FITOTE-03169; No of Pages 11 Fitoterapia xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Fitoterapia journal homepage: www.elsevier.com/locate/fitote

Review

4Q2

Jichao Chen, Wenlong Li, Hequan Yao, Jinyi Xu ⁎

5 6

State Key Laboratory of Natural Medicines, China Pharmaceutical University, 24 Tong Jia Xiang, Nanjing 210009, PR China Department of Medicinal Chemistry, China Pharmaceutical University, 24 Tong Jia Xiang, Nanjing 210009, PR China

R O O

3

Insights into drug discovery from natural products through structural modification

2Q1

23 24 22 25 26

Keywords: Natural products Druggability Structural modification

28 27

Contents

Natural products (NPs) have played a key role in drug discovery and are still a prolific source of novel lead compounds or pharmacophores for medicinal chemistry. Pharmacological activity and druggability are two indispensable components advancing NPs from leads to drugs. Although naturally active substances are usually good lead compounds, most of them can hardly satisfy the demands for druggability. Hence, these structural phenotypes have to be modified and optimized to overcome existing deficiencies and shortcomings. This review illustrates druggability optimization of NPs through structural modification with some successful examples. © 2015 Published by Elsevier B.V.

D

Article history: Received 5 March 2015 Accepted in revised form 19 April 2015 Accepted 20 April 2015 Available online xxxx

a b s t r a c t

E

10 11 12 13 14

C E R

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1. Potential of natural products . . . . . . . . . . . . . . . . . 1.2. Challenges with natural product research . . . . . . . . . . . 2. General principles for structural modification of natural products . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Examples for structural modification of natural products . . . . 2.1.1. Improvement of physico-chemical properties . . . . . 2.1.2. Improvement of plasma stability . . . . . . . . . . . 2.1.3. Improvement of metabolic stability . . . . . . . . . 2.1.4. Improvement of crossing the blood–brain barrier (BBB) 2.1.5. Improvement of potency and selectivity . . . . . . . 2.1.6. Employment of prodrug strategy . . . . . . . . . . . 2.1.7. The intellectual property attainment status of NPs . . . 3. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

N C

O

R

1.

U

29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44

i n f o

T

a r t i c l e

P

7 9

F

1

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0 0 0

. . . . . . . . . . . .

0 0 0 0 0 0 0 0 0 0 0 0

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

45 ⁎ Corresponding author at: State Key Laboratory of Natural Medicines, The Department of Medicinal Chemistry, China Pharmaceutical University, Nanjing 210009, China. Tel.: +86 25 83271299; fax: +86 25 83302827. E-mail address: [email protected] (J. Xu).

http://dx.doi.org/10.1016/j.fitote.2015.04.012 0367-326X/© 2015 Published by Elsevier B.V.

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

15 16 17 18 19 20 21

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

64 65

87

Structural diversity is a striking feature of NPs accounting for their lasting importance in drug discovery [5,6]. With increasing druggable targets in the postgenomic era, chemical diversity of screening libraries plays a major role in the fierce competition among pharmaceutical companies. In this respect, NPs are a rather indispensable complement to synthetic compound collections [7,8]. NPs are also featured by steric complexity and differ from synthetic compounds with respect to the statistical distribution of functionalities [8]. They interrogate a different and wider chemical space, and possess a broader dispersion of structural and physicochemical properties than synthetic compounds [9–11]. Some academic groups and companies have made their attempts to synthesize increasingly complex structures to match the chemical space occupied by NPs, however, about 83% of core ring scaffolds present in NPs are still absent from commercially available molecules and screening libraries [12]. In addition, most NPs show more favorable ADME/T properties compared to synthetic molecules, although they often deviate from “druglikeness” criteria, such as Lipinski's Rule of Five [13,14]. Moreover, NPs recognized as ‘privileged structures’ also provide attractive scaffolds for combinatorial synthesis and library design [15–17], and serve as chemical probes for the validation of new drug targets [18].

88

1.2. Challenges with natural product research

89 90

Despite the proven track record of natural products in drug discovery and their uncontested unique structural diversity, there are still several problems associated with NPs. Typical limitations of NP leads are low solubility or chemical instability, which especially hamper the development of parenteral drugs [19]. Furthermore, many natural products are complex structures with high molecular weight. “Heavy” structures break Lipinski's rules and will most likely exhibit no absorption from the gut into the blood, therefore impeding oral formulation [20]. Additionally, the intellectual property situation is often less clear in unmodified NPs. Although naturally active substances are usually good lead compounds, most of them can hardly satisfy the demands for druggability. NP leads are

72 73 74 75 76 77 78 79 80 81 82 83 84 85 86

91 92 93 94 95 96 97 98 99 100 101

C

70 71

E

68 69

R

66 67

R

60 61

O

58 59

C

56 57

N

54 55

U

52 53

2. General principles for structural modification of natural products

106 107

The ultimate goal of structural modification of NPs is to obtain new drugs. Pharmacological activity and druggability are two essential factors for drug innovation. Pharmacological activity is definitely indispensable, and druggability is destined by physico-chemical, biochemical, pharmacokinetic and safety properties of drugs. Structural modification of NPs should, therefore, be involved in all the contents of the above two aspects. It is necessary to perform selective modification of NPs according to the deficiency or shortcomings of the structure, the activity, and physico-chemical and pharmacokinetic properties. Generally, the following principles should be followed: increasing potency and selectivity, improving physicochemical properties such as solubility, distribution, ionizability, etc., enhancing the chemical and metabolic stability, improving biochemical properties, improving pharmacokinetic properties including absorption, distribution, metabolism and excretion, eliminating or reducing side-effects, and attaining intellectual properties. Some successful examples of NPs through structural modification are illustrated as follows.

108

F

1.1. Potential of natural products

50 51

O

63

49

R O

62

Over the past century, a number of natural compounds extracted from animals, plants, microbes and marine organisms have been used to treat human diseases [1,2]. The endeavor involved well-known drugs such as penicillin (antibacterial), morphine (analgesic), artemisinin (antimalarial) and paclitaxel (anticancer). Review of natural products (NPs) over the 30 years from 1981 to 2010 revealed that approximately 40% of the developed therapeutic agents approved by FDA were NPs, their derivatives, or synthetic mimetics related to NPs [3]. Despite increasing competition from combinatorial and classical compound libraries, there has been a steady introduction of NP-derived drugs in the last years. A total of 19 NPs-based drugs were approved for marketing worldwide between 2005 and 2010, covering infectious (bacterial, fungal, parasitic and viral), immunological, cardiovascular, neurological, inflammatory and related diseases, and oncology [4].

102 103

P

47 48

frequently optimized through structural modification to achieve the final candidate. Of the 58 NP drugs launched in the period of 1981–2011, 31 are structural analogues of the native NP leads [21].

D

1. Introduction

T

46

E

2

104 105

109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126

2.1. Examples for structural modification of natural products

127

2.1.1. Improvement of physico-chemical properties

128

2.1.1.1. Increase of low solubility. Low water solubility limits absorption and causes poor oral bioavailability. Although many cutting-edge technologies have been developed to formulate insoluble compounds over the years, medicinal chemists would like to solve drug delivery problems with “covalent bonds” [22], to improve solubility through structural modifications including adding ionizable or polar groups. Typically, a basic amine or a carboxylic acid is introduced to the structure. Artemisinin (1), a sesquiterpene lactone peroxide originally isolated from Chinese traditional medicine qinghao (Artemisia annua L.) [23], has shown potent activity in a variety of diseases, most notably malaria and, more recently, several types of cancers [24] and cytomegalovirus [25]. However, its poor solubility in water or oil caused difficulty in its rescue of severe malaria patients. To address this, oil-soluble artemether (2) and water-soluble sodium artesunate (3) were approved in China in 1987. The sodium salt of a carboxylic acid analogue achieved higher solubility, yet in this case the compound was unstable, resulting from the facile hydrolysis of the ester linkage. Although the carboxylic group in compound 4 was linked to the artemisinin nucleus via a more stable ethereal linkage, their antimalarial activity was much less active than that of sodium artesunate [26]. Ultimately, the amine analogues (maleates or oxlates of 5) attained better solubility and stability and were active after oral dosing [27]. Despite immense efforts, no new artemisinin derivative has been developed. Artemisone (6) was a 10-alkylaminoartemisinin analogue whose preparation entailed chemistry distinct to that leading to other derivatives and which extended the efficacy limit beyond

129 130

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

158 159 160

those of the known artemisinins [28]. The drug is currently in phase III clinical trials.

3

solubility (N 10 mg/mL). The compound has a water-soluble dipeptide ester group at the C-20 position of 17, which is stable at low pH but rapidly converts to the active drug 17 at physiological pH by intramolecular cyclization of the dipeptide moiety [38].

202 203 204 205 206 207

171 172 173 174 175 176 177 178 179

187 188 189 190 191 192 193 194 195 196 197 198 199 200 201

P

211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226

O N C

185 186

Camptothecin (11), a pentacyclic alkaloid isolated from Camptotheca acuminata [36], acts as a potent topoisomerase I inhibitor (Top I) showing strong antitumor activity both in vitro and in vivo [37]. Yet, the clinical application of 11 as an anticancer agent was limited due to its nonmechanismrelated toxicity and an extremely poor solubility. Two strategies had been taken to circumvent this solubility issue: (i) synthesizing camptothecin analogues with an amino functional group, e.g., topotecan (12) and Dx-8951(13), and (ii) synthesizing water-soluble prodrugs of lipophilic camptothecin analogues, e.g., irinotecan (14) [38]. Both 12 and 14 have been approved for the treatment of several types of cancer patients [39]. Moreover, water-soluble quaternary salt camptothecin analogues (15) were also prepared with comparable or superior Top I inhibitory activities but lower cytotoxicities in relation to 11 [40]. Additionally, another analogue TP300 (16) exhibited a broader antitumor spectrum and more potent antitumor activity than 12 with improved

U

183 184

209 210

227

R

R

180 181 182

2.1.1.2. Simplification of complex structure. NPs with complex structure often remain a challenge for their chemical synthesis, especially some of those with limited availability of source materials from which they are extracted hampered their clinical application. Furthermore, complex NPs often possess higher molecular weights and bring about poor oral absorbability. Hence, it is necessary to develop structurally simpler compounds derived from NPs while retaining their bioactivity. In this respect, Morphine (18) is particularly telling. 18 inspired the development of potent analgesics through the preparation of structurally simplified analogues such as morphinanes (19), benzomorphanes (20), phenylpiperidines (21), and the simplest one, methadone (22). Despite their decreasing complexity, analgesic activity is retained to different degrees and with different affinities for opioid receptor subtypes [41]. As such, strategies for structural simplification include decreasing the molecular size and eliminating unnecessary chiral centers.

D

169 170

E

167 168

T

165 166

Oridonin (7) is a complex ent-kaurane diterpenoid extracted from the traditional Chinese herb Isodon rubescens [29]. 7 has demonstrated great potential in the treatment of various human cancers due to its unique and safe anticancer pharmacological profile [30–32], and has recently been reported to have a potent antimycobacterial activity [33]. Nevertheless, the sparing solubility (1.29 mg/mL) limited its extensive use in clinic. By introducing various water-soluble amino acids and carboxylic acids into the C-14 position of 7, 1-O- and 14-O- derivatives of oridonin were obtained with improved solubility (N50 mg/mL) and stronger cytotoxicity against six cancer cell lines (BGC-7901, SW-480, HL-60, BEL-7402, A549 and B16) than that of 7 (IC50 = 26–40 μM), especially compounds 8 and 9 with IC50 values of 0.84 μM in HL-60 cell and 1.00 μM in BEL-7402 cell, respectively [34]. Recently, synthesis of the nitrogen-enriched oridonin derivatives with thiazole-fused A-ring (10) led to potent antiproliferative effects against breast, pancreatic and prostate cancer cells with low micromolar to submicromolar IC50 values as well as markedly enhanced aqueous solubility (N40 mg/mL) [35].

C

163 164

E

161 162

R O O

F

208

Largazole (23), a depsipeptide natural product isolated from the cyanobacterium Symploca sp. [42], is a histone deacetylase (HDAC) inhibitor with potent antiproliferative activity and selectivity for cancer cells. One study showed that C7-demethyl largazole was equipotent to 23 [43], indicating that the C7-methyl is unimportant for activity. Another study displayed that related thiazole–thiazole analogue 24 exhibited an intermediate HDAC inhibitory activity compared to largazole thiol [44], suggesting that thiazoline–thiazole moiety could be replaced with bisthiazole fragment. Based on these findings, a series of simplified analogues were developed by retaining the bisthiazole cap group while cleaving across the β-hydroxy and thiazole fragments. Among them, 25 showed a comparable HDAC inhibitory activity (IC50 = 70 nM) to that of 23 (IC50 = 13.7 nM) and was efficacious when administered orally [45].

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243

4

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

disease-modifying agent approved in 2010 for the treatment of 294 relapsing–remitting multiple sclerosis [58]. 295 296

244 245

259 260 261 262 263 264 265 266 267 268 269 270 271 272

F O

R O

257 258

P

255 256

D

253 254

2.1.1.3. Improvement of chemical stability. Chemical instability in bioassay buffer solution reduces the compound concentration and produces decomposition products that may be active. Oxygen, water, light, trace metals, and materials leached from the glass or plastic containers can react with the compound. The structural modifications that can improve chemical stability depend on the conditions and functional groups, including elimination, replacement or modification of unstable groups. Liphagal (33), a meroterpenoid natural product collected from the sponge Aka coralliphaga, was discovered in a screening program designed to find new isoform-selective PI3K inhibitors. 33 inhibited PI3Kα with an IC50 of 100 nM and showed an approximately 10-fold selectivity for PI3Kα compared with PI3Kγ in a fluorescent polarization enzyme bioassay [59]. It had been observed that the 6, 7 A/B ring of 33 can be rearranged to give spiro 6, 6 ring analogue 34 under acidic conditions, which led to the decrease of activity and selectivity. When ring B contracted to a six-membered ring resulted in the more stable analogue 35 with an IC50 of 66 nM and 27-fold selectivity for P13 Kα/ Kγ. At the same time, in the structure of 35, the C-8 methyl was eliminated to avoid the problem of lowered yields associated with generating mixtures of epimers at that position [60].

E

251 252

298

T

250

C

248 249

Halichondrin B (26), a large polyether macrolide discovered from several sponge sources, exhibits growth inhibition on a panel of various cancer cell lines at nanomolar concentrations [46]. It suppressed the microtubule growth phase without affecting the shortening phase, and caused tubulin sequestration into non-productive aggregates [47]. This unprecedented mechanism of action also resulted in excellent activities on cancer cells resistant to other antimicrotubule agents like taxanes [48]. However, compound supply was the major hurdle for clinical development due to its complex structure since the beginning. In order to produce simpler analogues and to identify the minimum pharmacophore of 26, the testing of intermediates from the total synthesis revealed that growth inhibitory activity on DLD-1 cells could be traced back to the right half macrolactone fragment 27 [49]. Unfortunately, 27 was devoid of activity in human tumor xenogra models in contrast to 26. Subsequent studies displayed that simple furan or pyran fragments could replace the western fused pyran moiety [50]. Further functional group exploration and the replacement of the ester linkage in the macrolactone ring by a ketone led to the identification of eribulin (28), an analogue that showed the desired potency on a panel of cancer cell lines without reversing mitotic block [51]. On account of its remarkable biological profile, the drug was approved for metastatic breast cancer in 2010 in the US and 2011 in Europe [52].

E

246 247

300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324

279 280 281 282 283 284 285 286 287 288 289 290 291 292 293

C

277 278

ISP-1 (29) was identified as the immunosuppressive agent from culture broths of the fungus Isaria sinclairii. It has been reported that 29 was 5- to 10-fold more efficacious than cyclosporin A in vitro and in vivo [53,54]. Structure-activity relationships (SARs) of 29 and the closely related mycestericins revealed that structural simplifications, such as reducing the double bond and removing the keto and 4-hydroxy functional groups, had no effect on activity. The removal of stereogenic centers resulted in 30 with a hydroxymethyl group instead of the carboxylic acid function of 29 [55]. Interestingly, this analogue was demonstrated to be more active and less toxic in vivo. Subsequent improvement of the activity and safety profile was acquired by lessening the linear alkyl chain length from C18 to C14, leading to the simplified analogue 31 [56]. The number of rotatable bonds was decreased through the introduction of a 1, 4-phenyl moiety into the linear alkyl chain. A systematic shift of the phenyl ring along the alkyl chain finally resulted in optimum compound 32 with further efficacy improvement, which turned out to be 100 times more potent than cyclosporin A [57]. As a result, 32 was the first oral

N

275 276

U

274

O

R

R

273

299

326 327

Curcumin (36), a yellow principal polyphenol curcuminoid, was extracted from the popular Indian spice turmeric (Curcuma longa) [61]. 36 shows a variety of biological and cellular activities including antioxidant, anti-inflammatory, anticancer and hypocholesterolemic properties [62]. However, poor bioavailability and pharmacokinetic profiles due to its instability under alkaline and light conditions have limited its clinical application. 36 can exist in solution under physiological pH conditions as a tautomeric mixture of keto and enol forms, and the enol form (37) was found to be responsible for the instability of 36 [63–65]. It was reported previously that the presence of the β-diketone moiety may be essential for the biological activity of 36 [66,67]. However, recent studies reported that curcumin derivatives without the β-diketone retained their anti-proliferative activities. By deleting the β-diketone moiety, mono-carbonyl analogues (38) of curcumin exhibited enhanced stability in vitro and greatly improved pharmacokinetic profiles in vivo [68]. In

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

345 346 347 348 349

addition, the β-diketone moiety via conjugation with thioureas formed divinylpyrimidinethiones (39) [69] or was replaced by isoxazole (40) or pyrazole (41) [70] with increased chemical stability.

5

in many trials from Phase I to Phase III against a variety of 391 carcinomata such as solid tumor [83] and prostate cancer [84]. 392

393

350

F

395

359 360 361 362 363 364 365 366 367

374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390

O

372 373

2.1.2. Improvement of plasma stability Compounds with certain functional groups can decompose in the bloodstream. Unstable compounds often have high clearance and short half-life, leading to poor in vivo pharmacokinetics and disappointing pharmacological performance. The plasma stability could be increased by substituting an amide for an ester or eliminating the hydrolyzable group. Epothilone B (50) is a 16-membered macrolide antibiotic produced by the myxobacterium Sorangium cellulosum [79]. Preclinical experiments have shown that 50 has potent antineoplastic activities against a wide range of tumor cell lines in vitro. Like the taxanes, 50 promoted tumor cell death by stabilizing microtubules and inducing apoptosis [80]. However, this promising in vitro activity did not translate into robust in vivo preclinical antitumor efficacy due to its poor plasma stability and unfavorable pharmacokinetic properties. Ixabepilone (51) is an analogue designed for good plasma stability, high in vivo efficacy and increased water solubility. The lactone is replaced with a lactam which is not susceptible to hydrolysis by esterases, conferring plasma stability on 51 with increased half-life from 19 to 52 h [81]. The drug was approved for monotherapy against metastatic breast cancer by the FDA in 2007 [82], and is currently

N C

370 371

R O O

Camptothecin (CPT, 11) was terminated in clinical trials because of its poor aqueous solubility, low metabolic stability of lactone and high in vivo hepatoxicity. Although subsequent structural modification of natural CPT has generated three antitumor agents (i.e., topotecan, irinotecan and belotecan) and a number of drug candidates with improved aqueous solubility and decreased hepatoxicity, their highly electrophilic α-hydroxylactone of the E ring can be rapidly hydrolyzed to the biologically inactive carboxylate form under the physiological conditions [88]. Furthermore, the hydrolyzed CPT carboxylate binds tightly to serum albumin, which limits the fraction of drug in the active lactone form [89]. Analogues with improved intrinsic stability were explored later. This effort has led to the development of several promising analogues including seven-membered β-hydroxy CPT (54) [90], five-membered α-hydroxy keto CPT (55) [91] and α-fluoro ether CPT (56) [92]. In the structure of 56, α-fluoro ether was designed as a metabolically stable bioisostere of lactone.

U

369

R

R

E

C

368

P

357 358

396 397 398 399 400 401 402 403 404 405 406 407 408 409

D

355 356

E

353 354

Combretastatin A-4 (CA-4, 42), a naturally microtubulardestabilizing agent isolated from the South African tree Combretum caffrum, exhibits potent antitumour and antivascular properties both in vitro and in vivo [71]. The cis-double bond in 42 is liable and prone to isomerize to the more thermally stable trans-isomer, resulting in complete loss of the cytotoxicity. To circumvent this problem, a series of cis-restricted CA-4 analogues were developed. Replacement of the ethylene bridge with various heterocycles such as imidazole 43 [72], furan 44 [73], triazole 45 [74] and β-lactam 46 [75] provided a rigid scaffold, thus preventing cis-trans isomerisation. Further SARs study has shown that the cis-stilbene configuration of CA-4 is not essential for its biological activity. The two phenyl rings can be separated by non-cyclic linkers such as methylene 47 [76], carbonyl 48 [77] and sulfide 49 [78], which retain potent antitumour activity.

T

351 352

Bottromycin A2 (52), an antibacterial peptide isolated from the fermentation broth of Streptomyces bottropensis [85], possesses potent antibacterial activities, including drugresistant strains such as methicillin resistant Staphylococcus aureus (MRSA) and vancomycin-resistant Enterococci (VRE) strains [86]. However, it does not show good in vivo efficacy resulting from hydrolysis of the terminal methyl ester moiety. Replacement of the ester moiety with the amide, urea or ketone led to analogues with significantly improved stability in mouse plasma, and the ketone analogue 53 exhibited potent activity against S. aureus, MRSA and VRE, comparable to that of vancomycin [87].

411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432

2.1.3. Improvement of metabolic stability Natural drugs often encounter formidable challenges to their stability in vivo, which imposes significant limitations on the structures of NPs. Structures that are highly active in vitro

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

433 434 435 436

454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477

F

O 2.1.4. Improvement of crossing the blood–brain barrier (BBB) BBB plays a vital role in every disease process involving the CNS and is a fatal issue in the development of all CNS drugs. In order to penetrate into the brain tissue, they must pass through the BBB. Many of the compounds that might otherwise be effective in treating CNS diseases are excluded from reaching a sufficient concentration in the brain tissue and producing the desired therapeutic effect. The brain penetration could be enhanced by lowering P-glycoprotein efflux, increasing molecular lipophilicity or reducing molecular weight. Paclitaxel (66), a structurally complex microtubulestabilizing agent discovered from the bark of the Pacific Yew, has become one of the most active cancer chemotherapeutic drugs known [108]. Although its clinical success is remarkable, 66 is not an effective agent for primary or metastatic brain cancer due to its inability to cross the BBB [109]. A primary mechanism limiting the distribution of 66 into the brain is active efflux by P-glycoprotein (Pgp) [110]. Accordingly, permeation of 66 could be substantially improved by inhibiting Pgp-mediated efflux. The introduction of the succinate at the C10 position (67) reduced apparent Pgp interactions, which imparted a 10-fold increase in brain penetration (8.47 × 10−7 cm/s) compared with 66 (0.845 × 10−7 cm/s) [111].

U

N

C

478

479 480 481 482 483 484 485 486 487 488 489

Phlorizin (61), a dihydrochalcone glycoside first isolated in 1835 from the root bark of the apple tree [101], plays a key role in the history of diabetes mellitus. Its principal pharmacological action is to produce renal glycosuria and block intestinal glucose absorption through inhibition of the sodium-glucose co-transporter 2 located in the proximal renal tubule and mucosa of the small intestine [102]. Further development of 61 is limited by its low metabolic stability and toxic effects. The efforts to overcome the shortcomings led to first-generation analogues such as T-1095 (62) [103] and sergliflozin etabonate (63) [104]. The removal, substitution or modification of

492 493 494 495 496 497 498 499 500 501 502

R O

452 453

490 491

P

450 451

D

448 449

phenolic hydroxyl groups prevented phlorizin analogues from phase II metabolism. In addition, the glucose moiety was designed as the carbonate prodrug to further increase metabolic stability. However, no first-generation compound has been launched, due to the intrinsic metabolic lability of the O-glycoside bond, especially in humans, resulting in insufficient plasma half-life [105]. Substitution of the O-glycoside with the C-glycoside avoided hydrolysis by β-glucosidases, resulting in second-generation analogues including dapagliflozin (64) [106] and canagliflozin (65) [107] which were approved in Europe in 2012 and the US in 2013, respectively.

T

446 447

C

444 445

E

442 443

R

441

R

439 440

may not make good drugs because they are susceptible to metabolism in vivo. There are two types of reactions including PhaseIandIImetabolisms. The major reaction of phaseImetabolism includes oxidation and reduction, while the major reactions of phaseIImetabolism contain glucuronidation and sulfation. Both of these types of reactions produce more polar products with higher aqueous solubility, so they are more readily excreted from the body via bile and urine. Metabolism increases clearance, reduces exposure, and is a major cause of low bioavailability. Structural modifications that reduce compound binding or reactivity at the labile site will increase metabolic stability. Strategies commonly include blocking metabolic site by adding other blocking groups, removing labile functional groups and replacing unstable groups. Lipoxin A4 (LXA4, 57) is a structurally and functionally distinct natural eicosanoid with potent immunomodulatory and anti-inflammatory activities [93,94]. 57 has been shown to suppress inflammation upon binding to its receptor (ALXR or hFPRL1), a G-protein-coupled receptor known to play a key role in modulating inflammation [95]. Therapeutic application of 57, however, is greatly hampered due to its chemical instability and rapid metabolism in vivo. First-generation analogues were designed to minimize rapid inactivation of 57 via ω-oxidation or oxidation at the 15(S)-alcohol by prostaglandin dehydrogenase (PGDH) [96]. Among which, 58 was obtained by substituting a fluorinated phenoxy group for the alkyl tail and by inverting the stereochemistry at C-15 to the Risomer (15-epi) based on the fact that 15-epi-LXA4 was equipotent in vitro assays to 57 but was a poorer substrate for PGDH [97]. Although 57 was efficacious in various models, it was cleared from the mouse within 15 min after intravenous injection by β-oxidation [98]. To overcome this catabolic pathway, a second generation 3-oxa-LXA4 analogue 59 was designed and synthesized, which possessed a single stable pharmacophore but still had a short half-life in vivo due to its light- or acid-sensitive trihydroxytetraene structure [99]. Replacement of the tetraene moiety with the trienyne unit led to a more stable analogue 60, which is a topically and orally active anti-inflammatory agent currently under development [99,100].

O

437 438

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

E

6

503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527

Compound 68, a novel triterpene monoglycoside isolated from the extracts of the black cohosh plant (Actaea racemosa), acts as a γ-secretase modulator showing a comparatively selective pharmacology (Aβ42 IC50 = 0.1 μM, Aβ40 IC50 = 2.7 μM) via screening of natural product extracts for

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

528 529 530 531 532

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

537 538 539 540 541 542 543 544 545 546

lines containing K562, MCF-7, Bel-7402, and MGC-803 (IC50 = 1.12, 0.68, 0.50 1.09 μM, respectively) than that of 7 (IC50 = 4.76, 14.60, 7.48 and 5.69 μM, respectively) and benzoic acid mustard (IC50 N 140 μM), and also exhibited an approximately 8-fold higher selective cytotoxicity toward the cancer cells, which was higher than those of 7 (2.5-fold) and clinically used chlorambucil (2-fold) [119].

547

560 561 562 563 564 565 566 567 568 569 570 571 572

U

573

574 575 576 577 578

P

D

E T

C

558 559

E

556 557

R

554 555

R

552 553

Oleanolic acid (78), a triterpenoid compound extracted from many Asian herbs, such as Fructus ligustri lucidi, Fructus forsythiae, Radix ginseng, and Akebia trifoliate, exhibits hepatoprotective, anti-inflammatory, antitumor, antioxidant and antiglycative activities [120]. 78 has been shown to inhibit tumor initiation and promotion, and to induce tumor cell differentiation and apoptosis. Despite its positive mode of action, the antitumor activity of 78 is relatively weak. By introducing a nitrate NO donor and L-valine at the 3-positon and 28-positon of 78, compound 79 showed effective activities against four human cancer cell lines (HepG2, MCF-7, A549, HCT-116), especially in A549 cells with an IC50 value 2-fold lower (IC50 = 7.5 μM) than that of the positive control cisplatin (IC50 = 16.5 μM) [121]. The introduction of a nitrate NO donor improves the antitumor activity because of the synergistic action of the cytotoxic NO donor and 78. Furthermore, the L-valine moiety can be selectively transported by peptide transporter 1 to the tumor cells, resulting in increase of selectivity. In addition, another promising prodrug 80 obtained through the introduction of a glutathione s-transferase π-activated O2 -diazeniumdiolate NO donor and liver-specific galactose at the 3-positon and 28-positon of 78, significantly improved the potency and selectivity of 78 [122].

583 584 585 586

588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614

O

550 551

2.1.5. Improvement of potency and selectivity The therapeutic effectiveness of most natural drugs today is encumbered by the low potency or undesired side-effects associated with their poor selectivity toward targeted cells. Intensive efforts have been made to endow natural agents with enhanced efficacy and selectivity. (±)-XJP (71), (±)-7, 8-dihydroxy-3-methyl-isochroman4-one, a novel natural polyphenolic compound isolated from the banana (Musa sapientum L.) peel, displayed potent antihypertensive activity in both acute and therapeutic antihypertensive tests in renal hypertensive rats [115,116]. (±)-XJP-B (72), an analogue of 71, was more active than 71 in spontaneously hypertensive rats (SHRs). However, the antihypertensive effects of both 71 and 72 are still not potent enough for therapeutic use. By coupling nitric oxide (NO)-donor moieties with 71 and 72, compounds 73, 74 and 75 reduced blood pressure by nearly 40% in SHRs, which was obviously superior to that of the lead compounds and comparable to that of reference drug captopril [117]. Moreover, by connecting N-substituted isopropanolamine functionalities, the classic side chains of β-blockers, to a phenolic oxygen of 71 or 72, compound 76 was achieved and exhibited potent β1-adrenoceptor blocking effect comparable to the well-known antihypertensive drug propranolol [118].

581 582

N C

548 549

579 580

F

535 536

their ability to reduce Aβ42 relative to Aβ40 in a cell-based assay [112]. However, the high molecular weight (MW) and tPSA (155 Å2, a property that has been strongly correlated with poor CNS permeability) of 68 limits its clinical development for CNS drugs. Replacement of the C3 glycoside with a MW-decreasing and more lipophilic morpholine led to an analogue 69 with significantly improved CNS exposure (B/P = 1.7, Aβ42 IC50 = 0.07 μM, Aβ40 IC50 = 2.3 μM, F = 37%) [113]. Furthermore, conversion of the C24 acetate to an ether and substitution of the C3 morpholine nitrogen with N-methyl cyclobutanamine provide an analogue 70 with a better pharmacokinetic profile (B/P = 0.55, Aβ42 IC50 = 0.1 μM, Aβ40 IC50 = 2.1 μM, F = 75%) [114].

R O O

533 534

7

Although oridonin (7) showed a unique and safe antitumor activity, the development of 7 for cancer therapy is hampered, to a large extent, by its relatively moderate potency. Compound 77 achieved by conjugation of 7 with benzoic acid mustard, showed more potent activities against four human cancer cell

2.1.6. Employment of prodrug strategy The prodrug design is a versatile, powerful method that can be applied to NPs with the aim of increasing aqueous solubility, improving lipophilicity, strengthening metabolic stability, enhancing transporter-mediated absorption, such as 79, and achieving site-specific delivery, such as 80. Rapamycin (81), a macrolide initially isolated from Streptomyces hygroscopicushas [123], is a highly specific mTOR inhibitor displaying antifungal, immunosuppressive, anticancer, neuroprotective and antiaging activities [124]. However, the poor aqueous solubility and metabolic instability impeded clinical development of 81 in cancer therapy [125]. Multiple analogues of 81 have been designed for improving its solubility and stability.

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

615 616 617 618 619 620 621 622 623 624 625 626 627

8

628 629 630 631 632

Temsirolimus (82), designed as a water-soluble ester prodrug with increased stability, demonstrated rapamycin-comparable antineoplastic potency against a panel of National Cancer Institute (NCI) human tumor cell lines, with an IC50 value frequently b10 nM [126]. In 2007, the drug was approved by the FDA for intravenous administration in oncology [127].

derivatives [136], and others being applied for (±)-XJP 675 derivatives [137], β-elemene derivatives [138], etc. 676 3. Conclusion

677

NPs as secondary metabolites have played a key role in drug discovery and constitute a prolific source of novel lead compounds or pharmacophores for medicinal chemistry. As compared to synthetic molecules, NPs possess structural diversity and complexity, more stereogenic centers and fewer halogen or nitrogen atoms. They form a rather heterogeneous class of compounds differentiated by their source organisms, biosphere of origin, and biological role. Due to the high potential to show excellent biological activities, NPs have been the major source of therapeutic agents. Druggability optimization of NPs is intent to increase solubility, potency, stability and selectivity, etc. Under the premise of maintaining or enhancing the activity, selective modifications of NPs overcome corresponding deficiencies and shortcomings according to the principles and rules of pharmacology and medicinal chemistry, advancing NPs from leads to available drugs.

678

Acknowledgments

694

This study was financially supported by a grant from the National Natural Science Funds (No. 81373280; 81302635; 81273377) and the Project Program of State Key Laboratory of Natural Medicines, China Pharmaceutical University (No. SKLNMZZCX201404).

695

References

700

649 650 651 652 653 654 655 656

C

658

664 665 666 667 Q3 668 669 Q4 670 671 672 673 674

2.1.7. The intellectual property attainment status of NPs NPs are now harder to get patented, which greatly affects NP intellectual property strategy. The US Patent and Trademark Office (USPTO) on March 4, 2014 issued new guidelines that natural law, material or phenomenon could not obtain patent protection [133]. Known NPs discovered with new activity or even NPs isolated first won't boost the chances of patent eligibility. While natural analogues that be modified with substituents, side chains, point mutations and so on, overstep the rule [134]. Thus it is essential for proper modification of natural structures so as to attain intellectual properties. Recently, we have carried out a large body of investigations on the structural modification of NPs and some achievements have also been made including three granted patents for oridonin derivatives [135], four for 23-hydroxybetulinic acid

U

662 663

N

659

660 661

P

D

T

647 648

C

645 646

E

643 644

R

641 642

R

639 640

Berberine (83), an isoquinoline alkaloid extracted from Chinese herbs Coptis chinensis, has been extensively used as traditional medicines for the treatment of gastroenteritis and secretory diarrhea [128], and has also been found to have a variety of pharmacological and biological activities, such as antifungal, antineoplastic, anti-inflammatory, anti-hyperglycemic and anti-hyperlipidemic effects [129]. However, for its hydrophilic in nature, 83 is absorbed poorly in the intestines, thus leading to low oral bioavailability [130]. Berberrubine (84), which is an active metabolite of 83 after first pass metabolism, showed a slightly lower lipidlowing activity than that of 83 [131,132]. The ester or ether prodrugs were designed and prepared with improved lipophilicity and oral bioavailability through modification of phenolic hydroxyl of 84. Among them, compound 85 bearing a palmitate exhibited a moderate Log P value and hydrolysis rate in blood, and reduced blood CHO and LDL-c by 35.8% and 45.5%, respectively, similar to that by 83 [132].

O

637 638

E

636

R O

O

F

633 634

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

[1] Baker DD, Chu M, Oza U, Rajgarhia V. The value of natural products to future pharmaceutical discovery. Nat Prod Rep 2007;24:1225–44. [2] Cragg GM, Newman DJ. Natural products: a continuing source of novel drug leads. Biochim Biophys Acta 1830;2013:3670–95. [3] Newman DJ, Cragg GM. Natural products as sources of new drugs over the 30 years from 1981 to 2010. J Nat Prod 2012;75:311–35. [4] Mishra BB, Tiwari VK. Natural products: an evolving role in future drug discovery. Eur J Med Chem 2011;46:4769–807. [5] Galloway WR, Diáz-Gavilán M, Isidro-Llobet A, Spring DR. Synthesis of unprecedented scaffold diversity. Angew Chem Int Ed 2009;48:1194–6. [6] Clardy J, Walsh C. Lessons from natural molecules. Nature 2004;432: 829–37. [7] Feher M, Schmidt JM. Property distributions: differences between drugs, natural products, and molecules from combinatorial chemistry. J Chem Inf Comput Sci 2003;43:218–27. [8] Henkel T, Brunne RM, Müller H, Reichel F. Statistical investigation into the structural complementarity of natural products and synthetic compounds. Angew Chem Int Ed 1999;38:643–7. [9] Rosén J, Gottfries J, Muresan S, Backlund A, Oprea TI. Novel chemical space exploration via natural products. J Med Chem 2009;52:1953–62. [10] Eberhardt L, Kumar K, Waldmann H. Exploring and exploiting biologically relevant chemical space. Curr Drug Targets 2011;12:1531–46. [11] Grabowski K, Schneider G. Properties and architecture of drugs and natural products revisited. Curr Chem Biol 2007;1:115–27. [12] Hert J, Irwin JJ, Laggner C, Keiser MJ, Shoichet BK. Quantifying biogenic bias in screening libraries. Nat Chem Biol 2009;5:479–83. [13] Keseru GM, Makara GM. The influence of lead discovery strategies on the properties of drug candidates. Nat Rev Drug Discov 2009;8:203–12. [14] Dandapani S, Marcaurelle LA. Grand challenge commentary: accessing new chemical space for ‘undruggable’ targets. Nat Chem Biol 2010;6: 861–3. [15] Quinn RJ, Carroll AR, Pham NB, Baron P, Palframan ME, Suraweera L, et al. Developing a druglike natural product library. J Nat Prod 2008;71:464–8. [16] Breinbauer R, Manger M, Scheck M, Waldmann H. Natural product guided compound library development. Curr Med Chem 2002;9:2129–45. [17] Morrison KC, Hergenrother PJ. Natural products as starting points for the synthesis of complex and diverse compounds. Nat Prod Rep 2014;31: 6–14.

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

679 680 681 682 683 684 685 686 687 688 689 690 691 692 693

696 697 698 699

701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

N C

O

R

R

E

E

D

P

R O O

F

[45] Chen F, Chai H, Su MB, Zhang YM, Li J, Xie X, et al. Potent and orally efficacious bisthiazole-based histone deacetylase inhibitors. ACS Med Chem Lett 2014;5:628–33. [46] Hirata Y, Uemura D. Halichondrins—antitumor polyether macrolides from a marine sponge. Pure Appl Chem 1986;58:701–10. [47] Dabydeen D, Burnett JC, Bai R, Verdier-Pinard P, Hickford SJH, Pettit GR, et al. Comparison of the activities of the truncated halichondrin B analog NSC 707389 (E7389) with those of the parent compound and a proposed binding site on tubulin. Mol Pharmacol 2006;70:1866–75. [48] Overmoyer B. Options for the treatment of patients with taxanerefractory metastatic breast cancer. Clin Breast Cancer 2008;8: S61–70. [49] Yu MJ, Zheng W, Seletsky BM, Littlefield BA, Kishi Y. Case history. Annu Rep Med Chem 2011;46:227–41. [50] Seletsky BM, Wang Y, Hawkins LD, Palme MH, Habgood GJ, DiPietro LV, et al. Structurally simplified macrolactone analogues of halichondrin B. Bioorg Med Chem Lett 2004;14:5547–50. [51] Zheng W, Seletsky BM, Palme MH, Lydon PJ, Singer LA, Chase CE, et al. Macrocyclic ketone analogues of halichondrin B. Bioorg Med Chem Lett 2004;14:5551–4. [52] Aftimos P, Awada A. Survival benefit of eribulin mesylate in heavily pretreated metastatic breast cancer: what next? Adv Ther 2011;28: 973–85. [53] Fujita T, Inoue K, Yamamoto S, Ikumoto T, Sasaki S, Toyama R, et al. S Fungal metabolites. Part 11. A potent immunosuppressive activity found in Isaria sinclairii metabolite. J Antibiot 1994;47:208–15. [54] Strader CR, Pearce CJ, Oberlies NH. Fingolimod (FTY720): a recently approved multiple sclerosis drug based on a fungal secondary metabolite. J Nat Prod 2011;74:900–7. [55] Fujita T, Yoneta M, Hirose R, Sasaki S, Inoue K, Kiuchi M, et al. Simple compounds, 2-alkyl-2-amino-1,3-propanediols have potent immunosuppressive activity. Bioorg Med Chem Lett 1995;5:847–52. [56] Fujita T, Hamamichi N, Kiuchi M, Matsuzaki T, Kitao Y, Inoue K, et al. Determination of absolute configuration and biological activity of new immunosuppressants, mycestericins D, E, F and G. J Antibiot 1996;49: 846–53. [57] Kataoka H, Sugahara K, Shimano K, Teshima K, Koyama M, Fukunari A, et al. FTY720, sphingosine 1-phosphate receptor modulator, ameliorates experimental autoimmune encephalomyelitis by inhibition of T cell infiltration. Cell Mol Immunol 2005;2:439–48. [58] Pelletier D, Hafler DA. Fingolimod for multiple sclerosis. N Engl J Med 2012;366:339–47. [59] Marion F, Williams DE, Patrick BO, Hollander I, Mallon R, Kim SC, et al. Liphagal, a selective inhibitor of PI3 kinaserisolated from the sponge aka coralliphaga: structure elucidation and biomimetic synthesis. Org Lett 2006;8:321–4. [60] Pereira AR, Strangman WK, Marion F, Feldberg L, Roll D, Mallon R, et al. Synthesis of phosphatidylinositol 3-kinase(PI3k)inhibitory analogues of the sponge meroterpenoid liphagal. J Med Chem 2010;53:8523–33. [61] Govindarajan VS. Turmeric—chemistry, technology, and quality. Crit Rev Food Sci Nutr 1980;12:199–301. [62] Salem M, Rohani S, Gillies ER. Curcumin, a promising anti-cancer therapeutic: a review of its chemical properties, bioactivity and approaches to cancer cell delivery. RSC Adv 2014;4:10815–29. [63] Tomren MA, Masson M, Loftsson T, Tonnesen HH. Studies on curcumin and curcuminoids XXXI. Symmetric and asymmetric curcuminoids: stability, activity and complexation with cyclodextrin. Int J Pharm 2007; 338:27–34. [64] Leung MH, Kee TW. Effective stabilization of curcumin by association to plasma proteins: human serum albumin and fibrinogen. Langmuir 2009; 25:5773–7. [65] Wang YJ, Pan MH, Cheng AL, Lin LI, Ho YS, Hsieh CY, et al. Stability of curcumin in buffer solutions and characterization of its degradation products. Pharm Biomed Anal 1997;15:1867–78. [66] Simon A, Allais DP, Duroux JL, Basly JP, Durand Fontanier S, Delage C. Inhibitory effect of curcuminoids on MCF-7 cell proliferation and structure–activity relationships. Cancer Lett 1998;129:111–6. [67] Dinkova-Kostova AT, Talalay P. Relation of structure of curcumin analogs to their potencies as inducers of Phase 2 detoxification enzymes. Carcinogenesis 1999;20:911–4. [68] Liang G, Shao LL, Wang Y, Zhao CG, Chu YH, Xiao J, et al. Exploration and synthesis of curcumin analogues with improved structural stability both in vitro and in vivo as cytotoxic agents. Bioorg Med Chem 2009;17: 2623–31. [69] Chen L, Magesh S, Wang H, Yang CS, Kong ANT, Hu LQ. Design and synthesis of novel iminothiazinylbutadienols and divinylpyrimidinethiones as ARE inducers. Bioorg Med Chem 2014;24:940–3. [70] Chakraborti S, Dhar G, Dwivedi V, Das A, Poddar A, Chakraborti G, et al. Stable and potent analogues derived from the modification of the dicarbonyl moiety of curcumin. Biochemistry 2013;52:7449–60.

T

C

[18] Carlson EE. Natural products as chemical probes. ACS Chem Biol 2010;5: 639–53. [19] von Nussbaum F, Brands M, Hinzen B, Weigand S, Hbich D. Antibacterial natural products in medicinal chemistry—exodus or revival ? Angew Chem Int Ed 2006;45:5072–129. [20] Lipinski CA, Lombardo F, Dominy BW, Feeney PJ. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv Drug Deliv Rev 2012;64:4–17. [21] Koehn FE. Biosynthetic medicinal chemistry of natural product drugs. Med Chem Commun 2012;3:854–65. [22] Curatolo W. Physical chemical properties of oral drug candidates in the discovery and exploratory development settings. Pharm Sci Technol Today 1998;1:387–93. [23] Liu JM, Ni MY, Fan YF, Tu YY, Wu ZH, Wu YL, et al. The structure and reactions of arteannuin. Acta Chim Sin 1979;37:129–41. [24] Zhu SQ, Liu WH, Ke XX, Li JF, Hu RJ, Cui HJ, et al. Artemisinin reduces cell proliferation and induces apoptosis in neuroblastom. Oncol Rep 2014;32: 1094–100. [25] He R, Mott BT, Rosenthal AS, Genna DT, Posner GH, Arav-Boger R. An artemisinin-derived dimer has highly potent anti-cytomegalovirus (CMV) and anti-cancer activities. PLoS One 2011;6:e24334. [26] Lin AJ, Klayman DL, Milhous WK. Antimalarial activity of new watersoluble dihydroartemisinin derivatives. J Med Chem 1987;30:2147–50. [27] Li Ying, Zhu YM, Jiang HJ, Pan JP, Wu GS, Wu JM. Synthesis and antimalarial activity of artemisinin derivatives containing an amino group. J Med Chem 2000;43:1635–40. [28] Haynes RK, Fugmann B, Stetter J, Rieckmann K, Heilmann HD, Chan HW, et al. Artemisone—a highly active antimalarial drug of the artemisinin class. Angew Chem Int Ed 2006;45:2082–8. [29] Fujita E, Nagao Y, Kaneko K, Nakazawa S, Kuroda H. The antitumor and antibacterial activity of the Isodon diterpenoids. Chem Pharm Bull 1976; 24:2118–27. [30] Li CY, Wang EQ, Cheng Y, Bao JK. Oridonin: an active diterpenoid targeting cell cycle arrest, apoptotic and autophagic pathways for cancer therapeutics. Int J Biochem Cell Biol 2011;43:701–4. [31] Zhou G, Kang H, Wang L, Gao L, Liu P, Xie J, et al. Oridonin, a diterpenoid extracted from medicinal herbs, targets AML1-ETO fusion protein and shows potent antitumor activity with low adverse effects on t(8;21) leukemia in vitro and in vivo. Blood 2007;109:3441–50. [32] Bohanon FJ, Wang XF, Ding CY, Ding Y, Radhakrishnan GL, Rastellini C, et al. Oridonin inhibits hepatic stellate cell proliferation and fibrogenesis. J Surg Res 2014;190:55–63. [33] Xu ST, Li DH, Pei LL, Yao H, Wang CQ, Can H, et al. Design, synthesis and antimycobacterial activity evaluation of natural oridonin derivatives. Bioorg Med Chem Lett 2014;24:2811–4. [34] Xu JY, Yang JY, Ran Q, Wang L, Liu J, Wang ZX, et al. Synthesis and biological evaluation of novel 1-O- and 14-O-derivatives of oridonin as potential anticancer drug candidates. Bioorg Med Chem Lett 2008;18:4741–4. [35] Ding CY, Zhang YS, Chen HJ, Yang ZD, Wild C, Chu LL, et al. Novel nitrogen-enriched oridonin analogues with thiazole-fused A-ring: protecting group-free synthesis, enhanced anticancer profile, and improved aqueous solubility. J Med Chem 2013;56:5048–58. [36] Wall ME, Wani MC, Cook CE, Palmer KH, McPhail AT, Sim GA. Plant antitumor agents. I. The isolation and structure of camptothecin, a novel alkaloidal leukemia and tumor inhibitor from camptotheca acuminate. J Am Chem Soc 1966;88:3888–90. [37] Oberlies NH, Kroll DJ. Camptothecin and taxol: historic achievements in natural products research. J Nat Prod 2004;67:129–35. [38] Liew ST, Yang LX. Design, synthesis and development of novel camptothecin drugs. Curr Pharm Des 2008;14:1078–97. [39] Ohwada J, Ozawa S, Kohchi M, Fukuda H, Murasaki C, Suda H, et al. Synthesis and biological activities of a pH-dependently activated watersoluble prodrug of a novel hexacyclic camptothecin analog. Bioorg Med Chem Lett 2009;19:2772–6. [40] Li QY, Deng XQ, Zu YG, Lv HY, Su L, Yao LP, et al. Cytotoxicity and topo I targeting activity of substituted 10—nitrogenous heterocyclic aromatic group derivatives of SN-38. Eur J Med Chem 2010;45:3200–6. [41] Lachance H, Wetzel S, Kumar K, Waldmann H. Charting, navigating, and populating natural product chemical space for drug discovery. J Med Chem 2012;55:5989–6001. [42] Taori K, Paul VJ, Luesch H. Structure and activity of largazole, a potent antiproliferative agent from the Floridian marine cyanobacterium Symploca sp.. J Am Chem Soc 2008;130:1806–7. [43] Souto JA, Vaz E, Lepore I, Poppler AC, Franci G, Alvarez R, et al. Synthesis and biological characterization of the histone deacetylase inhibitor largazole and C7-modified analogues. J Med Chem 2010;53:4654–67. [44] Bowers AA, West N, Newkirk TL, Troutman-Youngman AE, Schreiber SL, Wiest O, et al. Synthesis and histone deacetylase inhibitory activity of largazole analogs: alteration of the zinc-binding domain and macrocyclic scaffold. Org Lett 2009;11:1301–4.

U

739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818

9

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898

[100]

[101] [102] [103]

P

[104]

C

E

R

R

O

C

N

F

[99]

O

[98]

leukotriene B4 12-hydroxydehydrogenase is a multifunctional eicosanoid oxidoreductase. J Biol Chem 2000;275:25372–80. Phillips ED, Chang HF, Holmquist CR, McCauley JP. Synthesis of methyl(5S,6R,7E,9E,11Z,13E,15S)-16-(4-fluorophenoxy)-5,6,15-trihydroxy-7,9,11,13-hexadecatetraenoate, an analogue of 15R-lipoxin A4. Bioorg Med Chem Lett 2003;13:3223–6. Clish CB, O'Brien JA, Gronert K, Stahl GL, Petasis NA, Serhan CN. Local and systemic delivery of a stable aspirin-triggered lipoxin prevents neutrophil recruitment in vivo. Proc Natl Acad Sci U S A 1999;96:8247–52. Guilford WJ, Bauman JG, Skuballa W, Bauer S, Wei GP, Davey D, et al. Novel 3-oxa lipoxin A4 analogues with enhanced chemical and metabolic stability have anti-inflammatory activity in vivo. J Med Chem 2004;47: 2157–65. Levy BD, Lukacs NW, Berlin AA, Schmidt B, Guilford WJ, Serhan CN, et al. Lipoxin A4 stable analogs reduce allergic airway responses via mechanisms distinct from CysLT1 receptor antagonism. FASEB J 2007;21: 3877–84. Petersen C. Analyse des Phloridzins. Ann Acad Sci Fr 1835;15:178. Ehrenkranz JRL, Lewis NG, Kahn CR, Roth J. Phlorizin: a review. Diabetes Metab Res Rev 2005;21:31–8. Tsujihara K, Hongu M, Saito K, Kawanishi H, Kuriyama K, Matsumotu M, et al. Na(+)-glucose cotransporter (SGLT) inhibitors as antidiabetic agents. 4. Synthesis and pharmacological properties of 4'-dehydroxyphlorizin derivatives substituted on the B ring. J Med Chem 1999;42:5311–24. Katsuno K, Fujimori Y, Takemura Y, Hiratochi M, Itoh F, Komatsu Y, et al. Sergliflozin, a novel selective inhibitor of low-affinity sodium glucose cotransporter (SGLT2), validates the critical role of SGLT2 in renal glucose reabsorption and modulates plasma glucose level. J Pharmacol Exp Ther 2007;320:323–30. Isaji M. Bone: from a reservoir of minerals to a regulator of energy metabolism. Kidney Int Suppl 2011;79:S14–9. Meng M, Ellsworth BA, Nirschl AA, McCann PJ, Patel M, Girotra RN, et al. Discovery of dapagliflozin: a potent, selective renal sodium-dependent glucose cotransporter 2 (SGLT2) inhibitor for the treatment of type 2 diabetes. J Med Chem 2008;51:1145–9. Nomura S, Sakamaki S, Hongu M, Kawanishi E, Koga Y, Sakamoto T, et al. Discovery of canagliflozin, a novel Cglucoside with thiophene ring, as sodium-dependent glucose cotransporter 2 inhibitor for the treatment of type 2 diabetes mellitus. J Med Chem 2010;53:6355–60. Jordan MA. Mechanism of action of antitumor drugs that interact with microtubules and tubulin. Curr Med Chem Anticancer Agents 2002;2:1–7. Kemper EM, van Zandbergen AE, Cleypool C, Mos HA, Boogerd W, Beijnen JH, et al. Increased penetration of paclitaxel into the brain by inhibition of P-glycoprotein. Clin Cancer Res 2003;9:2849–55. Fischer H, Gottschlich R, Seelig A. Blood–brain barrier permeation: molecular parameters governing passive diffusion. J Membr Biol 1998; 165:201–11. Antonie R, Liu YB, Michaelis ML, Himes RH, Georg GI, Audus KL. Chemical modification of paclitaxel (Taxol) reduces P-glycoprotein interactions and increases permeation across the blood–brain barrier in vitro and in situ. J Med Chem 2005;48:832–8. Findeis MA, Schroeder F, McKee TD, Yager D, Fraering PC, Creaser SP, et al. Discovery of a novel pharmacological and structural class of gamma secretase modulators derived from the extract of Actaea racemosa. ACS Chem Neurosci 2012;3:941–51. Fuller NO, Hubbs JL, Austin WF, Creaser SF, McKee TD, Loureiro RMB, et al. Initial optimization of a new series of γ secretase modulators derived from a triterpene glycoside. ACS Med Chem Lett 2012;3:908–13. Hubbs JL, Fuller NO, Austin WF, Shen RC, Creaser SP, McKee TD, et al. Optimization of a natural product-based class of γ secretase modulators. J Med Chem 2012;55:9270–82. Qian H, Huang WL, Wu XM, Zhang HB, Zhou JP, Ye WC. A new isochroman-4-one derivative from the peel of Musa sapientum L. and its total synthesis. Chin Chem Lett 2007;18:1227–30. Liu J, Ren H, Xu JY, Bai RR, Yan Q, Huang WL, et al. Total synthesis and antihypertensive activity of (+/-)7,8-dihydroxy-3-methylisochromanone-4. Bioorg Med Chem Lett 2009;19:1822–4. Bai RR, Yang X, Zhu Y, Zhou ZW, Xie WJ, Yao HQ, et al. Novel nitric oxidereleasing isochroman-4-one derivatives: synthesis and evaluation of antihypertensive activity. Bioorg Med Chem 2012;20:6848–55. Bai RR, Huang XJ, Yang X, Hong W, Tang YQ, Yao HQ, et al. Novel hybrids of natural isochroman-4-one bearing N-substituted isopropanolamine as potential antihypertensive candidates. Bioorg Med Chem 2013;21: 2495–502. Xu ST, Pei LL, Wang CQ, Zhang YK, Li DH, Yao HQ, et al. Novel hybrids of natural oridonin-bearing nitrogen mustards as potential anticancer drug candidates. ACS Med Chem Lett 2014;5:797–802. Tsai SJ, Yin MC. Anti-oxidative, anti-glycative and antiapoptotic effects of oleanolic acid in brain of mice treated by dgalactose. Eur J Pharmacol 2012;689:81–8.

R O

[97]

D

[105] [106]

[107]

T

[71] Lin CM, Singh SB, Chu PS, Dempcy RO, Schmidt JM, Pettit GR, et al. Interactions of tubulin with potent natural and synthetic analogs of the antimitotic agent combretastatin: a structure–activity study. Mol Pharmacol 1988;34:200–8. [72] Pettit GR, Toki BE, Herald DL, Boyd MR, Hamel E, Pettit RK, et al. Antineoplastic agents. 410. Asymmetric hydroxylation of transcombretastatin A-4. J Med Chem 1999;42:1459–65. [73] Pirali T, Busacca S, Beltrami L, Imovilli D, Pagliai F, Miglio G, et al. Synthesis and cytotoxic evaluation of combretafurans, potential scaffolds for dual-action antitumoral agents. J Med Chem 2006;49: 5372–6. [74] Cafici L, Pirali T, Condorelli F, Del Grosso E, Massarotti A, Sorba G, et al. Solution-phase parallel synthesis and biological evaluation of combretatriazoles. J Comb Chem 2008;10:732–40. [75] Nathwani SM, Hughes L, Greene LM, Carr M, O'Boyle NM, McDonnell S, et al. Novel cis-restricted β-lactam combretastatin A-4 analogues display anti-vascular and anti-metastatic properties in vitro. Oncol Rep 2013;29: 585–94. [76] Alvarez R, Alvarez C, Mollinedo F, Sierra BG, Medarde M, Peláez R. Isocombretastatins A: 1,1-diarylethenes as potent inhibitors of tubulin polymerization and cytotoxic compounds. Bioorg Med Chem 2009;17: 6422–31. [77] Liou JP, Chang JY, Chang CW, Chang CY, Mahindroo N, Kuo FM, et al. Synthesis and structure–activity relationships of 3-aminobenzophenones as antimitotic agents. J Med Chem 2004;47:2897–905. [78] Barbosa EG, Bega LA, Beatriz A, Sarkar T, Hamel E, do Amaral MS, et al. A diaryl sulfide, sulfoxide, andsulfone bearing structural similarities to combretastatin A-4. Eur J Med Chem 2009;44:2685–8. [79] Gerth K, Bedorf N, Höfle G, Irschik H, Reichenbach H. Epothilons A and B: antifungal and cytotoxic compounds from Sorangium cellulosum (Myxobacteria). Production, physico-chemical and biological properties. J Antibiot 1996;49:560–3. [80] Altaha R, Fojo T, Reed E, Abraham J. Epothilones: a novel class of nontaxane microtubule-stabilizing agents. Curr Pharm Des 2002;8:1707–12. [81] Lee FY, Borzilleri R, Fairchild CR, Kamath A, Smykla R, Kramer R, et al. Preclinical discovery of ixabepilone, a highly active antineoplastic agent. Cancer Chemother Pharmacol 2008;63:157–66. [82] Conlin A, Fornier M, Hudis C, Kar S, Kirkpatrick P. Ixabepilone. Nat Rev Drug Discov 2007;6:953–4. [83] Deeken JF, Marshall JL, Pishvaian MJ, Hwang J, Ahlers CM, Clemens PL, et al. A phase I study of oral Ixabepilone in patients with advanced solid tumors. Cancer Chemother Pharmacol 2014;73:1071–8. [84] Smaglo BG, Pishvaian MJ. Profile and potential of Ixabepilone in the treatment of pancreatic cancer. Drug Des Devel Ther 2014;8:923–30. [85] Waisvisz JM, van der Hoeven MG, van Peppen J, Zwennis WCM, Bottromycin I. A new sulfur-containing antibiotic. J Am Chem Soc 1957; 79:4520–1. [86] Shimamura H, Gouda H, Nagai K, Hirose T, Ichioka M, Furuya Y, et al. Structure determination and total synthesis of bottromycin A2: a potent antibiotic against MRSA and VRE. Angew Chem Int Ed 2009;48:914–7. [87] Kobayashi K, Ichioka M, Hirose T, Nagai K, Matsumoto A, Matsui H, et al. Bottromycin derivatives: efficient chemical modifications of the ester moiety and evaluation of anti-MRSA and anti-VRE activities. Bioorg Med Chem Lett 2010;20:6116–20. [88] Burke TG, Mi Z. Ethyl substitution at the 7 position extends the half-life of 10-hydroxycamptothecin in the presence of human serum albumin. J Med Chem 1993;36:2580–2. [89] Burke TG, Mi Z. The structural basis of camptothecin interactions with human serum albumin: impact on drug stability. J Med Chem 1994;37: 40–6. [90] Bailly C. Homocamptothecins: potent topoisomerase I inhibitors and promising anticancer drugs. Crit Rev Oncol Hematol 2003;45:91–108. [91] Hautefaye P, Cimetière B, Pierré A, Léonce S, Hickman J, Laine W, et al. Synthesis and pharmacological evaluation of novel non-lactone analogues of camptothecin. Bioorg Med Chem Lett 2003;13:2731–5. [92] Miao ZY, Zhu LJ, Dong GQ, Zhuang CL, Wu YL, Wang SZ, et al. A new strategy to improve the metabolic stability of lactone: discovery of (20S, 21S) -21-fluorocamptothecins as novel, hydrolytically stable topoisomerase I inhibitors. J Med Chem 2013;56:7902–10. [93] Serhan CN. Lipoxins and aspirin-triggered 15-epi-lipoxin biosynthesis: an update and role in anti-inflammation and proresolution. Prostaglandins Other Lipid Mediat 2002;68–69:433–55. [94] Levy BD, Serhan CN. Exploring new approaches to the treatment of asthma: potential roles for lipoxins and aspirintriggered lipid mediators. Drugs Today (Barc) 2003;39:373–84. [95] Chiang N, Arita M, Serhan CN. Anti-inflammatory circuitry: lipoxin, aspirin-triggered lipoxins and their receptor ALX. Prostaglandins Leukot Essent Fatty Acids 2005;73:163–77. [96] Clish CB, Levy BD, Chiang N, Tai HH, Serhan CN. Oxidoreductases in lipoxin A4 metabolic inactivation: 15-oxaprostaglandin 13-reductase/

U

899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

E

10

[108] [109]

[110]

[111]

[112]

[113]

[114]

[115]

[116]

[117]

[118]

[119]

[120]

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058

J. Chen et al. / Fitoterapia xxx (2015) xxx–xxx

[133] Guidance for determining subject matter eligibility of claims reciting or involving laws of nature, natural phenomena, & natural products. http:// www.uspto.gov/patent/patents-announcements/guidance-determining-subject-matter-eligibility-claims-reciting-or; 2014. [134] Harrison C. Patenting natural products just got harder. Nat Biotechnol 2014;32:403–4. [135] (a) Xu JY, Wu XM, Yang JY, Ran J, Wang L, Wang ZX, et al. Preparation and application for oridonin derivatives. Chin. Pat. 200710133915.X; (b) Xu JY, Wang L, Zhang YH, Yao HQ, Li DH, Xu ST, et al. Preparation and application for oridonin derivatives as anticancer agents. Chin. Pat. 201010509348.5; (c) Xu JY, Wang L, Zhang YH, Yao HQ, Li DH, Xu ST, et al. Preparation and application for nitric oxide-releasing 14-Ooridonin derivatives. Chin. Pat. 201110178862.X. [136] (a) Wu XM, Xu JY, Bi Y, Zhou JP, Zhang LY, Yuan ST, et al. Preparation and application for 23-hydroxybetulinic acid derivatives. Chin. Pat. 200610040277.2; (b) Wu XM, Xu JY, Bi Y, Zhou JP, Zhang LY, Yuan ST, et al. Preparation and application for 23-hydroxybetulinic acid derivatives. Chin. Pat. 200810008418.1; (c) Wu XM, Xu JY, Zhu PQ, Yao HQ, Li Z, Sun F, et al. Preparation and application for 3,23,28- 23-hydroxybetulinic acid derivatives. Chin. Pat. 201010224698.7; (d) Wu XM, Xu JY, Zhu PQ, Yao HQ, Li Z, Sun F, et al. Preparation and application for heterocycyclefused 23-hydroxybetulinic acid derivatives. Chin. Pat. 201010224690.0. [137] (a) Xu JY, Bai RR, Jiang KT, Wu XM, Zhu Y, Xu D, et al. Preparation and application for nitric oxide-releasing isochroman-4-one derivatives as antihypertensive agents. Chin. Pat. 201210132644.7; (b) Xu JY, Bai RR, Liu J, Wu XM, Yang X, Hong W, et al. Preparation and application for isochroman-4-one bearing N-substituted as antihypertensive agents. Chin. Pat. 201210422187.5. [138] (a) Shang J, Xu JY, Xu H, Wang RY, Chen JC, Duan WL, et al. Preparation for 13-β-elemene derivatives and their application in treatment of atherosclerosis. Chin. Pat. 201410376842.7; (b) Shang J, Xu JY, Xu H, Wang RY, Duan WL, Chen JC, et al. Preparation for 14-β-elemene derivatives and their application in treatment of atherosclerosis. Chin. Pat. 201410377535.0.

D

P

R O O

F

[121] Fang L, Wang M, Gou SH, Liu XY, Zhang H, Cao F. Combination of amino acid/dipeptide with nitric oxide donating oleanolic acid derivatives as Pept1 targeting antitumor prodrugs. J Med Chem 2014;57:1116–20. [122] Fu JJ, Liu L, Huang ZJ, Lai YS, Ji H, Peng SX, et al. Hybrid molecule from O2 (2,4-dinitrophenyl)diazeniumdiolate and oleanolic acid: a glutathione s-transferase π-activated nitric oxide prodrug with selective antihuman hepatocellular carcinoma activity and improved stability. J Med Chem 2013;56:4641–55. [123] Vézina C, Kudelski A, Sehgal SN. Rapamycin (AY-22,989), a new antifungal antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle. J Antibiot 1975;28:721–6. [124] Park SR, Yoo YJ, Ban YH, Yoon YJ. Biosynthesis of rapamycin and its regulation: past achievements and recent progress. J Antibiot 2010;63: 434–41. [125] Boni JP, Hug B, Leister C, Sonnichsen D. Intravenous temsirolimus in cancer patients: clinical pharmacology and dosing considerations. Semin Oncol 2009;36:S18–25. [126] Ma WW, Jimeno A. Temsirolimus. Drugs Today (Barc) 2007;43:659–69. [127] Rini B, Kar S, Kirkpatrick P. Temsirolimus. Nat Rev Drug Discov 2007;6: 599–600. [128] Chinese Pharmacopeia Committee. Pharmacopeia of People's Republic of China, 1. Beijing: Chemical Industry Press; 2005 213–4. [129] Li P, Zhu WL, Chen KX. Advances in the study of berberine and its derivatives. Acta Pharm Sin 2008;43:773–87. [130] Zhou Y, Cao S, Wang Y, Xu P, Yan J, Bin W, et al. Berberine metabolites could induce low density lipoprotein receptor up-regulation to exert lipid-lowering effects in human hepatoma cells. Fitoterapia 2014;92: 230–7. [131] Li YH, Li Y, Yang P, Kong WJ, You XF, Ren G, et al. Design, synthesis, and cholesterol-lowering efficacy for prodrugs of berberrubine. Bioorg Med Chem 2010;18:6422–8. [132] Chen CM, Chang HC. Determination of berberine in plasma, urine and bile by high-performance liquid chromatography. J Chromatogr B 1995;665: 117–23.

E

1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092

11

U

N C

O

R

R

E

C

T

1127

Please cite this article as: Chen J, et al, Insights into drug discovery from natural products through structural modification, Fitoterapia (2015), http://dx.doi.org/10.1016/j.fitote.2015.04.012

1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126

Insights into drug discovery from natural products through structural modification.

Natural products (NPs) have played a key role in drug discovery and are still a prolific source of novel lead compounds or pharmacophores for medicina...
2MB Sizes 0 Downloads 11 Views