising beneficial

15

effects, it is believed that OA, UA and BA protect against diabetes-related comorbidities due to

16

their anti-atherogenic, anti-inflammatory and anti-oxidant properties. We also highlight the

17

protective effect of OA, UA and BA against oxidative damage which may be very relevant for

18

the treatment and/or prevention of T2DM. In the present review, we make an integrative

19

description of the anti-diabetic properties of OA, UA and BA, evaluating the potential use of

20

these TAs as food supplements or pharmaceutical agents, in order to prevent and / or treat

21

T2DM.

22

KEYWORDS: Oleanolic acid, ursolic acid, betulinic acid, anti-diabetic, hepatic disease,

23

cardiovascular disease, anti-obesity.

Apartado 6158, 7801-908 Beja, Portugal. *Corresponding Author Phone: + 351 284314399, email: [email protected] CNC, Center for Neuroscience and Cellular Biology, UC-Biotech Building, Biocant Park, University of Coimbra, 3060-107 Cantanhede, Portugal.

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 69

24

INTRODUCTION

25

Diabetes is a chronic metabolic disorder that is currently assuming epidemic proportions.

26

According to statistics, it was estimated that about 9% of the adult worldwide population had

27

diabetes in 2014, with 90% among those having type 2 diabetes mellitus (T2DM).1 Type 2

28

diabetes is characterized by hyperglycemia resulting from insulin resistance, as well as

29

insufficient insulin secretion by pancreatic β-cells. Accumulating evidence demonstrates that

30

T2DM is associated with obesity,2 as well as to development of several comorbidities, including

31

hepatic, cardiac and renal disorders.3 Management of T2DM requires an integrated approach,

32

which includes hyperglycemia and obesity treatments, as well as prevention of diabetic

33

comorbidities. However, the currently available anti-diabetic drugs have limited efficacy and/or

34

safety concerns, and only temporarily improve blood glucose levels, failing in the treatment of

35

obesity, as well as in the prevention of diabetes complications. Therefore, the identification of

36

new therapeutic agents, which may have the potential to prevent, or treat diabetes and its related

37

comorbidities, are an unmet need.

38

Consumption of different natural compounds is known to result in anti-diabetic effects, offering

39

exciting possibilities for the future development of successful therapies. Terpenoids are a diverse

40

group of phytochemical agents derived from squalene or related acyclic and cyclic 30-carbon

41

precursors, which are currently receiving attention because of their wide range of biological

42

activity.4 An increasingly number of publications has focused on pentacyclic triterpenoids,

43

lupane, oleanane and ursane-type structures, particularly oleanolic (3β-hydroxyolean-12-en-28-

44

oic, OA), ursolic (3β-hydroxyurs-12-en-28-oic, UA) and betulinic (3β-hydroxy-lup-20-en-28-

45

oic, BA) acids (Figure 1) due to their anti-diabetic properties (reviewed in5-7).

2 ACS Paragon Plus Environment

Page 3 of 69

Journal of Agricultural and Food Chemistry

46

Oleanolic, ursolic and betulinic acids have several important biological activities, including anti-

47

tumoral, anti-inflammatory and anti-oxidant effects,4 as well as hepato-,8-11 cardio-12-14 and

48

nephroprotective8,

49

potential application for T2DM management, and its associated comorbidities. The biological

50

effects mainly result from their anti-hyperglycemic,16-19 anti-hyperlipidemic,20-23 anti-

51

atherogenic,24-26 anti-oxidant27-29 and anti-inflammatory27,

52

that OA, UA and BA, or plants extracts rich in those TAs reduce the absorption and uptake of

53

glucose, lower endogenous glucose production, increase insulin biosynthesis, secretion and

54

sensitivity, and protect against diabetic complications.6, 7 These protective effects occur without

55

noticeable hepatotoxicity, at the dose used in the experimental studies, as indicated by the

56

decreased values of aspartate aminotransferase and alanine aminotransferase measured.21,

57

This relevant fact is contrary to most currently used anti-diabetic drugs, including sulfonylureas,

58

α-glucosidase inhibitors, biguanides and thiazolidinediones, which can cause liver toxicity.33, 34

59

The low toxicity profile of these compounds has also been evidenced in several in vivo studies,

60

where OA (10-1700 mg/kg)35-39, UA (25-2500 mg/kg)10, 40, 41 and BA (25-50 mg/kg)42 protected

61

the liver from oxidative damage, at concentrations equivalent or higher to those used in diabetic

62

studies, when administered to rodents pre-treated with different hepatotoxicants. The protective

63

effect of OA, UA and BA against oxidative damage may be critical for their use in the treatment

64

and prevention of diabetes and their comorbidities.

65

In this context, in the present unique review, we make an integrative description of the

66

hypoglycemic, anti-obesity properties of OA, UA and BA, as well as their effects on diabetic

67

comorbidities in the management of T2DM. We here underline the main targets of action and

68

limitations of these compounds, exploring the current knowledge and open questions. The

15

roles. The three triterpenic acids (TAs) described above also have a

30, 31

action. Published data indicate

22, 32

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 69

69

present work focuses on the main issues that research in this field should take into account,

70

emphasizing what should be clarified in the future, and evaluating the potential use of these TAs

71

as food supplements or pharmaceutical agents, in order to prevent and / or treat T2DM.

72

MECHANISMS BEHIND HYPOGLYCEMIC EFFECTS

73

Numerous research described hypoglycemic effects for OA, UA and BA by using in vitro and in

74

vivo models. The description of hypoglycemic effects of these three TAs is summarized in Table

75

1.

76

Absorption and uptake of glucose:

77

Inhibition of carbohydrate hydrolysis, and down-regulation of glucose transporters in the

78

gastrointestinal tract decreases post-prandial hyperglycemia,43, 44 suppressing, at least in part, the

79

development of T2DM.45

80

Oleanolic, ursolic and betulinic acids inhibit pancreatic α-amylase activity in vitro ([TAs]: 5–50

81

µg/mL ≈ 10–100 µM),46,

82

(HFD)-induced obese mice.32, 48, 49 Additionally, other authors also showed that these three TAs

83

have also the capacity to decrease α-glucosidase (IC50: 0.1–231 µM)47, 50-54 activity in vitro. Also,

84

OA inhibits human pancreatic α-amylase activity (IC50 0.1 mg/mL ≈ 200 µM),55 and decreases

85

post-prandial hyperglycemia in diabetic rats by inhibiting the activities of α-amylase (IC50 3.60

86

µM), and α-glucosidase (IC50 12.40 µM), and simultaneously down-regulating the expression of

87

sodium-dependent glucose co-transporter (SGLT-1) and glucose transporter -2 (GLUT-2) in the

88

small intestine16 (Table 1). Taken together, these results suggest that inhibition of α-amylase and

89

α-glucosidase, as well as down-regulation of glucose transporters by the above mentioned TAs

90

can delay glucose absorption in the small intestine in diabetes, consequently preventing post-

47

and in vivo (5-20 mg/kg/day during 15 weeks) on high-fat diet

4 ACS Paragon Plus Environment

Page 5 of 69

Journal of Agricultural and Food Chemistry

91

prandial plasma glucose increase. However, the reduced number of animal testing and the lack of

92

human studies raise questions about the true effectiveness of these compounds in vivo,

93

highlighting the need for more studies in living systems.

94

Endogenous glucose production and glycogen synthesis:

95

T2DM increases endogenous glucose production and reduces the synthesis of glycogen by

96

increasing gluconeogenesis, glycogenolysis, and by reducing glycogenesis.56

97

Oleanolic, ursolic and betulinic acids (IC50: 14.9, 9 and 43 µM, respectively) inhibit glycogen

98

phosphorylase (GP),57 the rate-limiting step of glycogenolysis, which decreases glycogenolysis

99

and also stimulates glycogen synthesis.58 Moreover, UA treatment (5 mg/kg/day or 0.01-0.05%,

100

w/w) increases glycogen content by increasing hepatic glucokinase activity in vivo mice

101

models,17,

102

glucose levels by inhibiting hepatic glucose-6-phosphatase (G6Pase) activity,17, 20, 59 the ultimate

103

step in glucose production.61 In line with increased hepatic glycogen, UA (2 mg/kg/day) also

104

increases glycogen deposition in healthy rats,62 most likely by promoting the phosphorylation

105

and inactivation of glycogen synthase kinase-3 (GSK-3), an enzyme which inhibits glycogen

106

synthesis pathway.63 Additionally, OA, UA and BA (10 µM) stimulate glucose uptake and

107

glycogen synthesis in human HepG2 hepatoma cells via 5′ AMP-activated protein kinase

108

(AMPK)-GSK-3β pathway.64 Once activated, AMPK, an energy sensor that regulates cellular

109

metabolism, leads to the inhibition of glucose production by increasing GSK-3β

110

phosphorylation. Activation of AMPK decreased gluconeogenesis by regulating gluconeogenic

111

targets, such as phosphoenolpyruvate carboxykinase (PEPCK) and G6Pase.65 In vivo treatment

112

with OA (20 mg/kg/day) decreased G6Pase and PEPCK protein contents,18 which may result in

113

lower gluconeogenesis rates. A similar in vitro study, but using BA instead (10 and 20 µM) 5

20, 59

a key regulatory step in liver glycogen formation,60 and also decreases blood

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 69

114

demonstrated an effect on hepatic gluconeogenesis by decreasing PEPCK and G6Pase mRNA,

115

resulting from AMPK activation19. Another study also evidenced that once activated by BA, the

116

AMPK pathway promoted glucose utilization and glucose transport across the cell membrane, by

117

induction of glycolysis and by up-regulation of GLUT-1 and -2 protein expression in mouse

118

embryonic fibroblasts. This study also demonstrated that glycolysis stimulation by BA was

119

accompanied by a reduction of glucose oxidation66 (Table 1). According to the results, OA, UA

120

and BA reduced glucose availability in the body by limit endogenous glucose production, also

121

promoting glycogen synthesis in T2DM, contribute to the hypoglycemic effect of these

122

compounds. However, future studies are warranted to confirm these results using more in vivo

123

assays, including in humans.

124

Insulin resistance:

125

T2DM mainly results from insulin resistance and β-cell dysfunction. The improvement of insulin

126

sensitivity represents a good strategy to prevent T2DM development.67

127

Numerous studies have supported the notion that OA, UA and BA improve insulin responses by

128

acting as insulin secretagogue and insulinomimetic agents in different contexts.68-73 According to

129

different reports, the three TAs interfere with insulin biosynthesis, secretion and sensitivity

130

involving

131

phosphatases,74-79 by stimulating phosphatidylinositol 3-kinase-dependent (PI3KB)/protein

132

kinase Akt pathway,72, 73, 80-82 by acting as TGR-5 receptors 83-86, by improving the survival of β-

133

cells68, 69, 87 and/or by acting as anti-inflammatory and anti-oxidant agents 27, 28, 31.

134

Protein tyrosine phosphatases and PI3KB/Akt pathway: Several in vitro studies provided

135

evidence that OA, UA and BA directly inhibited protein tyrosine phosphatase 1B (PTP-1B),74-79

136

thus improving insulin resistance in T2DM and obesity.88-91 PTP-1B is a molecule that 6

multi-target

mechanisms,

mainly

by inhibiting

ACS Paragon Plus Environment

different

protein

tyrosine

Page 7 of 69

Journal of Agricultural and Food Chemistry

137

negatively regulates insulin signaling. Insulin regulates glucose homeostasis through: i) binding

138

to its receptor to initiate a signaling cascade; ii) activating and promoting phosphorylation of

139

insulin receptor substrate proteins; and iii) mediating the PI3K/Akt pathway.92, 93 Akt pathway

140

activation facilitated glucose uptake into adipose tissue, cardiac muscle and skeletal muscle, by

141

mediating the translocation of glucose transporter -4 (GLUT-4) to the plasma membrane, and by

142

inhibiting gluconeogenesis.5 However, PTP-1B inhibited the PI3K/Akt signaling pathway

143

resulting in insulin resistance. The mechanism is thought to result from inhibition of GLUT-4

144

translocation to the plasma membrane, decreasing glucose re-uptake and gluconeogenesis.

145

Therefore, the direct inhibition of PTP-1B by OA, UA and BA may improve insulin sensitivity.

146

Nevertheless, it was described that UA acted directly in the insulin receptor (IR).80 In vitro

147

studies performed by Jung et al. in cultured adipocytes and Chinese-hamster ovary cells

148

expressing human IR80 showed that UA is an insulin sensitizer, leading to an increase in IRβ

149

auto-phosphorylation and to a subsequent activation of downstream PI3K signaling pathway.

150

The activation of this pathway would result in phosphorylation and inactivation of GSK-3,

151

leading to an increase of glycogen synthesis,63 as described before. Also, OA increased IR and

152

insulin receptor substrate-1 (IRS-1) mRNA and upregulated total IRS-1 protein expression in

153

adipose tissue insulin resistance rats, suggesting that this compound improves insulin resistance

154

via the IRS-1/PI3k/Akt pathway.81 Regarding UA (10-20 µM), this molecule was described to

155

increase insulin sensitivity, by promoting GLUT-4 translocation in 3T3-L1 adipocytes,80,

156

probably by activating the PI3K pathway.82 In a similar way, in vivo studies showed that

157

treatment with UA or BA (10 mg/kg) increases glycogen content and glucose uptake in muscle

158

by acting as insulin secretagogue and insulinomimetic agents via PI3K and MAPK, stimulating

159

GLUT-4 synthesis and translocation in muscle.72, 73 All these effects are summarized in Table 1.

82

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 69

160

TGR5 receptors: Accumulating evidence suggest that OA,83 UA and BA84 act in vitro as

161

selective TGR5 agonists (EC50 1.04-2.25 µM). TGR5 is a G-protein-coupled receptor identified

162

as the first cell surface receptor involved in energy homeostasis and which is activated by bile

163

acids. Its activity ameliorates insulin resistance.94 Indeed, OA (10-30 µM) activated TGR5 in

164

enteroendocrine cells, leading to glucagon-like peptide-1 (GLP-1) release.85 GLP-1 is a potent

165

glucose-dependent insulin-stimulating hormone,95 that stimulates β-cell proliferation and pro-

166

insulin gene expression,96 and which inhibits glucagon expression.97 Oleanolic acid (10-50 µM)

167

also activated TGR5 in pancreatic β-cells, leading to TGR5-mediated glucose-stimulated insulin

168

release 86 (Table 1). Contradictorily, a human phase I clinical trial showed that the hypoglycemic

169

effect of intraduodenal OA perfusions (1 mM) was not accompanied by the release of GLP-1 in

170

healthy male volunteers (University Hospital, Basel, Switzerland, coded NCT01674946).

171

Improvement of β-cell survival: Current evidence shows that decreased function of β-cells

172

always precedes the development of T2DM and induces profound metabolic dysfunction.

173

Therefore, β-cells function preservation or recovery can be an effective therapeutic strategy to

174

prevent or treat T2DM.98 It is known that UA (0.05%, w/w) preserved the integrity of β-cells,68

175

and islet architecture69 in the pancreas, contributing to prevent the metabolic dysfunction of β-

176

cells and consequently the development of T2DM. Oleanolic acid (0.5 mg/day during 2 days)

177

prolonged the survival of transplanted islets due to its anti-inflammatory, immune-regulatory and

178

anti-oxidant properties. Beneficial effects were suggested to be mediated by decreased γ-

179

interferon-inducible protein 10 and IL-4 in serum, the number of T cells secreting γ-interferon,

180

IL-4, IL-7, and IL-2, and the infiltration of CD4 and CD8 cells, as well as decreased reactive

181

oxygen species (ROS) production,87 supporting the notion that the anti-inflammatory and anti-

8 ACS Paragon Plus Environment

Page 9 of 69

Journal of Agricultural and Food Chemistry

182

oxidant effects of these compounds can also contribute to preserve the integrity of β-cells and

183

consequently the development of T2DM.

184 185

Anti-inflammatory and anti-oxidant properties:

186

Different studies demonstrated that OA,99 UA30 and BA31 down-regulate nuclear factor kappa B

187

(NF-κB) signaling pathway in different cell models, which may decrease several inflammatory

188

cytokines, and consequently improve insulin response in diabetic patients100, supporting the

189

notion which the anti-inflammatory effects of these compounds may in turn prevent T2DM.

190

Similarly, it was also described that the reduction of inflammatory cytokines tumor necrosis

191

factor -α (TNF-α) and IL-6 mediated by reduction of NF-kB promoted by OA, may in turn

192

down-regulate the expression of IRS-1 and GLUT-4 in HepG2 cell line,27 evidencing the crucial

193

role between the reduction of oxidative stress and the suppression of inflammatory response in

194

the prevention of T2DM. Another study also suggested that OA (20 mg/kg/day during 2 weeks)

195

improves hepatic insulin resistance by acting as anti-inflammatory, down-regulating IL-1β, IL-6

196

and TNF-α1, and as anti-oxidant by increasing superoxide dismutase (SOD), catalase (CAT)

197

protein contents and mitochondrial reduced glutathione pool in diabetic mice. In the same study,

198

OA also increased nuclear factor (erythroid-derived 2)-like 2 (Nrf2) protein expression18, which

199

promoted the transcriptional induction of several anti-oxidant genes and suppression of

200

inflammatory responses101, contributing thus to the improvement of insulin resistance102. The

201

authors suggested that this increase of Nrf2 may in turn be responsible for reducing

202

mitochondrial redox balance, also improving mitochondrial dysfunction, and also promoting the

203

activation of AMPK signaling in diabetic mice18 . Although this study suggested that Nrf-2

204

mediated protection may be a key mechanism in improving insulin resistance induced by OA, no

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 69

205

more studies showed this effect in T2DM. However, the induction of Nrf2-dependent genes was

206

detected when OA (22.5-90 mg/kg/day for 3 or 4 days)103-105 and UA (25 and 50 mg/kg/daily for

207

one week)41 were administered as hepatoprotective agents against different hepatotoxicants in

208

vivo studies. This supported the notion the two TAs suppressed inflammatory and oxidative

209

response through the Nrf2 pathway, and perhaps contributed to reduce insulin resistance through

210

that same mechanism. Another study also reported that UA (0.125%, 0.25%, 0.5% during 6

211

weeks) promoted anti-inflammatory effects by decreasing TNF-α, monocyte chemotactic

212

protein-1 (CCL2/MCP-1), IL-1β, IL-2, IL-6 and IL-8, and anti-oxidant effects, mediated by

213

increasing SOD, CAT, and glutathione peroxidase (GPx) activities, and by decreasing

214

malondialdehyde (MDA), which also indirectly contributed to improve insulin resistance.28 This

215

study supported the notion that UA contributed to improve insulin resistance due to its anti-

216

inflammatory28,

217

inflammatory31 and anti-oxidant properties29, 31 are extensively described there are no studies that

218

suggest the direct contribution of those effects in the improvement of insulin resistance. Instead a

219

recent study indicated that BA was not able to activate Nrf2 pathway in Chinese hamster ovary

220

cells66. Altogether, the data supports the notion that the anti-inflammatory and anti-oxidant

221

properties of these compounds contribute, at least in part, to prevent development of T2DM,

222

although doubts still exist on the role of Nrf2 regulation.

223

30

and anti-oxidant properties28,

40, 41

. Regarding BA, despite its anti-

ANTI-OBESITY EFFECTS AND THEIR MECHANISMS OF ACTION

224

In association with improved glucose tolerance and insulin sensitivity, OA, UA and BA are also

225

anti-obesity promoters (Table 2).

226

Several studies demonstrated the improvement of glucose tolerance being often accompanied by

227

alterations in the lipid profile, by decreasing total cholesterol, triglycerides (TG) LDLc contents 10 ACS Paragon Plus Environment

Page 11 of 69

Journal of Agricultural and Food Chemistry

228

and by increasing the levels of HDLc in streptozotocin-106 and alloxan-21 -diabetic rats treated

229

with OA (60-200 mg/kg/ day, during 40 days). The use of UA and BA (10 mg/kg/day during 15

230

days) was associated with decreased glucose, total cholesterol and TG contents in HFD-induced

231

obese mice.32, 48, 49 Additionally, treatment with OA, BA (5-20 mg/kg/ day, during 15 weeks) and

232

UA (5-10 mg/kg/ day or 0.05% w/w during 5-15 weeks) also reduced body weight, visceral

233

adiposity32,

234

adipocyte size20 in obese diabetic rodents. Accumulating evidence suggest that the three TAs act

235

as anti-obesity agents by reducing lipid absorption,107, 108 improving lipid homeostasis109-114 and

236

promoting food intake, and body weight regulation.32,

237

humans places into question the true hypolipidemic effects of these compounds, clearly

238

highlighting the need for future human studies.

239

Lipid absorption:

240

Post-prandial hyperlipidemia in diabetes is a risk factor for the development of hyperlipidemia

241

complications. Hence, the absorption of dietary lipids by the small intestine plays an important

242

role in the regulation of post-prandial hyperlipidemia.

243

Ursolic107 and betulinic acids107, 108 exert hyperlipidemic effects, at least in part, by inhibiting

244

pancreatic lipase (IC50 values between 15.83 to 21.10 µM), which prevent the absorption of

245

lipids from the small intestine. Other studies showed that OA and BA have inhibitory activities

246

on human cholesterol acyltransferase-2 (hACAT-2),115 an isoenzyme responsible for cholesterol

247

absorption in the intestine.116 When comparing OA and BA, the latter exhibited more potent

248

hACAT-2 inhibitory activity, with an IC50 value of 28.8 µM.115 Taken together, these results

249

suggest that OA and BA also decrease the lipid absorption process in intestinal mucosal cells

250

(Table 2).

48, 49

, free fatty acids22 and hepatic lipid accumulation,20,

48, 49

69

restoring the normal

However, the lack of studies in

11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 69

251

Lipogenesis and fatty acid β-oxidation:

252

The inhibition of free fatty acids (FFA) synthesis (lipogenesis)117 and increasing fatty acid β-

253

oxidation118 are two potential pathways which contribute to the reduction of hepatic fat

254

accumulation in T2DM patients. Compounds which inhibit the transcription of sterol-regulatory-

255

element-binding protein-1c (SREBP-1c), and consequently mediate the down- regulation of

256

hepatic genes involved in lipogenesis, such as acetyl-coA carboxylase (ACC) and fatty acid

257

synthase (FAS),119 result in a reduction of FFA synthesis. Additionally, agents that stimulate the

258

up-regulation of carnitine palmitoyltransferase-1 (CPT-1),120 acyl-CoA oxidase (ACO)121 and

259

peroxisome proliferator activated receptor-α (PPAR-α),122 increase fatty acid β-oxidation. An in

260

vivo study showed that OA (25-50 mg/kg/ day) reduced the expression of lipogenic genes

261

including ACC,109, 110 glycerol-3-phosphate acyltransferase,110 and FAS,109 thus inhibiting lipid

262

synthesis, and consequently the retention of lipids in hepatocytes. Treatment with UA (5

263

mg/kg/day) down-regulated SREBP-1c, FAS and ACC transcripts, while promoting the

264

inhibition of lipogenesis in liver, and increased fatty acid oxidation (by upregulating CPT-1 and

265

PPAR-α).22 Other studies described that the hypolipidemic effects of UA in vitro and in vivo are

266

primarily achieved by the activation of hepatic PPAR-α, which acts as an antagonist. Once

267

activated, PARP-α promotes fatty acid uptake and fatty acid β-oxidation, and decreases

268

lipogenesis. Ursolic acid also increased the expression of fatty acid transport protein 4, enhances

269

hepatic fatty acid uptake and induced acyl-CoA synthetase-1, CPT-1, and ACO1, improving

270

hepatic fatty acid β-oxidation.23, 111 Besides that, UA decreased FAS and SREBP-1c transcripts,

271

while inhibiting lipogenesis.23 Interestingly, Jia et al. showed that UA induces hepatic autophagy,

272

which could also contribute to the degradation of lipid droplets in hepatocytes, and consequently

273

regulate hypolipidemic effects of UA.23 It was also proposed that UA also reduces intracellular 12 ACS Paragon Plus Environment

Page 13 of 69

Journal of Agricultural and Food Chemistry

274

fat storage in skeletal muscle cells, by increasing FFA uptake and β-oxidation via activation of

275

AMPK pathway and upregulation of uncoupling protein (UCP)-3

276

membrane protein involved in mitochondrial uncoupling and fatty acid metabolism124,

277

contributing thus to explain also the anti-obesity effects induced by this compound. Interestingly,

278

BA (20-40 µM) decreased lipogenesis and lipid accumulation, suppressing SREBP-1, FAS and

279

SCD-1 mRNA expression, and SREBP-1 nuclear translocation through the AMPK pathway125.

280

This strengthens the notion that the activation of AMPK may be involved in hypoglycemic and

281

hypolipidemic effects of these compounds. hypolypidemic effects of the compounds are

282

described in Table 2.

283

Adipogenesis and lipolysis:

284

Several studies demonstrated that OA112, 114, 126, UA113, 127, 128 and BA108, 129 regulate adipogenesis

285

by inhibiting the expression and/or activation of different adipogenic transcriptional factors,

286

which may in turn reduce obesity and ultimately prevent the insulin resistance in T2DM.

287

Oleanolic acid (1-25 µM) attenuated adipogenesis by reducing protein and, mRNA expression of

288

peroxisome proliferator receptor-γ (PPAR-γ) and CCAAT element binding protein α (C/EBP-

289

α).112 Regarding UA (2.5–10 µM), this molecule was shown to decrease protein expression of

290

C/EBP-α and -β, PPAR-γ, and SREBP-1c, promoting the increase of ACC phosphorylation, as

291

well as CPT-1 protein expression, and decrease protein contents of FAS and fatty-acid-binding

292

protein-4, through liver kinase B1/AMPK pathway. Therefore, these results also suggested that

293

UA reduces adipogenesis by stimulating fatty acid oxidation, and promoting also the inhibition

294

of fatty acids synthesis.113 Other studies reported that OA (31.7 µM)114 and BA (IC50 9.6 µM)129

295

decrease adiposity by inhibiting the activity of diacylglycerol acetyltransferase, involved in TG

296

deposition in adipocytes. However, OA and UA (25–100 µM) also decreased adiposity by 13 ACS Paragon Plus Environment

123

, an inner mitochondrial

Journal of Agricultural and Food Chemistry

Page 14 of 69

297

increasing lipolysis in fat cells, by promoting hormone-sensitive lipase translocation,126, 127 the

298

rate-limiting enzyme in adipose tissue lipolysis, that hydrolyzes stored triglycerides into

299

monoglycerides and FFAs.130 In vitro studies demonstrated that UA (1–1000 µM) and BA (2.5–

300

10 µM) also promote lipolysis by inhibiting phosphodiesterase activity,108, 128 thus preventing or

301

reversing obesity and insulin resistance in T2DM.

302

Regulation of food intake and body weight:

303

Adipocytes and the gastrointestinal tract release a wide range of hormones, such as adiponectin,

304

resistin, leptin and ghrelin, which regulate appetite and energy metabolism. Altered production

305

or secretion of these hormones caused by excess adipose tissue and body weight contribute to

306

obesity and insulin resistance.131 Therefore, the regulation of adiponectin, resistin, leptin and

307

ghrelin contribute to avoid obesity and to improve insulin sensitivity in T2DM. Oleanolic,

308

ursolic and betulinic acids (5-20 mg/kg/day during 15 days) increased leptin and also decreased

309

the amount of ghrelin synthetized on HFD-induced obesity in mice,32,

310

appetite in this animal model. In contrast, Jia et al. showed that UA (50 and 200 mg/kg)

311

decreases leptin levels on HFD-induced obese mice, as a consequence of the reduction in

312

adipocyte tissue observed in this study. In the same work, it was also reported that UA increases

313

adiponectin,23 regulating lipid and glucose metabolism. Oleanolic acid (1-25 µM) suppressed the

314

production of resistin during adipogenic differentiation of 3T3-L1 adipocytes,132 suggests that

315

this is a mechanism which contributes to the improvement of obesity and insulin resistance in

316

T2DM (Table 2).

317

48, 49

thus regulating

COMORBIDITIES

14 ACS Paragon Plus Environment

Page 15 of 69

Journal of Agricultural and Food Chemistry

318

T2DM leads to the progression of macrovascular and microvascular comorbidities, including

319

atherosclerosis, cardiovascular and hepatic diseases, as well as nephropathy, which are the main

320

risk factors for death among diabetic patients.3 Therefore, the prevention of these related

321

complications may reduce the patients’ mortality risk.

322

Hepatic Diseases:

323

End-stage liver disease is one of the main causes of death in T2DM patients. Among liver

324

diseases in diabetes, non-alcoholic fatty liver disease (NAFLD) is an initial stage of a continuing

325

spectrum of liver alterations, which ultimately can be fatal.133-135 Considering that OA109, 110 and

326

UA22, 23, 111 decrease lipogenesis and stimulate fatty acid β-oxidation in hepatocytes (Table 2), it

327

is possible that both OA and UA provide effective treatment options for NAFLD, thus avoiding

328

progression of disease. In agreement with this, Zhou et al. reported that OA administration (100

329

mg/kg day during 4 weeks) activates AMPK,136 an energy sensor which regulates fatty acid

330

oxidation,137 and reduces lipogenesis and lipid accumulation in the liver.138 Once activated by

331

OA, AMPK induces the inhibition of SREBP and also leads to decrease hepatic ACC, FAS and

332

stearoyl-CoA desaturase-1 (SCD-1) in T2DM mice (Table 3), thus normalizing hepatic lipid

333

metabolism.136 Betulinic acid activates AMPK, which in turn leads to suppression of mRNA

334

expression and nuclear translocation of SREBP-1, repression of FAS and SCD-1 gene

335

expression, and increases PPAR-α and CD36 gene levels in human HepG2 hepatoma cells,

336

primary hepatocytes, and liver tissue from HFD-fed mice. These results suggest that BA can also

337

effectively improve intracellular lipid accumulation in liver cells, suppressing de novo

338

lipogenesis and increasing lipolysis through AMPK pathway activation.125 However, since UA

339

promotes FFA β-oxidation in muscle by activation of AMPK pathway123, a similar effect is

340

expected in the liver. Nonetheless, it is known that UA improves HFD-induced hepatic steatosis 15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 69

341

by regulating lipid metabolism through PPAR-α pathway, decreasing mRNA/protein content for

342

FAT/CD36, SREBP-1c, ACC, and FAS and increasing mRNA/protein content of CPT-1. It was

343

suggested that UA improves insulin resistance by decreasing insulin level and improving the

344

homeostasis model assessment of insulin resistance index (HOMA-IR), by decreasing

345

inflammatory signaling by down-regulating TNF-α, CCL2/MCP-1, IL-1β, IL-2, IL-6 and IL-8,

346

and stimulating the oxidative stress network, increasing SOD, CAT and GPx activities, resulting

347

in decreased MDA content,28 as described in Table 3. Taken all together, those effects may

348

contribute to the improvement of hepatic steatosis and reduction of metabolic dysfunctions on

349

NAFLD.

350

Additionally to hypolipidemic effects, the anti-oxidant activity of TAs may also prevent hepatic

351

dysfunction in diabetes. It has been proposed that the hepatoprotective effects of OA,139 and

352

UA10 are also mediated by a direct ROS scavenging role, i.e. decreasing hydroxyl and

353

superoxide radicals and by increasing the expression/activity of anti-oxidant enzymes, including

354

SOD, CAT, glutathione reductase, GPx and thioredoxin peroxidase (TPx), overall resulting in

355

lipid peroxidation inhibition. Although the mechanisms of anti-oxidant effects are still not

356

clarified, it has been proposed that anti-oxidant and anti-inflammatory effects of OA and UA are

357

mediated, to a great extent, through the activation of Nrf2 transcription, as described above.

358

Nevertheless, the known role of Nrf2 on the regulation of hepatic lipogenesis123, associated to the

359

expression of anti-inflammatory genes, lead us to propose that the activation of Nrf2 pathway by

360

OA and UA may in turn prevent the development of NAFLD by promoting anti-inflammatory

361

and anti-oxidant pathways reducing, at the same time, hepatic lipid accumulation.

16 ACS Paragon Plus Environment

Page 17 of 69

Journal of Agricultural and Food Chemistry

362

Although BA is unable to activate the Nrf2 pathway, it has been evidenced that its anti-oxidant

363

mechanisms responsible for the hepatoprotective activities may also be related with an

364

improvement of the tissue redox system, thus decreasing lipid peroxidation.29

365

Cardiovascular diseases:

366

Diabetes is associated with increased atherosclerosis and the development of cardiovascular

367

disease (CVD). Several studies suggested that OA, UA and BA are potential therapeutic agents

368

for diabetic-related CVD, mainly to prevent the development of atherosclerotic lesions and

369

myocardial dysfunction. Besides their hypolipidemic effects, these TAs have also anti-

370

atherogenic properties. Oleanolic and ursolic acids protected against low density lipoprotein

371

oxidation,140 and reduce pro-inflammatory cytokine production, including NF-kB, TNF-α and

372

IL-1β.24,

373

pathway by OA promoted the reduction of low density lipoprotein oxidation, contributing thus to

374

prevent atherosclerosis123. Other studies suggested that UA exert at least some of its anti-

375

atherogenic effects by preventing the accumulation of inflammatory monocytes in the blood,

376

reducing in vivo monocyte chemotactic activity, thus decreasing plaque size, and macrophage

377

content in atherosclerotic lesions in diabetic mice.25, 26 It was previously shown that BA vascular

378

protection occurs by inhibition of ROS and NF- kB, which consequently reduces the expression

379

of cell adhesion molecules such as vascular cell adhesion molecule-1, intracellular adhesion

380

molecule-1, and endothelial-selectin, contributing to suppression inflammatory responses31,

381

suggesting that the anti-oxidants and anti-inflammatory effects of these compounds are crucial to

382

prevent the development of atherosclerosis. Relevant data on anti-atherogenic effects are

383

summarized in Table 3.

141

According with this, Jiang et al. showed that the activation of Nrf2 signaling

17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 69

384

Another potential mechanism that may delay the development of atherosclerosis may involve a

385

decrease in blood pressure level. The anti-hypertensive effects of OA and UA may be mainly

386

attributed to their potent anti-hyperlipidemic and anti-oxidant effects, combined with diuretic,

387

natriuretic and saluretic effects, mostly due to the inhibition of Na+ and K+ reabsorption in the

388

early portion of the distal tubule.142 However, it was also reported that the anti-hypertensive and

389

anti-atherogenic properties of these compounds may also result from their effects on endothelium

390

dependent vasorelaxation. Impaired endothelial function, due to decreased bioavailability of NO

391

is a hallmark of atherosclerosis. In opposition, an increase in NO induces vasorelaxation,

392

preventing the development of atherosclerosis.143 In this context, UA and BA (0.1-100 µM)

393

displayed vasorelaxing properties, mediated by an increase of NO release and relaxation by

394

activation of endothelial nitric oxide synthase (eNOS).144-146 It was previously described that BA

395

increased the production of NO and the activation of eNOS mediated by the AMPK signaling

396

pathway147, suggesting that this compound used the same pathway to regulate also hepatic

397

glucose production19 (Table 1) and intracellular lipid accumulation125 (Table 2). Nevertheless, it

398

was also supposed that UA and BA (1–10 µM) upregulates eNOS expression, and

399

simultaneously, reduces nicotinamide adenine dinucleotide phosphate (NADPH) oxidase

400

expression in human endothelial cells.148, 149 As NADPH oxidase represents a major source of

401

ROS in CVD,150 this decreased expression may limit vascular oxidative stress, and increase

402

bioactive NO, which contributes to prevent the development of CVD. In opposition, Messner et

403

al. showed that UA (3-50 µM) induced pro-atherogenic effects, promoting DNA-damage, p53-

404

mediated mitochondrial- and caspase-dependent human endothelial cell apoptosis, accelerating

405

atherosclerotic plaque formation in vivo (Table 3). These authors supposed that pro-atherogenic

406

activities observed in this study are exerted at least in part by UA metabolite effects, instead

18 ACS Paragon Plus Environment

Page 19 of 69

Journal of Agricultural and Food Chemistry

407

directly by UA.11 However, further studies are needed to identify their metabolites and clarify

408

these effects.

409

Other studies also evidenced that OA12 and UA151 exhibit cardioprotective effects against

410

hyperglycemia-induced myocardial damage, at least in part due to their anti-oxidant and anti-

411

apoptotic properties. Although the precise mechanisms for these TAs-mediated anti-oxidant

412

effects in the heart are unclear, earlier data described that OA (0.6–1.2 mmol/kg/day during 3

413

days ≈ 274–550 mg/kg/day)152 and UA (35 mg/kg/ day during 8 weeks)151 increase the

414

expression of several enzymatic and non-enzymatic anti-oxidants, including α-tocopherol, GPx,

415

CAT, TPx and SOD in the heart tissue. Oleanolic and ursolic acids are also believed to act as

416

direct free radical scavengers and to reduce lipid peroxidation.153 Thus, it is possible that TAs

417

anti-oxidant effects may prevent cardio-metabolic complications in diabetes. In agreement,

418

Mapanga et al. proposed that OA12 anti-oxidant effects attenuate myocardial apoptosis and

419

thereby leading to improvement in contractile functional recovery following ischemia-

420

reperfusion of hyperglycemic rat hearts. In vitro, OA is a potent cardioprotective agent against

421

high-glucose-induced injury, acting as anti-oxidant, anti-apoptotic and anti-inflammatory

422

agent.13 Although there are no studies concerning UA and BA cardioprotective effects in

423

diabetes, it is known that BA14 and UA154 prevent cardiomyocyte apoptosis.

424

Nephropathy:

425

Diabetic renal injury, or the so-called diabetic nephropathy, is an important diabetic

426

complication, which exacerbate the severity and the mortality of T2DM. Non-enzymatic

427

glycation with formation of Maillard reaction products, also known as advanced glycation end-

428

products (AGEs), including glycated hemoglobin, Nε-(carboxymethyl)lysine (CML) and 19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 69

429

glycated albumin have been implicated in the pathogenesis of diabetic nephropathy.155,

430

Particularly, it is suggested that the accumulation of CML and glycated albumin contributes to

431

the deterioration of diabetic nephropathy.155 Oleanolic acid treatment (100 mg/kg/day during 2

432

weeks) on HFD- and streptozotocin-induced diabetes in mice decreases the loss of glucose in

433

urine and improves kidney structure.157 Other studies also indicated that OA and UA (0.01-0.2%

434

for 10 – 12 weeks) improved kidney function in diabetic mice,158,

435

structure and by inhibiting the polyol pathway. The inhibition of this pathway probably occurred

436

by decreasing renal sorbitol and fructose concentrations,159 and by decreasing renal aldose

437

reductase and renal sorbitol dehydrogenase activities,17, 59, 159 thereby inhibiting AGEs formation

438

and preventing kidney and liver injury. Besides that, these two TAs also decrease plasma HbA1c,

439

pentosidine and CML, as well as urinary glycated albumin in streptozotocin-diabetic mice.159

440

These results were different from those observed in vitro, where OA inhibited formation of

441

pentosidine and CML, but UA only suppressed CML, without affect pentosidine160 (Table 3).

442

Considering also that the oxidative and inflammatory reactions contribute to glycative

443

processes161, the anti-inflammatory and anti-oxidant properties of these three TAs can also

444

contribute to reduce AGE production, although there are no studies evidencing these effects.

445

Despite that, UA was found to decrease the receptor for AGE (RAGE) expression in brain of

446

aging mice162, which subsequently interact with AGEs in the circulation and tissues to form

447

AGE-RAGE complex. As it was described that AGE-RAGE complex generate pro-

448

inflammatory, pro-thrombotic molecules and ROS (reviewed in161), it is expected that the

449

reduction of AGE-RAGE complex contributes to reduce the glycative stress, and delays the

450

development of CVD and diabetic renal injury. It is clear that human studies are needed to

159

156

by preserving kidney

20 ACS Paragon Plus Environment

Page 21 of 69

Journal of Agricultural and Food Chemistry

451

further confirm the anti-glycative effect of these TAs in the progression of development of

452

diabetic co-morbidities.

453

PRESENT AND FUTURE PRESPECTIVES

454

Oleanolic, ursolic and betulinic acids are three TAs that display a wide range of promising anti-

455

diabetic properties, without demonstrating apparent hepatotoxic effects at the dosages

456

administered, contrarily to most anti-diabetic agents currently used. A large number of studies

457

evidenced that OA, UA and BA act as hypoglycemic and anti-obesity agents mainly through (i)

458

reducing the absorption of glucose; (ii) decreasing endogenous glucose production and

459

increasing glycogen synthesis; (iii) increasing insulin sensitivity; (iv) improving lipid

460

homeostasis; and (v) promoting body weight regulation. Besides these promising beneficial

461

effects, OA, UA and BA may also contribute to protect against diabetes comorbidities, such as

462

kidney injury, cardio-metabolic and hepatic diseases (Figure 2). It is interesting to note that many

463

of their multiple anti-diabetic effects occur by mediation of common signaling pathways, which

464

may in turn result in different outcomes. Among multiple effects, it should be noted that AMPK

465

and Nrf2 pathways are involved in many mechanisms of action of these compounds, being

466

AMPK signaling responsible by hypoglycemic, hypolipidemic and anti-atherosclerotic

467

mechanisms of these TAs, while Nrf2 may modulate many of anti-oxidant, anti-inflammatory

468

and hypolipidemic effects of OA and UA, making them important action targets of these

469

compounds.

470

Although the high potential of these three TAs in the treatment and prevention of T2DM is very

471

promising, clinical trials with OA, UA and BA are practically non-existent, with the exception of

472

a 2005 Phase II clinical trial with UA which at the University of Guadalajara, México

473

(#NCT02337933). The trial was performed to evaluate its effects on insulin sensitivity and 21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 69

474

metabolic syndrome, as characterized by overweight/obesity, insulin resistance, hyperglycemia,

475

dyslipidemia and hypertension, and by an inflammatory state, which together increases the risk

476

for CVD or T2DM (University of Guadalajara, México, coded NCT02337933).The results of

477

that clinical trial are still undisclosed. The very limited number of clinical puts into question the

478

true effect of these compounds in humans, highlighting the need to clarify the true efficacy,

479

safety and tolerability of these compounds and the doses required to achieve these anti-diabetic

480

effects of these TAs, before they can be used in T2DM management.

481

Another major limitation in the use of these compounds is their low bioavailability, supported by

482

the fact that the peak plasma concentration after administration of OA163, UA164 and BA165 (10-

483

100 mg/kg) never exceed 1% of the oral dose administered in rats, possibly due to their poor

484

gastrointestinal absorption, limiting their human applications. To improve the bioavailability and

485

efficacy of these compounds, some strategies have been used, either by developing new systems

486

for drug delivery in order to increase the bioavailability of these compounds, or through the

487

developing also of new derivative structures. Numerous efforts have been made to improve the

488

bioavailability of these compounds by developing non-covalent complexes with hydrophilic

489

cyclodextrins, as well as the use of nanosuspensions166. Ursolic acid liposomes (UALs) was until

490

recently the only system approved to be used in a phase I clinical trial (# 2009L00634, China) to

491

improve bioavailability in patients with advanced solid tumors167. This strategy may increase the

492

therapeutic use of these compounds, although its use with OA and BA, as well as their effects in

493

diabetic patients still remains unstudied.

494

Another strategy has been the development of OA168, UA169 and BA84 derivatives to improve anti-

495

hypoglicemic, anti-inflammatory, anti-oxidant effects of these natural compounds. However, the

496

only derivative so far studied in clinical trials was the semi-synthetic OA derivative, bardoxolone

22 ACS Paragon Plus Environment

Page 23 of 69

Journal of Agricultural and Food Chemistry

497

methyl (C-28 methyl ester of 2-cyano-3,12-dioxoolean-1,9-dien-28-oic acid), also designated as

498

CDDO-Me or RTA 402. Bardoxolone methyl successfully completed Phase II clinical trials

499

(coded NCT01574365) for the treatment of patients with type 2 diabetes-associated chronic

500

kidney disease.170,

501

chronic kidney disease were halted due to a higher incidence of cardiovascular events.172 Despite

502

this, development of new TAs derivatives may eventually represent an excellent strategy to treat

503

diabetes in the near future.

504

Apart from these strategies, a very recent study reported also the ability of OA (100 mg/kg day,

505

during 4 weeks) to synergistically potentiate anti-diabetic effects of metformin (250 mg/kg day,

506

during 4 weeks), by improving glucose and insulin homeostasis in diabetic mice.173 It is

507

reasonable to speculate that these compounds may be used to increase the anti-diabetic effects

508

and possibly to counteract hepatotoxicity derived from commonly used therapeutics. However,

509

OA possible hepatotoxic effect should be confirmed, since it has been reported that OA, at

510

higher concentrations (≥ 90 mg/kg/ day), caused cholestatic liver injury in mice.174, 175 Regarding

511

UA and BA, there are no studies reporting their synergistic effects when combined with anti-

512

diabetic agents, neither their potential hepatotoxicity. In this regard, the highly promising effect

513

of these compounds should push clinical assays in order to confirm the efficacy, safety and

514

tolerability of OA, UA and BA individually or in combination with anti-diabetic agents.

515

In summary, this review provides a unique and holistic view on the anti-diabetic properties of

516

OA, UA and BA in the management of T2DM, highlighting the strengths and weaknesses to the

517

future application these compounds in humans. A strong point of the three compounds is their

518

global anti-diabetic effects mediated by specific mechanisms of action, a peculiar characteristic

519

of these compounds which makes them a true asset for treatment and prevention of T2DM.

171

However, Phase III trials (# NCT01351675) with patients in end-stage

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 69

520

Potential weaknesses involve their low bioavailability, and the very limited number of clinical

521

trials. More strategies should be designed to increase the bioavailability of these natural

522

compounds, testing also their effects in clinical trials with T2DM patients. In this perspective, a

523

forcing need is also the clarification of the signaling pathways involved in anti-diabetic effects,

524

searching new therapeutic targets that are involved in Nrf2 and AMPK signaling pathways so

525

that new derivatives with selective anti-diabetic effects can explore those signaling pathways.

526

Therefore, by merging all the information described above, we expected that the development of

527

new strategies to improve the bioavailability and efficacy of these TAs in managing T2DM. An

528

optimal future scenario would be that these compounds could be used as pharmaceutical agents

529

and/or as food supplements for the prevention and/or management of T2DM.

530 531

AUTHOR INFORMATION

532

Corresponding Author

533

*Tel: (+351) 284314399, [email protected]

534 535

Funding sources

536

This work is funded by FEDER funds through the Operational Programme Competitiveness

537

Factors - COMPETE and national funds by FCT - Foundation for Science and Technology under

538

the projects NEucBark– New valorization strategies for Eucalyptus spp. Bark Extracts

539

(PTDC/AGR-FOR/3187/2012), and strategic project UID/NEU/04539/2013 (CNC).

540 541

Notes

542

The authors declare no competing financial interest.

543 24 ACS Paragon Plus Environment

Page 25 of 69

Journal of Agricultural and Food Chemistry

544

ABBREVIATIONS

545

ACC, acetyl-coA carboxylase; ACO, acyl-CoA oxidase; AGEs, advanced glycation end-

546

products; AMPK, 5′ AMP-activated protein kinase; BA, betulinic acid; CAT, catalase; C/EBP-α

547

and –β, CCAAT element binding protein α and β; CCL2/MCP-1, monocyte chemotactic protein-

548

1;

549

cardiovascular disease; eNOS, endothelial nitric oxide synthase; FAS, fatty acid synthase; FFA,

550

free fatty acid; GLP-1, glucagon-like peptide-1; GLUT-2 and -4, glucose transporter -2 and -4;

551

GP, glycogen phosphorylase; G6Pase, glucose-6-phosphatase; GPx, glutathione peroxidase;

552

GSK-3, glycogen synthase kinase-3; hACAT-2, human cholesterol acyltransferase -2; HFD,

553

high-fat diet; HOMA-IR, homeostasis model assessment of insulin resistance index; IR, insulin

554

receptor; IRS-1, insulin receptor substrate -1; MDA, malondialdehyde; NADPH, nicotinamide

555

adenine dinucleotide phosphate; NAFLD, non-alcoholic fatty liver disease; NF-κB, factor

556

nuclear kappa B; Nrf2, nuclear factor erythroid 2-related factor 2; OA, oleanolic acid; PEPCK,

557

phosphoenolpyruvate carboxykinase; PI3KB, phosphatidylinositol 3-kinase-dependent; PPAR-α

558

and γ, peroxisome proliferator activated receptor-α and γ; PTP-1B, protein tyrosine phosphatase

559

1B; RAGE, receptor for AGE; ROS, reactive oxygen species; SCD-1, stearoyl-CoA desaturase-

560

1; SGLT-1, sodium-dependent glucose co-transporter; SOD, superoxide dismutase; SREBP-1c,

561

sterol-regulatory-element-binding protein-1c; TAs, triterpenic acids; T2DM, type 2 diabetes

562

mellitus; TG, triglycerides; TNF-α1, tumor necrosis factor –α1; TPx, thioredoxin peroxidase;

563

UA, ursolic acid, UCP-3, uncoupling protein-3.

CML,

Nε-(carboxymethyl)lysine;

CPT-1,

carnitine

palmitoyltransferase-1;

CVD,

564

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

565

Page 26 of 69

REFERENCES

566

(1) WHO, Diabetes Fact sheet N°312, 2015.

567

(2) Biggs, M. L.; Mukamal, K. J.; Luchsinger, J. A.; Ix, J. H.; Carnethon, M. R.; Newman, A. B.;

568

de Boer, I. H.; Strotmeyer, E. S.; Mozaffarian D.; Siscovick, D. S. Association between

569

adiposity in midlife and older age and risk of diabetes in older adults. Jama 2010, 303, 2504-

570

2512.

571

(3) Forouhi, N. G.; Wareham, N. J. Epidemiology of diabetes. Medicine 2014, 42, 698-702.

572

(4) Domingues, R. M. A.; Guerra, A.R.; Duarte, M., Freire, C. S. R.; Neto, C. P.; Silva, C. M. S.;

573

Silvestre, A.J.D. Bioactive triterpenic acids: from agroforestry biomass residues, to promising

574

therapeutic tools. Mini-Rev. Org. Chem. 2014, 11, 382-399.

575

(5) Camer, D.; Yu, Y.; Szabo, A.; Huang, X. F. The molecular mechanisms underpinning the

576

therapeutic properties of oleanolic acid, its isomer and derivatives for type 2 diabetes and

577

associated complications. Mol. Nutr. Food Res. 2014, 58, 1750-1759.

578

(6) Castellano, J. M.; Guinda, A.; Delgado, T.; Rada, M.; Cayuela, J. A. Biochemical basis of the

579

antidiabetic activity of oleanolic acid and related pentacyclic triterpenes. Diabetes 2013, 62,

580

1791-1799.

581

(7) Alqahtani, A.; Hamid, K.; Kam, A.; Wong, K. H.; Abdelhak, Z.; Razmovski-Naumovski, V.;

582

Chan, K.; Li, K. M.; Groundwater, P. W.; Li, G. Q. The pentacyclic triterpenoids in herbal

583

medicines and their pharmacological activities in diabetes and diabetic complications. Curr.

584

Med. Chem. 2013, 20, 908-931.

585

(8) Abdel-Zaher, A. O.; Abdel-Rahman, M. M.; Hafez, M. M.; Omran, F. M. Role of nitric oxide

586

and reduced glutathione in the protective effects of aminoguanidine, gadolinium chloride and

26 ACS Paragon Plus Environment

Page 27 of 69

Journal of Agricultural and Food Chemistry

587

oleanolic acid against acetaminophen-induced hepatic and renal damage. Toxicology 2007, 234,

588

124-134.

589

(9) Gao, J.; Tang, X.; Dou, H.; Fan, Y.; Zhao, X.; Xu, Q. Hepatoprotective activity of Terminalia

590

catappa L. leaves and its two triterpenoids. J. Pharm. Pharmacol. 2004, 56, 1449-1455.

591

(10) Martin-Aragon, S.; de las Heras, B.; Sanchez-Reus, M. I.; Benedi, Pharmacological

592

modification of endogenous antioxidant enzymes by ursolic acid on tetrachloride-induced liver

593

damage in rats and primary cultures of rat hepatocytes. J. Exp. Toxicol. Pathol. 2001, 53, 199-

594

206.

595

(11) Messner, B.; Zeller, I.; Ploner, C.; Frotschnig, S.; Ringer, T.; Steinacher-Nigisch, A.; Ritsch,

596

A.; Laufer, G.; Huck, C.; Bernhard, D. Ursolic acid causes DNA-damage, p53-mediated,

597

mitochondria- and caspase-dependent human endothelial cell apoptosis, and accelerates

598

atherosclerotic plaque formation in vivo. Atherosclerosis 2011, 219, 402-408.

599

(12) Mapanga, R. F.; Rajamani, U.; Dlamini, N.; Zungu-Edmondson, M.; Kelly-Laubscher, R.;

600

Shafiullah, M.; Wahab, A.; Hasan, M. Y.; Fahim, M. A.; Rondeau, P.; Bourdon, E.; Essop, M. F.

601

Oleanolic acid: a novel cardioprotective agent that blunts hyperglycemia-induced contractile

602

dysfunction. PloS one 2012, 7, e47322.

603

(13) Chan, C. Y.; Mong, M. C.; Liu, W. H.; Huang, C. Y.; Yin, M. C. Three pentacyclic

604

triterpenes protect H9c2 cardiomyoblast cells against high-glucose-induced injury. Free Radical

605

Res. 2014, 48, 402-411.

606

(14) Xia, A.; Xue, Z.; Li, Y.; Wang, W.; Xia, J.; Wei, T.; Cao, J.; Zhou, W. Cardioprotective

607

effect of betulinic Acid on myocardial ischemia reperfusion injury in rats. Evid. Based

608

Complement. Alternat. Med. 2014, 2014, 573745.

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 69

609

(15) Ma, J. Q.; Ding, J.; Xiao, Z. H.; Liu, C. M. Ursolic acid ameliorates carbon tetrachloride-

610

induced oxidative DNA damage and inflammation in mouse kidney by inhibiting the STAT3 and

611

NF-kappaB activities. Int. Immunopharmacol. 2014, 21, 389-395.

612

(16) Khathi, A.; Serumula, M. R.; Myburg, R. B.; Van Heerden, F. R.; Musabayane, C. T.

613

Effects of Syzygium aromaticum-derived triterpenes on postprandial blood glucose in

614

streptozotocin-induced diabetic rats following carbohydrate challenge. PloS one 2013, 8, e81632.

615

(17) Lee, J.; Lee, H. I.; Seo, K. I.; Cho, H. W.; Kim, M. J.; Park, E. M.; Lee, M. K. Effects of

616

ursolic acid on glucose metabolism, the polyol pathway and dyslipidemia in non-obese type 2

617

diabetic mice. Indian J. Exp. Biol., 2014, 52, 683-691.

618

(18) Wang, X.; Liu, R.; Zhang, W.; Zhang, X.; Liao, N.; Wang, Z.; Li, W.; Qin, X.; Hai, C.

619

Oleanolic acid improves hepatic insulin resistance via antioxidant, hypolipidemic and anti-

620

inflammatory effects. Mol. Cell. Endocrinol. 2013, 376, 70-80.

621

(19) Kim, S. J.; Quan, H. Y.; Jeong, K. J.; Kim do, Y.; Kim, G.; Jo, H. K.; Chung, S. H. J.

622

Beneficial effect of betulinic acid on hyperglycemia via suppression of hepatic glucose

623

production. Agric. Food Chem. 2014, 62, 434-442.

624

(20) Sundaresan, A.; Harini, R.; Pugalendi, K. V. Ursolic acid and rosiglitazone combination

625

alleviates metabolic syndrome in high fat diet fed C57BL/6J mice. Gen. Physiol. Biophys. 2012,

626

31, 323-333.

627

(21) Gao, D.; Li, Q.; Li, Y.; Liu, Z.; Fan, Y.; Liu, Z.; Zhao, H.; Li, J.; Han, Z. Antidiabetic and

628

antioxidant effects of oleanolic acid from Ligustrum lucidum Ait in alloxan-induced diabetic

629

rats. Phytother. Res. 2009, 23, 1257-1262.

28 ACS Paragon Plus Environment

Page 29 of 69

Journal of Agricultural and Food Chemistry

630

(22) Sundaresan, A.; Radhiga, T.; Pugalendi, K. V. Effect of ursolic acid and Rosiglitazone

631

combination on hepatic lipid accumulation in high fat diet-fed C57BL/6J mice. Eur. J.

632

Pharmacol. 2014, 741, 297-303.

633

(23) Jia, Y.; Kim, S.; Kim, J.; Kim, B.; Wu, C.; Lee, J. H.; Jun, H.; Kim, N.; Lee, D.; Lee, S.

634

Ursolic acid improves lipid and glucose metabolism in high-fat-fed C57BL/6J mice by activating

635

peroxisome proliferator-activated receptor alpha and hepatic autophagy. Mol. Nutr. Food Res.

636

2014, 344-354.

637

(24) Marquez-Martin, A.; De La Puerta, R.; Fernandez-Arche, A.; Ruiz-Gutierrez, V.; Yaqoob,

638

P. Modulation of cytokine secretion by pentacyclic triterpenes from olive pomace oil in human

639

mononuclear cells. Cytokine 2006, 36, 211-217.

640

(25) Ullevig, S. L.; Zhao, Q.; Zamora, D.; Asmis, R. Ursolic acid protects diabetic mice against

641

monocyte dysfunction and accelerated atherosclerosis. Atherosclerosis 2011, 219, 409-416.

642

(26) Ullevig, S. L.; Kim, H. S.; Nguyen, H. N.; Hambright, W. S.; Robles, A. J.; Tavakoli, S.;

643

Asmis, S. Ursolic acid protects monocytes against metabolic stress-induced priming and

644

dysfunction by preventing the induction of Nox4. Redox Biol. 2014, 2, 259-266.

645

(27) Li, M.; Han, Z.; Bei, W.; Rong, X.; Guo, J.; Hu, X. Oleanolic Acid Attenuates Insulin

646

Resistance via NF-kappaB to Regulate the IRS1-GLUT4 Pathway in HepG2 Cells. Evid. Based

647

Complement. Alternat. Med. 2015, 2015, 643102.

648

(28) Li, S.; Liao, X.; Meng, F.; Wang, Y.; Sun, Z.; Guo, F.; Li, X.; Meng, M.; Li, Y.; Sun, C.

649

Therapeutic role of ursolic acid on ameliorating hepatic steatosis and improving metabolic

650

disorders in high-fat diet-induced non-alcoholic fatty liver disease rats. PloS one, 2014, 9,

651

e86724.

29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 69

652

(29) Yi, J. ; Xia, W.; Wu, J.; Yuan, L.; Wu, J.; Tu, D.; Fang, J.; Tan, Z. Betulinic acid prevents

653

alcohol-induced liver damage by improving the antioxidant system in mice. J. Vet. Sci. 2014, 15,

654

141-148.

655

(30) Chun, J.; Lee, C.; Hwang, S. W.; Im, J. P.; Kim, J. S. Ursolic acid inhibits nuclear factor-

656

kappaB signaling in intestinal epithelial cells and macrophages, and attenuates experimental

657

colitis in mice. Life Sci., 2014, 110, 23-34.

658

(31) Yoon, J. J.; Lee, Y. J.; Kim, J. S.; Kang, D. G.; Lee, H. S. Protective role of betulinic acid

659

on TNF-alpha-induced cell adhesion molecules in vascular endothelial cells. Biochem. Biophys.

660

Res. Commun. 2010, 391, 96-101.

661

(32) de Melo, C. L.; Queiroz, M. G.; Fonseca, S. G.; Bizerra, A. M.; Lemos, T. L.; Melo, T. S.;

662

Santos, T. S.; Rao, V. S. Oleanolic acid, a natural triterpenoid improves blood glucose tolerance

663

in normal mice and ameliorates visceral obesity in mice fed a high-fat diet. Chem.-Biol. Interact.

664

2010, 185, 59-65.

665

(33) Chitturi, S.; George, J. Hepatotoxicity of commonly used drugs: nonsteroidal anti-

666

inflammatory drugs, antihypertensives, antidiabetic agents, anticonvulsants, lipid-lowering

667

agents, psychotropic drugs. Semin. Liver Dis. 2002, 22, 169-183.

668

(34) Scheen, A. J. Pharmacokinetic and toxicological considerations for the treatment of diabetes

669

in patients with liver disease. Expert Opin. Drug Metab. Toxicol. 2014, 10, 839-857.

670

(35) Yim, T. K.; Wu, W. K.; Pak, W. F.; Ko, K. M. Hepatoprotective action of an oleanolic acid-

671

enriched extract of Ligustrum lucidum fruits is mediated through an enhancement on hepatic

672

glutathione regeneration capacity in mice. Phytother. Res. 2001, 15, 589-592.

30 ACS Paragon Plus Environment

Page 31 of 69

Journal of Agricultural and Food Chemistry

673

(36) Jeong, H. G. Inhibition of cytochrome P450 2E1 expression by oleanolic acid:

674

hepatoprotective effects against carbon tetrachloride-induced hepatic injury. Toxicol Lett. 1999,

675

105, 215-222.

676

(37) Bai, X.; Qiu, A.; Guan, J.; Shi, Z. Antioxidant and protective effect of an oleanolic acid-

677

enriched extract of A. deliciosa root on carbon tetrachloride induced rat liver injury. Asia Pac. J.

678

Clin. Nutr. 2007, 16 Suppl 1, 169-173.

679

(38) Liu, J.; Liu, Y.; Klaassen, C. D. Protective effect of oleanolic acid against chemical-induced

680

acute necrotic liver injury in mice. Zhongguo yao li xue bao, 1995, 16, 97-102.

681

(39) Liu, J.; Liu, Y.; Mao, Q.; Klaassen, C. D. The effects of 10 triterpenoid compounds on

682

experimental liver injury in mice. Fundam. Appl. Toxicol. 1994, 22, 34-40.

683

(40) Ma, J. Q.; Ding, J.; Zhang, L.; Liu, C. M. Ursolic acid protects mouse liver against CCl4-

684

induced oxidative stress and inflammation by the MAPK/NF-kappaB pathway. Environ. Toxicol.

685

Pharmacol. 2014, 37, 975-983.

686

(41) Ma, J. Q.; Ding, J.; Zhang, L.; Liu, C. M. Protective effects of ursolic acid in an

687

experimental model of liver fibrosis through Nrf2/ARE pathway. Clin. Res. Hepatol.

688

Gastroenterol. 2015, 39, 188-197.

689

(42) Zheng, Z. W.; Song, S. Z.; Wu, Y. L.; Lian, L. H.; Wan, Y.; Nan, J. X. Betulinic acid

690

prevention of d-galactosamine/lipopolysaccharide liver toxicity is triggered by activation of Bcl-

691

2 and antioxidant mechanisms. J. Pharm. Pharmacol. 2011, 63, 572-578.

692

(43) Kellett, G. L.; Brot-Laroche, E. Apical GLUT2: a major pathway of intestinal sugar

693

absorption. Diabetes 2005, 54, 3056-3062.

31 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 69

694

(44) Van de Laar, F. A.; Lucassen, P. L.; Akkermans, R. P.; Van de Lisdonk, E. H.; Rutten , G.

695

E.; Van Weel, C. Alpha-glucosidase inhibitors for type 2 diabetes mellitus. Cochrane Database

696

Syst. Rev. 2005, CD003639.

697

(45) Kumar, S.; Narwal, S.; Kumar, V.; Prakash, O.; alpha-glucosidase inhibitors from plants: A

698

natural approach to treat diabetes. Pharmacogn. Rev. 2011, 5, 19-29.

699

(46) Ali, H.; Houghton, P. J.; Soumyanath, A. alpha-Amylase inhibitory activity of some

700

Malaysian plants used to treat diabetes; with particular reference to Phyllanthus amarus. J.

701

Ethnopharmacol. 2006, 107, 449-455.

702

(47) Kumar, S.; Kumar, V.; Prakash, O.; Enzymes inhibition and antidiabetic effect of isolated

703

constituents from Dillenia indica. BioMed Res. Int. 2013, 2013, 382063.

704

(48) Rao, V. S.; de Melo, C. L.; Queiroz, M. G.; Lemos, T. L.; Menezes, D. B.; Melo, T. S.;

705

Santos, F. A.; Ursolic acid, a pentacyclic triterpene from Sambucus australis, prevents abdominal

706

adiposity in mice fed a high-fat diet. J. Med. Food 2011, 14, 1375-1382.

707

(49) de Melo, C. L.; Queiroz, M. G.; Arruda Filho, A. C.; Rodrigues, A. M.; de Sousa, D. F.;

708

Almeida, J. G.; Pessoa, O. D.; Silveira, E. R.; Menezes, D. B.; Melo, T. S.; Santos, F. A.; Rao,

709

V. S. Betulinic acid, a natural pentacyclic triterpenoid, prevents abdominal fat accumulation in

710

mice fed a high-fat diet. J. Agric. Food Chem. 2009, 57, 8776-8781.

711

(50) Jabeen, B.; Riaz, N.; Saleem, M.; Naveed, M. A.; Ashraf, M.; Alam, U.; Rafiq, H. M.;

712

Tareen, R. B.; Jabbar, A.; Isolation of natural compounds from Phlomis stewartii showing alpha-

713

glucosidase inhibitory activity. Phytochemistry 2013, 96, 443-448.

714

(51) Rani, M. P.; Raghu, K. G.; Nair, M. S.; Padmakumari, K. P. Isolation and identification of

715

alpha-glucosidase and protein glycation inhibitors from Stereospermum colais. Appl. Biochem.

716

Biotechnol. 2014, 173, 946-956. 32 ACS Paragon Plus Environment

Page 33 of 69

Journal of Agricultural and Food Chemistry

717

(52) Wang, Z. W.; Wang, J. S.; Luo, J.; Kong, L. Y. alpha-Glucosidase inhibitory triterpenoids

718

from the stem barks of Uncaria laevigata. Fitoterapia 2013, 90, 30-37.

719

(53) He, Q. Q.; Yang, L.; Zhang, J. Y.; Ma, J. N.; Ma, C. M. Chemical constituents of gold-red

720

apple and their alpha-glucosidase inhibitory activities. J. Food Sci. 2014, 79, C1970-1983.

721

(54) Salah El Dine, R.; Ma, Q.; Kandil, Z. A.; El-Halawany, A. M. Triterpenes as uncompetitive

722

inhibitors of alpha-glucosidase from flowers of Punica granatum L. Nat. Prod. Res. 2014, 28,

723

2191-2194.

724

(55) Komaki, E.; Yamaguchi, S.; Maru, I.; Kinoshita, M.; Kakehi, K.; Ohta, Y.; Tsukada, Y.

725

Identification of Anti-α-Amylase Components from Olive Leaf Extracts. Food Sci. Technol. Res.

726

2003, 9, 35-39.

727

(56) Hundal, R. S.; Krssak, M.; Dufour, S.; Laurent, D.; Lebon, V.; Chandramouli, V.; Inzucchi,

728

S. E.; Schumann, W. C.; Petersen, K. F.; Landau, B. R.; Shulman, G. I. Mechanism by which

729

metformin reduces glucose production in type 2 diabetes. Diabetes 2000, 49, 2063-2069.

730

(57) Wen, X.; Sun, H.; Liu, J.; Cheng, K.; Zhang, P.; Zhang, L.; Hao, J.; Zhang, L.; Ni, P.;

731

Zographos, S. E.; Leonidas, D. D.; Alexacou, K. M.; Gimisis, T.; Hayes, J. M.; Oikonomakos, N.

732

G. Naturally occurring pentacyclic triterpenes as inhibitors of glycogen phosphorylase: synthesis,

733

structure-activity relationships, and X-ray crystallographic studies. J. Med. Chem. 2008, 51,

734

3540-3554.

735

(58) Aiston, S.; Coghlan, M. P.; Agius, L. Inactivation of phosphorylase is a major component of

736

the mechanism by which insulin stimulates hepatic glycogen synthesis. Eur. J. Biochem. 2003,

737

270, 2773-2781.

33 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 34 of 69

738

(59) Jang, S. M.; Kim, M. J.; Choi, M. S.; Kwon, E. Y.; Lee, M. K. Inhibitory effects of ursolic

739

acid on hepatic polyol pathway and glucose production in streptozotocin-induced diabetic mice.

740

Metab. Clin. Exp. 2010, 59, 512-519.

741

(60) Postic, C.; Shiota, M.; Niswender, K. D.; Jetton, T. L.; Chen, Y.; Moates, J. M.; Shelton, K.

742

D.; Lindner, J.; Cherrington, A. D.; Magnuson, M. A. Dual roles for glucokinase in glucose

743

homeostasis as determined by liver and pancreatic beta cell-specific gene knock-outs using Cre

744

recombinase. J. Biol. Chem., 1999, 274, 305-315.

745

(61) Foster, J. D.; Pederson, B. A.; Nordlie, R. C. Glucose-6-phosphatase structure, regulation,

746

and function: an update. Proc. Soc. Exp. Biol. Med. 1997, 215, 314-332.

747

(62) Azevedo, M. F.; Camsari, C.; Sa, C. M.; Lima, C. F.; Fernandes-Ferreira, M.; Pereira-

748

Wilson, C. Ursolic acid and luteolin-7-glucoside improve lipid profiles and increase liver

749

glycogen content through glycogen synthase kinase-3. Phytother. Res. 2010, 24 Suppl 2, S220-

750

224.

751

(63) Lee, J.; Kim, M. S. The role of GSK3 in glucose homeostasis and the development of

752

insulin resistance. Diabetes Res. Clin. Pract. 2007, 77 Suppl 1, S49-57.

753

(64) Ha do, T.; Tuan, D. T.; Thu, N. B.; Nhiem, N. X.; Ngoc, T. M.; Yim, N.; Bae, K. Palbinone

754

and triterpenes from Moutan Cortex (Paeonia suffruticosa, Paeoniaceae) stimulate glucose

755

uptake and glycogen synthesis via activation of AMPK in insulin-resistant human HepG2 Cells.

756

Bioorg. Med. Chem. Lett., 2009, 19, 5556-5559.

757

(65) Viollet, B.; Foretz, M.; Guigas, B.; Horman, S.; Dentin, R.; Bertrand, L.; Hue, L.; Andreelli,

758

F. Activation of AMP-activated protein kinase in the liver: a new strategy for the management of

759

metabolic hepatic disorders. J. Physiol. 2006, 574, 41-53.

34 ACS Paragon Plus Environment

Page 35 of 69

Journal of Agricultural and Food Chemistry

760

(66) Heiss, E. H.; Kramer, M. P.; Atanasov, A. G.; Beres, H.; Schachner, D.; Dirsch, V. M.

761

Glycolytic switch in response to betulinic Acid in non-cancer cells. PloS one, 2014, 9, e115683.

762

(67) Gerich, J. E. Redefining the clinical management of type 2 diabetes: matching therapy to

763

pathophysiology. Eur. J. Clin. Invest. 2002, 32 Suppl 3, 46-53.

764

(68) Jang, S. M.; Yee, S. T.; Choi, J.; Choi, M. S.; Do, G. M.; Jeon, S. M.; Yeo, J.; Kim, M. J.;

765

Seo, K. I.; Lee, M. K. Ursolic acid enhances the cellular immune system and pancreatic beta-cell

766

function in streptozotocin-induced diabetic mice fed a high-fat diet. Int. Immunopharmacol.

767

2009, 9, 113-119.

768

(69) Jayaprakasam, B.; Olson, L. K.; Schutzki, R. E.; Tai, M. H.; Nair, M. G. Amelioration of

769

obesity and glucose intolerance in high-fat-fed C57BL/6 mice by anthocyanins and ursolic acid

770

in Cornelian cherry (Cornus mas). J. Agric. Food Chem. 2006, 54, 243-248.

771

(70) Teodoro, T.; Zhang, L.; Alexander, T.; Yue, J.; Vranic, M.; Volchuk, A. Oleanolic acid

772

enhances insulin secretion in pancreatic beta-cells. FEBS Lett. 2008, 582, 1375-1380.

773

(71) Zhang, Y.; Jayaprakasam, B.; Seeram, N. P.; Olson, L. K.; DeWitt, D.; Nair, M. G. Insulin

774

secretion and cyclooxygenase enzyme inhibition by cabernet sauvignon grape skin compounds.

775

J. Agric. Food Chem. 2004, 52, 228-233.

776

(72) Castro, A. J.; Frederico, M. J.; Cazarolli, L. H.; Bretanha, L. C.; Tavares Lde, C.; Buss Zda,

777

S.; Dutra, M. F.; de Souza, A. Z.; Pizzolatti, M. G.; Silva, F. R. Betulinic acid and 1,25(OH)(2)

778

vitamin D(3) share intracellular signal transduction in glucose homeostasis in soleus muscle. Int.

779

J. Biochem. Cell Biol. 2014, 48, 18-27.

780

(73) Castro, A. J.; Frederico, M. J.; Cazarolli, L. H.; Mendes, C. P.; Bretanha, L. C.; Schmidt, E.

781

C.; Bouzon, Z. L.; de Medeiros Pinto, V. A.; da Fonte Ramos, C.; Pizzolatti, M. G.; Silva, F. R.

782

The mechanism of action of ursolic acid as insulin secretagogue and insulinomimetic is mediated

35 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 36 of 69

783

by cross-talk between calcium and kinases to regulate glucose balance. Biochim. Biophys. Acta

784

2014, 51-61.

785

(74) Zhang, Y. N.; Zhang, W.; Hong, D.; Shi, L.; Shen, Q.; Li, J. Y.; Li, J.; Hu, L. H. Oleanolic

786

acid and its derivatives: new inhibitor of protein tyrosine phosphatase 1B with cellular activities.

787

Bioorganic & medicinal chemistry 2008, 16, 8697-8705.

788

(75) Ramirez-Espinosa, J. J.; Rios, M. Y.; Lopez-Martinez, S.; Lopez-Vallejo, F.; Medina-

789

Franco, J. L.; Paoli, P.; Camici, G.; Navarrete-Vazquez, G.; Ortiz-Andrade, R.; Estrada-Soto, S.

790

Antidiabetic activity of some pentacyclic acid triterpenoids, role of PTP-1B: in vitro, in silico,

791

and in vivo approaches. Eur. J. Med. Chem. 2011, 46, 2243-2251.

792

(76) Na, M.; Yang, S.; He, L.; Oh, H.; Kim, B. S.; Oh, W. K.; Kim, B. Y.; Ahn, J. S. Inhibition

793

of protein tyrosine phosphatase 1B by ursane-type triterpenes isolated from Symplocos

794

paniculata. Planta Med. 2006, 72, 261-263.

795

(77) Zhang, W.; Hong, D.; Zhou, Y.; Zhang, Y.; Shen, Q.; Li, J. Y.; Hu, L. H.; Li, J. Ursolic acid

796

and its derivative inhibit protein tyrosine phosphatase 1B, enhancing insulin receptor

797

phosphorylation and stimulating glucose uptake. Biochim. Biophys. Acta 2006, 1760, 1505-1512.

798

(78) Li, D.; Li, W.; Higai, K.; Koike, K. Protein tyrosine phosphatase 1B inhibitory activities of

799

ursane- and lupane-type triterpenes from Sorbus pohuashanensis. J. Nat. Med. 2014, 68, 427-

800

431.

801

(79) Choi, J. Y.; Na, M.; Hyun Hwang, I.; Ho Lee, S.; Young Bae, E.; Yeon Kim, B.; Seog Ahn,

802

J. Isolation of betulinic acid, its methyl ester and guaiane sesquiterpenoids with protein tyrosine

803

phosphatase 1B inhibitory activity from the roots of Saussurea lappa C.B.Clarke. Molecules

804

2009, 14, 266-272.

36 ACS Paragon Plus Environment

Page 37 of 69

Journal of Agricultural and Food Chemistry

805

(80) Jung, S. H.; Ha, Y. J.; Shim, E. K.; Choi, S. Y.; Jin, J. L.; Yun-Choi, H. S.; Lee, J. R.

806

Insulin-mimetic and insulin-sensitizing activities of a pentacyclic triterpenoid insulin receptor

807

activator. Biochem. J. 2007, 403, 243-250.

808

(81) Li, Y.; Wang, J.; Gu, T.; Yamahara J.; Li, Y. Oleanolic acid supplement attenuates liquid

809

fructose-induced adipose tissue insulin resistance through the insulin receptor substrate-

810

1/phosphatidylinositol 3-kinase/Akt signaling pathway in rats. Toxicol. Appl. Pharmacol. 2014,

811

277, 155-163.

812

(82) He, Y.; Li, W.; Li, Y.; Zhang, S.; Wang, Y.; Sun, C. Ursolic acid increases glucose uptake

813

through the PI3K signaling pathway in adipocytes. PloS one 2014, 9, e110711.

814

(83) Sato, H.; Genet, C.; Strehle, A.; Thomas, C.; Lobstein, A.; Wagner, A.; Mioskowski, C.;

815

Auwerx, J.; Saladin, R. Anti-hyperglycemic activity of a TGR5 agonist isolated from Olea

816

europaea. Biochem. Biophys. Res. Commun. 2007, 362, 793-798.

817

(84) Genet, C.; Strehle, A.; Schmidt, C.; Boudjelal, G.; Lobstein, A.; Schoonjans, K.; Souchet,

818

M.; Auwerx, J.; Saladin, R.; Wagner, A. Structure-activity relationship study of betulinic acid, a

819

novel and selective TGR5 agonist, and its synthetic derivatives: potential impact in diabetes. J.

820

Med. Chem. 2010, 53, 178-190.

821

(85) Bala, V.; Rajagopal, S.; Kumar, D. P.; Nalli, A. D.; Mahavadi, S.; Sanyal, A. J.; Grider, J.

822

R.; Murthy, K. S. Release of GLP-1 and PYY in response to the activation of G protein-coupled

823

bile acid receptor TGR5 is mediated by Epac/PLC-epsilon pathway and modulated by

824

endogenous H2S. Front. Physiol 2014, 5, 420.

825

(86) Kumar, D. P.; Rajagopal, S.; Mahavadi, S.; Mirshahi, F.; Grider, J. R.; Murthy, K. S.;

826

Sanyal, A. J. Activation of transmembrane bile acid receptor TGR5 stimulates insulin secretion

827

in pancreatic beta cells. Biochem. Biophys. Res. Commun. 2012, 427, 600-605.

37 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 38 of 69

828

(87) Nataraju, A.; Saini, D.; Ramachandran, S.; Benshoff, N.; Liu, W.; Chapman W.;

829

Mohanakumar, T. Oleanolic Acid, a plant triterpenoid, significantly improves survival and

830

function of islet allograft. Transplantation 2009, 88, 987-994.

831

(88) Norris, K.; Norris, F.; Kono, D. H.; Vestergaard, H.; Pedersen, O.; Theofilopoulos, A. N.;

832

Moller, N. P. Expression of protein-tyrosine phosphatases in the major insulin target tissues.

833

FEBS Lett. 1997, 415, 243-248.

834

(89) Asante-Appiah, E.; Kennedy, B. P. Protein tyrosine phosphatases: the quest for negative

835

regulators of insulin action. Am. J. Physiol.: Endocrinol. Metab. 2003, 284, E663-670.

836

(90) Galic, S.; Hauser, C.; Kahn, B. B.; Haj, F. G.; Neel, B. G.; Tonks, N. K.; Tiganis, T.

837

Coordinated regulation of insulin signaling by the protein tyrosine phosphatases PTP1B and

838

TCPTP. Mol. Cell. Biol. 2005, 25, 819-829.

839

(91) Nakagawa, Y.; Aoki, N.; Aoyama, K.; Shimizu, H.; Shimano, H.; Yamada, N.; Miyazaki, H.

840

Receptor-type protein tyrosine phosphatase epsilon (PTPepsilonM) is a negative regulator of

841

insulin signaling in primary hepatocytes and liver. Zool. Sci. 2005, 22, 169-175.

842

(92) White, M. F. The IRS-signalling system: a network of docking proteins that mediate insulin

843

action. Mol. Cell. Biochem. 1998, 182, 3-11.

844

(93) Virkamaki, A.; Ueki, K.; Kahn, C. R. Protein-protein interaction in insulin signaling and the

845

molecular mechanisms of insulin resistance. J. Clin. Invest. 1999, 103, 931-943.

846

(94) Trauner, M.; Claudel, T.; Fickert, P.; Moustafa, T.; Wagner, M. Bile acids as regulators of

847

hepatic lipid and glucose metabolism. Dig. Dis. 2010, 28, 220-224.

848

(95) MacDonald, P. E.; El-Kholy, W.; Riedel, M. J.; Salapatek, A. M.; Light, P. E.; Wheeler, M.

849

B. The multiple actions of GLP-1 on the process of glucose-stimulated insulin secretion.

850

Diabetes 2002, 51 Suppl 3, S434-442.

38 ACS Paragon Plus Environment

Page 39 of 69

Journal of Agricultural and Food Chemistry

851

(96) Buteau, J.; Foisy, S.; Joly, E.; Prentki, M. Glucagon-like peptide 1 induces pancreatic beta-

852

cell proliferation via transactivation of the epidermal growth factor receptor. Diabetes 2003, 52,

853

124-132.

854

(97) Mojsov, S. Glucagon-like Peptide-1 (GLP-1) and the Control of Glucose Metabolism in

855

Mammals and Teleost Fish. Amer. Zool 2000, 40, 246-258.

856

(98) Puddu, A.; Sanguineti, R.; Mach, F.; Dallegri, F.; Viviani, G. L.; Montecucco, F. Update on

857

the protective molecular pathways improving pancreatic beta-cell dysfunction. Mediators of

858

inflammation 2013, 2013, 750540.

859

(99) Lee, W.; Yang, E. J.; Ku, S. K.; Song, K. S.; Bae, J. S. Anti-inflammatory effects of

860

oleanolic acid on LPS-induced inflammation in vitro and in vivo. Inflammation 2013, 36, 94-

861

102.

862

(100) Qatanani, M.; Lazar, M. A.; Mechanisms of obesity-associated insulin resistance: many

863

choices on the menu. Genes Dev. 2007, 21, 1443-1455.

864

(101) Hybertson, B. M.; Gao, B.; Bose, S. K.; McCord, J. M. Oxidative stress in health and

865

disease: the therapeutic potential of Nrf2 activation. Mol. Aspects Med. 2011, 32, 234-246.

866

(102) Zhang, Z.; Zhou, S.; Jiang, X.; Wang, Y. H.; Li, F.; Wang, Y. G.; Zheng, Y.; Cai, L. The

867

role of the Nrf2/Keap1 pathway in obesity and metabolic syndrome. Rev. Endocr. Metab.

868

Disord. 2015, 16, 35-45.

869

(103) Reisman, S. A.; Aleksunes, L. M.; Klaassen, C. D. Oleanolic acid activates Nrf2 and

870

protects from acetaminophen hepatotoxicity via Nrf2-dependent and Nrf2-independent

871

processes. Biochem. Pharmacol. 2009, 77, 1273-1282.

39 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 40 of 69

872

(104) Liu, J.; Wu, Q.; Lu, Y. F.; Pi, J. New insights into generalized hepatoprotective effects of

873

oleanolic acid: key roles of metallothionein and Nrf2 induction. Biochem. Pharmacol. 2008, 76,

874

922-928.

875

(105) Lu, Y. F.; Liu, J.; Wu, K. C.; Klaassen, C. D. Protection against phalloidin-induced liver

876

injury by oleanolic acid involves Nrf2 activation and suppression of Oatp1b2. Toxicol. Lett.

877

2014, 232, 326-332.

878

(106) Gao, D.; Li, Q.; Li, Y.; Liu, Z.; Liu, Z.; Fan, Y.; Han, Z.; Li, J.; Li, K. Antidiabetic

879

potential of oleanolic acid from Ligustrum lucidum Ait. Can. J. Physiol. Pharmacol. 2007, 85,

880

1076-1083.

881

(107) Jang, D. S.; Lee, G. Y.; Kim, J.; Lee, Y. M.; Kim, J. M.; Kim, Y. S.; Kim, J. S. A new

882

pancreatic lipase inhibitor isolated from the roots of Actinidia arguta. Arch. Pharmacal Res.

883

2008, 31, 666-670.

884

(108) Kim, J.; Lee, Y. S.; Kim, C. S.; Kim, J. S. Betulinic acid has an inhibitory effect on

885

pancreatic lipase and induces adipocyte lipolysis. Phytother. Res. 2012, 26, 1103-1106.

886

(109) Liu, C.; Li, Y.; Zuo, G.; Xu, W.; Gao, H.; Yang, Y.; Yamahara, J.; Wang, J.; Li, Y.

887

Oleanolic Acid diminishes liquid fructose-induced Fatty liver in rats: role of modulation of

888

hepatic sterol regulatory element-binding protein-1c-mediated expression of genes responsible

889

for de novo Fatty Acid synthesis. Evid. Based Complement. Alternat. Med. 2013, 2013, 534084.

890

(110) Yunoki, K.; Sasaki, G.; Tokuji, Y.; Kinoshita, M.; Naito, A.; Aida, K.; Ohnishi, M. Effect

891

of dietary wine pomace extract and oleanolic acid on plasma lipids in rats fed high-fat diet and

892

its DNA microarray analysis. J. Agric. Food Chem. 2008, 56, 12052-12058.

40 ACS Paragon Plus Environment

Page 41 of 69

Journal of Agricultural and Food Chemistry

893

(111) Jia, Y.; Bhuiyan, M. J.; Jun, H. J.; Lee, J. H.; Hoang, M. H.; Lee, H. J.; Kim, N.; Lee, D.;

894

Hwang, K. Y.; Hwang, B. Y.; Choi, D. W.; Lee, S. J. Ursolic acid is a PPAR-alpha agonist that

895

regulates hepatic lipid metabolism. Bioorg. Med. Chem. Lett. 2011, 21, 5876-5880.

896

(112) Sung, H. Y.; Kang, S. W.; Kim, J. L.; Li, J.; Lee, E. S.; Gong, J. H.; Han, S. J.; Kang, Y.

897

H. Oleanolic acid reduces markers of differentiation in 3T3-L1 adipocytes. Nutr. Res. 2010, 30,

898

831-839.

899

(113) He, Y.; Li, Y.; Zhao, T.; Wang, Y.; Sun, C. Ursolic acid inhibits adipogenesis in 3T3-L1

900

adipocytes through LKB1/AMPK pathway. PloS one 2013, 8, e70135.

901

(114) Dat, N. T.; Cai, X. F.; Rho, M. C.; Lee, H. S.; Bae, K.; Kim, Y. H. The inhibition of

902

diacylglycerol acyltransferase by terpenoids from Youngia koidzumiana. Arch. Pharmacal Res.

903

2005, 28, 164-168.

904

(115) Lee, W. S.; Im, K. R.; Park, Y. D.; Sung, N. D.; Jeong, T. S. Human ACAT-1 and ACAT-

905

2 inhibitory activities of pentacyclic triterpenes from the leaves of Lycopus lucidus TURCZ.

906

Biol. Pharm. Bull. 2006, 29, 382-384.

907

(116) Rudel, L. L.; Lee, R. G.; Cockman, T. L. Acyl coenzyme A: cholesterol acyltransferase

908

types 1 and 2: structure and function in atherosclerosis. Curr. Opin. Lipidol. 2001, 12, 121-127.

909

(117) Beysen, C.; Murphy, E. J.; Nagaraja, H.; Decaris, M.; Riiff, T.; Fong, A.; Hellerstein, M.

910

K.; Boyle, P. J. A pilot study of the effects of pioglitazone and rosiglitazone on de novo

911

lipogenesis in type 2 diabetes. J. Lipid Res. 2008, 49, 2657-2663.

912

(118) Dong, J.; Xu, H.; Xu, H.; Wang, P. F.; Cai, G. J.; Song, H. F.; Wang, C. C.; Dong, Z. T.;

913

Ju, Y. J.; Jiang, Z. Y.; Nesfatin-1 stimulates fatty-acid oxidation by activating AMP-activated

914

protein kinase in STZ-induced type 2 diabetic mice. PloS one 2013, 8, e83397.

41 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 42 of 69

915

(119) Murase, T.; Misawa, K.; Minegishi, Y.; Aoki, M.; Ominami, H.; Suzuki, Y.; Shibuya, Y.;

916

Hase, T. Coffee polyphenols suppress diet-induced body fat accumulation by downregulating

917

SREBP-1c and related molecules in C57BL/6J mice. Am. J. Physiol.: Endocrinol. Metab. 2011,

918

300, E122-133.

919

(120) Zou, M. H.; Hou, X. Y.; Shi, C. M.; Kirkpatick, S.; Liu, F.; Goldman, M. H.; Cohen, R. A.

920

Activation of 5'-AMP-activated kinase is mediated through c-Src and phosphoinositide 3-kinase

921

activity during hypoxia-reoxygenation of bovine aortic endothelial cells. Role of peroxynitrite. J.

922

Biol. Chem. 2003, 278, 34003-34010.

923

(121) Vazquez, M.; Roglans, N.; Cabrero, A., Rodriguez, C.; Adzet, T.; Alegret, M.; Sanchez, R.

924

M.; Laguna, J. C. Bezafibrate induces acyl-CoA oxidase mRNA levels and fatty acid

925

peroxisomal beta-oxidation in rat white adipose tissue. Mol. Cell. Biochem. 2001, 216, 71-78.

926

(122) Tanaka, T.; Masuzaki, H.; Ebihara, K.; Ogawa, Y.; Yasue, S.; Yukioka, H.; Chusho, H.;

927

Miyanaga, F.; Miyazawa, T.; Fujimoto, M.; Kusakabe, T.; Kobayashi, N.; Hayashi, T.; Hosoda,

928

K.; Nakao, K.; Transgenic expression of mutant peroxisome proliferator-activated receptor

929

gamma in liver precipitates fasting-induced steatosis but protects against high-fat diet-induced

930

steatosis in mice. Metab. Clin. Exp. 2005, 54, 1490-1498.

931

(123) Chu, X.; He, X.; Shi, Z.; Li, C.; Guo, F.; Li, S.; Li, Y.; Na, L.; Sun, C. Ursolic acid

932

increases energy expenditure through enhancing free fatty acid uptake and beta-oxidation via an

933

UCP3/AMPK-dependent pathway in skeletal muscle. Mol. Nutr. Food Res. 2015, 59, 1491-1503.

934

(124) Rousset, S.; Alves-Guerra, M. C.; Mozo, J.; Miroux, B.; Cassard-Doulcier, A. M.;

935

Bouillaud, F.; Ricquier, D. The biology of mitochondrial uncoupling proteins. Diabetes 2004, 53

936

Suppl 1, S130-135.

42 ACS Paragon Plus Environment

Page 43 of 69

Journal of Agricultural and Food Chemistry

937

(125) Quan, H. Y.; Kim do, Y.; Kim, S. J.; Jo, H. K.; Kim, G. W.; Chung, S. H. Betulinic acid

938

alleviates non-alcoholic fatty liver by inhibiting SREBP1 activity via the AMPK-mTOR-SREBP

939

signaling pathway. Biochem. Pharmacol. 2013, 85, 1330-1340.

940

(126) Kong, T.; Li, S.; Li, Y.; Sun, C. [Oleanolic acid-stimulated lipolysis in primary adipocytes

941

and its mechanisms]. Weisheng Yanjiu, 2011, 40, 27-29.

942

(127) Li, Y.; Kang, Z.; Li, S.; Kong, T.; Liu, X.; Sun, C. Ursolic acid stimulates lipolysis in

943

primary-cultured rat adipocytes. Mol. Nutr. Food Res. 2010, 54, 1609-1617.

944

(128) Kim, J.; Jang, D. S.; Kim, H.; Kim, J. S. Anti-lipase and lipolytic activities of ursolic acid

945

isolated from the roots of Actinidia arguta. Arch. Pharmacal Res. 2009, 32, 983-987.

946

(129) Chung, M. Y.; Rho, M. C.; Lee, S. W.; Park, H. R.; Kim, K.; Lee, I. A.; Kim, D. H.; Jeune,

947

K. H.; Lee, H. S.; Kim, Y. K. Inhibition of diacylglycerol acyltransferase by betulinic acid from

948

Alnus hirsuta. Planta Med. 2006, 72, 267-269.

949

(130) Claus, T. H.; Lowe, D. B.; Liang, Y.; Salhanick, A. I.; Lubeski, C. K.; Yang, L.; Lemoine,

950

L.; Zhu, J.; Clairmont, K. B. Specific inhibition of hormone-sensitive lipase improves lipid

951

profile while reducing plasma glucose. J. Pharmacol. Exp. Ther. 2005, 315, 1396-1402.

952

(131) Jung, U. J.; Choi, M. S. Obesity and its metabolic complications: the role of adipokines

953

and the relationship between obesity, inflammation, insulin resistance, dyslipidemia and

954

nonalcoholic fatty liver disease. Int. J. Mol. Sci. 2014, 15, 6184-6223.

955

(132) Kim, H. S.; Sung, H. Y.; Kim, M. S.; Kim, J. L.; Kang, M. K.; Gong, J. H.; Park, H. S.;

956

Kang, Y. H. Oleanolic acid suppresses resistin induction in adipocytes by modulating Tyk-STAT

957

signaling. Nutr. Res. 2013, 33, 144-153.

958

(133) Angulo, P.; Lindor, K. D. Non-alcoholic fatty liver disease. J. Gastroenterol. Hepatol.

959

2002, 17 Suppl, S186-190.

43 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 44 of 69

960

(134) Choudhury, J.; Sanyal, A. J. Insulin resistance and the pathogenesis of nonalcoholic fatty

961

liver disease. Clin. Liver Dis. 2004, 8, 575-594, ix.

962

(135) James, O. F.; Day, C. P. Non-alcoholic steatohepatitis (NASH): a disease of emerging

963

identity and importance. J. Hepatol. 1998, 29, 495-501.

964

(136) Zhou, X.; Zeng, X. Y.; Wang, H.; Li, S.; Jo, E.; Xue, C. C.; Tan, M.; Molero, J. C.; Ye, J.

965

C. Hepatic FoxO1 acetylation is involved in oleanolic acid-induced memory of glycemic control:

966

novel findings from Study 2. PloS one 2014, 9, e107231.

967

(137) Lee, W. J.; Kim, M.; Park, H. S.; Kim, H. S.; Jeon, M. J.; Oh, K. S.; Koh, E. H.; Won, J.

968

C.; Kim, M. S.; Oh, G. T.; Yoon, M.; Lee, K. U.; Park, J. Y. AMPK activation increases fatty

969

acid oxidation in skeletal muscle by activating PPARalpha and PGC-1. Biochem. Biophys. Res.

970

Commun. 2006, 340, 291-295.

971

(138) Li, Y.; Xu, S.; Mihaylova, M. M.; Zheng, B.; Hou, X.; Jiang, B.; Park, O.; Luo, Z.; Lefai,

972

E.; Shyy, J. Y.; Gao, B.; Wierzbicki, M.; Verbeuren, T. J.; Shaw, R. J.; Cohen, R. A.; Zang, M.

973

AMPK phosphorylates and inhibits SREBP activity to attenuate hepatic steatosis and

974

atherosclerosis in diet-induced insulin-resistant mice. Cell Metab. 2011, 13, 376-388.

975

(139) Wang, X.; Ye, X. L.; Liu, R.; Chen, H. L.; Bai, H.; Liang, X.; Zhang, X. D.; Wang, Z.; Li,

976

W. L.; Hai, C. X. Antioxidant activities of oleanolic acid in vitro: possible role of Nrf2 and MAP

977

kinases. Chem.-Biol. Interact. 2010, 184, 328-337.

978

(140) Andrikopoulos, N. K.; Kaliora, A. C.; Assimopoulou, A. N.; Papageorgiou, V. P.

979

Inhibitory activity of minor polyphenolic and nonpolyphenolic constituents of olive oil against in

980

vitro low-density lipoprotein oxidation. J. Med. Food 2002, 5, 1-7.

981

(141) Checker, R.; Sandur, S. K.; Sharma, D.; Patwardhan, R. S.; Jayakumar, S.; Kohli, V.;

982

Sethi, G.; Aggarwal, B. B.; Sainis, K. B. Potent anti-inflammatory activity of ursolic acid, a

44 ACS Paragon Plus Environment

Page 45 of 69

Journal of Agricultural and Food Chemistry

983

triterpenoid antioxidant, is mediated through suppression of NF-kappaB, AP-1 and NF-AT. PloS

984

one 2012, 7, e31318.

985

(142) Somova, L. O.; Nadar, A.; Rammanan, P.; Shode, F. O. Cardiovascular,

986

antihyperlipidemic and antioxidant effects of oleanolic and ursolic acids in experimental

987

hypertension. Phytomedicine 2003, 10, 115-121.

988

(143) Sena, C. M.; Pereira, A. M.; Seica, R. Endothelial dysfunction - a major mediator of

989

diabetic vascular disease. Biochim. Biophys. Acta 2013, 1832, 2216-2231.

990

(144) Rodriguez-Rodriguez, R.; Stankevicius, E.; Herrera, M. D.; Ostergaard, L.; Andersen, M.

991

R.;

992

independent release of endothelium-derived nitric oxide. Br. J. Pharmacol. 2008, 155, 535-546.

993

(145) Fu, J. Y.; Qian, L. B.; Zhu, L. G.; Liang, H. T.; Tan, Y. N.; Lu, H. T.; Lu, J. F.; Wang, H.

994

P.; Xia, Q. Betulinic acid ameliorates endothelium-dependent relaxation in L-NAME-induced

995

hypertensive rats by reducing oxidative stress. Eur. J. Pharm. Sci. 2011, 44, 385-391.

996

(146) Qian, L. B.; Fu, J. Y.; Cai, X.; Xia, M. L. Betulinic acid inhibits superoxide anion-

997

mediated impairment of endothelium-dependent relaxation in rat aortas. Indian J. Pharmacol.

998

2012, 44, 588-592.

999

(147) Jin, S. W.; Choi, C. Y.; Hwang, Y. P.; Kim, H. G.; Kim, S. J.; Chung, Y. C.; Lee, K. J.;

1000

Jeong, T. C.; Jeong, H. G. Betulinic Acid Increases eNOS Phosphorylation and NO Synthesis via

1001

the Calcium-Signaling Pathway. J. Agric. Food Chem. 2016, 64, 785-791

1002

(148) Steinkamp-Fenske, K.; Bollinger, L.; Voller, N.; Xu, H.; Yao, Y.; Bauer, R.; Forstermann,

1003

U.; Li, H. Ursolic acid from the Chinese herb danshen (Salvia miltiorrhiza L.) upregulates eNOS

1004

and downregulates Nox4 expression in human endothelial cells. Atherosclerosis 2007, 195,

1005

e104-111.

Ruiz-Gutierrez, V.; Simonsen, U. Oleanolic acid induces relaxation and calcium-

45 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 46 of 69

1006

(149) Steinkamp-Fenske, K.; Bollinger, L.; Xu, H.; Yao, Y.; Horke, S.; Forstermann, U.; Li, H.

1007

Reciprocal regulation of endothelial nitric-oxide synthase and NADPH oxidase by betulinic acid

1008

in human endothelial cells. J. Pharmacol. Exp. Ther. 2007, 322, 836-842.

1009

(150) Pendyala, S.; Natarajan, V. Redox regulation of Nox proteins. Respir. Physiol. Neurobiol.

1010

2010, 174, 265-271.

1011

(151) Yang, J. J.; Gong, Y.; Shi, J.; Qi, M. Y. [Study on the effect of ursolic acid (UA) on the

1012

myocardial fibrosis of experimental diabetic mice]. Zhongguo Yingyong Shenglixue Zazhi 2013,

1013

29, 353-356.

1014

(152) Du, Y.; Ko, K. M. Oleanolic acid protects against myocardial ischemia-reperfusion injury

1015

by enhancing mitochondrial antioxidant mechanism mediated by glutathione and alpha-

1016

tocopherol in rats. Planta Med. 2006, 72, 222-227.

1017

(153) Senthil, S.; Chandramohan, G.; Pugalendi, K. V. Isomers (oleanolic and ursolic acids)

1018

differ in their protective effect against isoproterenol-induced myocardial ischemia in rats. Int. J.

1019

Cardiol. 2007, 119, 131-133.

1020

(154) Yang, Y.; Li, C.; Xiang, X.; Dai, Z.; Chang, J.; Zhang, M.; Cai, H.; Zhang, H.; Zhang, M.;

1021

Guo, Y.; Wu, Z. Ursolic acid prevents endoplasmic reticulum stress-mediated apoptosis induced

1022

by heat stress in mouse cardiac myocytes. J. Mol. Cell. Cardiol. 2014, 67, 103-111.

1023

(155) Gugliucci, A.; Bendayan, M. Renal fate of circulating advanced glycated end products

1024

(AGE): evidence for reabsorption and catabolism of AGE-peptides by renal proximal tubular

1025

cells. Diabetologia 1996, 39, 149-160.

1026

(156) Dunlop, M. Aldose reductase and the role of the polyol pathway in diabetic nephropathy.

1027

Kidney Int. Suppl., 2000, 77, S3-12.

46 ACS Paragon Plus Environment

Page 47 of 69

Journal of Agricultural and Food Chemistry

1028

(157) Zeng, X. Y.; Wang, Y. P.; Cantley, J.; Iseli, T. J.; Molero, J. C.; Hegarty, B. D.; Kraegen,

1029

E. W.; Ye, Y.; Ye, J. M. Oleanolic acid reduces hyperglycemia beyond treatment period with

1030

Akt/FoxO1-induced suppression of hepatic gluconeogenesis in type-2 diabetic mice. PloS one

1031

2012, 7, e42115.

1032

(158) Zhou, Y.; Li, J. S.; Zhang, X.; Wu, Y. J.; Huang, K.; Zheng, L. Ursolic acid inhibits early

1033

lesions of diabetic nephropathy. Int. J. Mol. Med. 2010, 26, 565-570.

1034

(159) Wang, Z. H.; Hsu, C. C.; Huang, C. N.; Yin, M. C. Anti-glycative effects of oleanolic acid

1035

and ursolic acid in kidney of diabetic mice. Eur. J. Pharmacol. 2010, 628, 255-260.

1036

(160) Yin, M. C.; Chan, K. C. Nonenzymatic antioxidative and antiglycative effects of oleanolic

1037

acid and ursolic acid. J. Agric. Food Chem. 2007, 55, 7177-7181.

1038

(161) Yin, M. C. Inhibitory effects and actions of pentacyclic triterpenes upon glycation.

1039

BioMedicine, 2015, 5, 13.

1040

(162) Lu, J.; Wu, D. M.; Zheng, Y. L.; Hu, B.; Zhang, Z. F.; Ye, Q.; Liu, C. M.; Shan, Q.; Wang,

1041

Y. J. Ursolic acid attenuates D-galactose-induced inflammatory response in mouse prefrontal

1042

cortex through inhibiting AGEs/RAGE/NF-kappaB pathway activation. Cerebral cortex, 2010,

1043

20, 2540-2548.

1044

(163) Jeong, D. W.; Kim, Y. H.; Kim, H. H.; Ji, H. Y.; Yoo, S. D.; Choi, W. R.; Lee, S. M.; Han,

1045

C. K.; Lee, H. S. Dose-linear pharmacokinetics of oleanolic acid after intravenous and oral

1046

administration in rats. Biopharm. Drug Dispos. 2007, 28, 51-57.

1047

(164) Chen, Q. ; Luo, S.; Zhang, Y.; Chen, Z. Development of a liquid chromatography-mass

1048

spectrometry method for the determination of ursolic acid in rat plasma and tissue: application to

1049

the pharmacokinetic and tissue distribution study. Anal. Bioanal. Chem. 2011, 399, 2877-2884.

47 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 48 of 69

1050

(165) Godugu, C.; Patel, A. R.; Doddapaneni, R.; Somagoni, J.; Singh, M. Approaches to

1051

improve the oral bioavailability and effects of novel anticancer drugs berberine and betulinic

1052

acid. PloS one, 2014, 9, e89919.

1053

(166) Chen, Y.; Liu, J.; Yang, X.; Zhao, X.; Xu, H. Oleanolic acid nanosuspensions: preparation,

1054

in-vitro characterization and enhanced hepatoprotective effect. J. Pharm. Pharmacol. 2005, 57,

1055

259-264.

1056

(167) Qian, Z.; Wang, X.; Song, Z.; Zhang, H.; Zhou, S.; Zhao, J.; Wang, H. A phase I trial to

1057

evaluate the multiple-dose safety and antitumor activity of ursolic acid liposomes in subjects

1058

with advanced solid tumors. Biomed. Res. Int. 2015, 2015, 809714.

1059

(168) Wang, Y. Y.; Yang, Y. X.; Zhe, H.; He, Z. X.; Zhou, S. F. Bardoxolone methyl (CDDO-

1060

Me) as a therapeutic agent: an update on its pharmacokinetic and pharmacodynamic properties.

1061

Drug Des. Devel. Ther. 2014, 8, 2075-2088.

1062

(169) Wu, P.; Zheng, J.; Huang, T.; Li, D.; Hu, Q.; Cheng, A.; Jiang, Z.; Jiao, L.; Zhao, S.;

1063

Zhang, K. Synthesis and Evaluation of Novel Triterpene Analogues of Ursolic Acid as Potential

1064

Antidiabetic Agent. PloS one, 2015, 10, e0138767.

1065

(170) Pergola, P. E.; Krauth, M.; Huff, J. W.; Ferguson, D. A.; Ruiz, S.; Meyer, C. J.; Warnock,

1066

D. G. Effect of bardoxolone methyl on kidney function in patients with T2D and Stage 3b-4

1067

CKD. Am. J. Nephrol. 2011, 33, 469-476.

1068

(171) Pergola, P. E.; Raskin, P.; Toto, R. D.; Meyer, C. J.; Huff, J. W.; Grossman, E. B.; Krauth,

1069

M.; Ruiz, S.; Audhya, P.; Christ-Schmidt, H.; Wittes, J.; Warnock, D. G.; Investigators, B. S. N.

1070

Bardoxolone methyl and kidney function in CKD with type 2 diabetes. Engl. J. Med. 2011, 365,

1071

327-336.

48 ACS Paragon Plus Environment

Page 49 of 69

Journal of Agricultural and Food Chemistry

1072

(172) de Zeeuw, D.; Akizawa, T.; Audhya, P.; Bakris, G. L.; Chin, M.; Christ-Schmidt, H.;

1073

Goldsberry, A.; Houser, M.; Krauth, M.; Lambers Heerspink, H. J.; McMurray, J. J.; Meyer, C.

1074

J.; Parving, H. H.; Remuzzi, G.; Toto, R. D.; Vaziri, N. D.; Wanner, B.; Wittes, J.; Wrolstad, D.;

1075

Chertow, G. M.; Investigators, B. T. Bardoxolone methyl in type 2 diabetes and stage 4 chronic

1076

kidney disease. N. Engl. J. Med. 2013, 369, 2492-2503.

1077

(173) Wang, X.; Chen, Y.; Abdelkader, D., Hassan, W.; Sun, H.; Liu, J. Combination therapy

1078

with oleanolic acid and metformin as a synergistic treatment for diabetes. J. Diabetes Res. 2015,

1079

2015, 973287.

1080

(174) Liu, J.; Lu, Y. F.; Zhang, Y.; Wu, K. C.; Fan, F.; Klaassen, C. D. Oleanolic acid alters bile

1081

acid metabolism and produces cholestatic liver injury in mice. Toxicol. Appl. Pharmacol. 2013,

1082

272, 816-824.

1083

(175) Lu, Y. F.; Wan, X. L.; Xu, Y; Liu, J. Repeated oral administration of oleanolic acid

1084

produces cholestatic liver injury in mice. Molecules 2013, 18, 3060-3071.

1085

49 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1086

Page 50 of 69

FIGURES AND LEGEND

1087 1088

Figure 1. Chemical structures of some bioactive triterpenic acids: oleanolic, ursolic and betulinic

1089

acids. Adapted from4.

1090

50 ACS Paragon Plus Environment

Page 51 of 69

Journal of Agricultural and Food Chemistry

1091 1092

Figure 2. Mechanisms of oleanolic (OA), ursolic (UA) and betulinic (BA) acids on diabetes type 2 and

1093

related comorbidities. The different hypoglycemic (orange continuous arrow) and anti-obesity (red

1094

dash arrows) properties of these TAs are schematized. These TAs promote the reduction of glucose

1095

absorption (1) by inhibition of pancreatic α-amylase and α-glucosidase or by down-regulation of

1096

sodium-dependent glucose co-transporter (SGLT-1) and glucose transporter 2 (GLUT2); as well as

1097

reducing the endogenous glucose production (EGP) and increase in glycogen synthesis (2), as

1098

consequence of the reduction of gluconeogenesis and glycogenolysis, and stimulation of glycogenesis.

1099

The improvement of insulin signaling (3) by OA, UA and BA is highlighted and mediated i) by

1100

inhibition of protein tyrosine phosphatase (PTP), ii) by improvement of β-cell survival and direct

1101

stimulation of phosphatidylinositol 3-kinase-dependent (PI3KB)/protein kinase Akt pathway, iii) by

1102

translocation of GLUT-4, iv) by activation TGR5 and glucagon-like peptide-1 (GLP-1) release, and

1103

also v) by inhibition of ROS production and suppression of inflammatory processes. Anti-obesity 51 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 52 of 69

1104

effects occur due to the reduction of lipid absorption (4) by inhibition of pancreatic α-lipase or human

1105

cholesterol acyltransferase -2 (hACAT-2); regulation of lipid metabolism (5) by reduction of

1106

lipogenesis and stimulation of fatty acid β-oxidation; and also by promoting reduction adipogenesis,

1107

stimulating the lipolysis in adipose tissue. The regulation of appetite (6) is another effect of these

1108

compounds, which occur as consequence of the reduction of ghrelin and resistin and the increase of

1109

adiponectin and leptin levels, although the effect of UA on leptin remain controversial. The protective

1110

properties of OA, UA and BA against diabetes-related comorbidities (blue dash arrows) may act on

1111

preventing the development of non-alcoholic fatty liver disease (NAFLD) (7), cardiovascular disease

1112

(CVD) (8) or diabetic nephropathy (9), as consequence of their hypolipidemic, anti-atherogenic,

1113

hypotensive, anti-inflammatory and anti-oxidant properties, and also by preserving kidney structure

1114

and promoting the inhibition of polyol pathway. ↕= regulates, ↑= enhances/stimulates/promotes, (+):

1115

up-regulates/activates, ↗= improves, ↓= reduces, (-) = downregulates,

= inhibits.

1116

52 ACS Paragon Plus Environment

Page 53 of 69

Journal of Agricultural and Food Chemistry

1117

TABLES

Biological action

TAs dosage and biological system used OA and UA: 5 µg/mL (~ 10 µM) Extracted from six Malaysian plants BA: 50 µg/mL (~100 µM) Extracted from leaves of Dillenia indica OA: 5-20 mg/kg/day during 15 weeks In vivo: HFD fed mice UA: 10 mg/kg/day during 15 weeks In vivo: HFD fed mice

Biological outcome to TAs

Reference

↓↓pancreatic α-amylase activity

46

↓↓pancreatic ߙ-amylase and ߙ-glucosidase activities

47

32 ↓↓pancreatic α-amylase activity, ↓blood glucose levels

BA: 10 mg/kg/day during 15 weeks In vivo: HFD fed mice

Absorption and uptake of glucose

48

49

OA: IC50 35 µM, UA: IC50 39 µM Extracted from Punica granatum L. flowers

54

OA: IC50 231.3 µM Extracted from Phlomis stewartii

50

UA: IC50 0.119 µM Extracted from Stereospermum colais

51 ↓↓α-glucosidase activity

UA: IC50 16 µM Extracted from stem barks of Uncaria laevigata

52

OA: IC50 49,5 µM, UA: IC50 47,6 µM, BA: IC50 14,0 µM Extracted from gold-red apple

53

OA: IC50 0.1 mg/mL (~200 µM) In vivo: male GK/Jcl type 2 diabetic rats and healthy human adult volunteers Extracted from olive leaf

↓↓pancreatic and salivary α-amylase activity, ↓blood glucose levels

55

OA: 80 mg/kg twice daily during 5 weeks (in vivo) 4.37–21.9 µM (in vitro) In vivo: STZ-DM rats

In vivo:↓blood glucose levels, ↓SGLT-1 and GLUT-2 expressions In vitro:↓↓α-amylase (IC50 of 3.60 µM) and α-glucosidase ( IC50 of 12.40 µM) activities

16

53 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Table 1. Summary of studies reporting the hypoglycemic effects of OA, UA and BA. Endogenous glucose production and synthesis of glycogen

OA: IC50 14.9 µM UA: IC50 9.0 µM BA: IC50 43.0 µM UA: 0.01-0.05% (w/w) during 4 weeks

Page 54 of 69

↓↓muscle GP activity

57

↓blood glucose level, ↓↓hepatic G6Pase and ↑↑GK activities, ↑GK/G6Pase ratio, ↑GLUT-2

17

1118

54 ACS Paragon Plus Environment

Page 55 of 69

Journal of Agricultural and Food Chemistry

In vivo: STZ/NA-DM mice UA: 5 mg/kg/day during 5 weeks In vivo: HFD C57BL/6J mice UA: 0.05% (w/w) during 4 weeks In vivo: HFD and STZ-DM mice

mRNA levels ↓glucose, HbA1c and insulin levels, ↑glycogen content, ↑↑GK and ↓↓G6Pase activities ↓blood glucose level, ↑hepatic glycogen content, ↑↑GK and ↓↓G6Pase activities

UA: 2 mg/kg/day during 7 days In vivo: healthy rats

↓ plasma glucose concentration, ↑liver glycogen levels, ↑phospho-GSK-3

62

↑glucose uptake, ↑glycogen synthesis ↑phospho-AMPK, ↑ phospho-GSK-3β

64

↑blood glucose levels, ↑insulin signaling, ↓ G6P and PEPCK protein expression, ↑PGC-1α mRNA expression, ↑phospho-AMPK, ↑mitochondrial density, ↓structural injury of mitochondria, ↑∆Ѱm, ↑ATP content

18

BA: 10-20 µM and 5-10 mg/kg during 3 weeks In vitro: Human HepG2 hepatoma cells In vivo: HFD ICR mice

In vitro: ↓hepatic glucose, ↓PGC-1α, PEPCK, and G6Pase gene expression,↑AMPK, ↓phospho-CREB, ↑phospho-CAMKK, In vivo: ↓plasma glucose, ↓insulin resistance index

19

BA: 10 µM In vitro: differentiated 3T3-L1 adipocytes, C2C12 myotubes, HUVEC, RAW264.7 macrophages, MEF and CHO cells

↑ glucose uptake, ~LDH, ATP levels, cell viability, ↑ higher extracellular acidification rate and ↑extracellular lactate, ↑mitochondrial content and mitochondrial ROS production, ↓oxygen consumption rate, ↑UCP-1 and -2 protein expression, ↑GLUT1 and GLUT2 protein expression, phosphor-AMPK ↑glucose levels, ↑ insulin and c-peptide, → β-cells,↑proliferation T lymphocytes, ↑IL-2 and IFN-γ, ↓TNF-α

68

OA, UA and BA: 10 µM In vitro: Human HepG2 hepatoma cells OA: 20 mg/kg/day during 2 weeks In vivo: Lep(db)(/)(db) obese DM mice

UA: 0.05% (w/w) during 4 weeks In vivo: HFD STZ-DM mice

20

59

69

Insulin resistance UA: 0.05% (w/w) during 8 weeks In vivo: HFD mice

↓glucose levels, ↑insulin level, →islet architecture

70

1119

55 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 56 of 69

OA: 30 and 50 µM In vitro: pancreatic β-cells (INS-1 832/13 cells) and rat islets

↑insulin secretion, ↑acute glucose-stimulated insulin secretion, ↑insulin protein and mRNA levels, ~ somastatin, glucagon, cellular Ca2+ and cAMP levels

71

OA: 5–50 µg/mL (~10 – 100 µM) In vitro: pancreatic β-cells Extracted from Cabernet Sauvignon grape skin

↑Insulin production

73

↗ glycemia, ↑insulin secretion, ↑glycogen content, ↑glucose uptake in muscle ↑GLUT-4 protein expression, ↑localization GLUT-4 at the plasma membrane ↗ glycemia, ↑insulin vesicle translocation, ↑insulin secretion, ↑glycogen content in muscle, ↑glucose uptake in muscle tissue ↑GLUT-4 protein and mRNA expression

74

BA: 0.1, 1 and 10 mg/kg In vivo: Male wistar rats

UA: 0.1, 1 and 10 mg/kg In vivo: Male wistar rats

OA: IC50 3.37-3.40 µM

↓↓PTP-1B (IC50 3.37 µM), ↓↓TCPTP (IC50 3.40 µM)

OA: IC50 9.50 µM, UA: IC50 2.30 µM

75

76

79 78

UA: IC50 3.8 µM Extracted from S. paniculata BA: IC50 0.7 µg/mL Extracted from roots of Saussurea lappa C.B.Clarke

↓↓PTP-1B

BA: IC50 3.5 µM Extracted from the fruits Sorbus pohuashanensis UA: IC50 2.73-3.33 µM

77

71

↓↓PTP-1B (IC50 3.08 µM), ↓↓TCPTP (IC50 3.33 µM), ↓↓SHP-2 (IC50 2,73 µM)

77

1120

56 ACS Paragon Plus Environment

Page 57 of 69

Journal of Agricultural and Food Chemistry

UA: 0.01–100 µg/mL (~ 0.02–200 µM) In vitro: CHO/IR cells and 3T3-L1 primary adipocytes

OA: 5 and 25 mg/kg/day during 10 weeks In vivo: Adipo-IR rats

UA: 2.5–10 µM In vitro: 3T3-L1 pre-adipocytes

OA: EC50 1.42 µM and 100 mg/kg/day during 7 and 14 days In vitro: CHO-TGR5 cell line In vivo: HFD C57BL/6J mouse Extracted from Olea europaea

OA: EC50 2.25 µM, UA: EC50 1.43 µM, BA: EC50 1.04 µM In vitro: CHO cells OA: 10-30 µM In vitro: Murine enteroendocrine cells (STC-1) OA: 10–50 µM In vitro: pancreatic β cell line MIN6 and also in mouse and human pancreatic islets

↑phospho-IR β-subunit in CHO/IR and adipocytes, ↑IR in CHO/IRs, ↑phospho-Akt and phospho -ERK in CHO/IRs, ↑phospho-Akt and phospho GSK-3 in adipocytes, ↑translocation of GLUT-4 in adipocytes insulin-mimetic agent (>50 µg/mL) insulin-sensitizer (