Structural Simplification of Natural Products

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

Review

Cite This: Chem. Rev. XXXX, XXX, XXX−XXX pubs.acs.org/CR

Structural Simplification of Natural Products


Shengzheng Wang,†,‡,# Guoqiang Dong,†,# and Chunquan Sheng*,†

Department of Medicinal Chemistry, School of Pharmacy, Second Military Medical University, 325 Guohe Road, Shanghai,
200433, P.R. China

Department of Medicinal Chemistry, School of Pharmacy, Fourth Military Medical University, 169 Changle West Road, Xi’an,
710032, P.R. China

ABSTRACT: Natural products (NPs) are important sources of clinical drugs due to
their structural diversity and biological prevalidation. However, the structural
Downloaded via UNIV OF MASSACHUSETTS AMHERST on February 7, 2019 at 22:18:25 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

complexity of NPs leads to synthetic difficulties, unfavorable pharmacokinetic profiles,


and poor drug-likeness. Structural simplification by truncating unnecessary
substructures is a powerful strategy for overcoming these limitations and improving
the efficiency and success rate of NP-based drug development. Herein, we will provide
a comprehensive review of the structural simplification of NPs with a focus on design
strategies, case studies, and new technologies. In particular, a number of successful
examples leading to marketed drugs or drug candidates will be discussed in detail to illustrate how structural simplification is
applied in lead optimization of NPs.

CONTENTS 3.3.1.5. Structural Simplification of Sanguinar-


ine S
1. Introduction B 3.3.1.6. Structural Simplification of Dynemicin S
2. Strategies for the Structural Simplification of NPs D 3.3.1.7. Structural Simplification of Merrilac-
3. Structural Simplification of NPs: Case Studies E tone A T
3.1. Structural Simplification Leading to Clinical 3.3.1.8. Structural Simplification of Pyripyro-
Drugs E pene A T
3.1.1. Structural Simplification of Morphine: A 3.3.1.9. Structural Simplification of Pladieno-
Classic Story in Medicinal Chemistry G lide B U
3.1.2. Structural Simplification of Schisandrin 3.3.1.10. Structural Simplification of Melico-
C: Discovery of Bicyclol I bisquinolinone B U
3.1.3. Structural Simplification of Myriocin: 3.3.1.11. Structural Simplification of Campo-
Discovery of Fingolimod I tothecin V
3.1.4. Structural Simplification of Trichostatin 3.3.1.12. Structural Simplification of Yohim-
A: Discovery of Vorinostat J bine V
3.1.5. Structural Simplification of Halichon- 3.3.1.13. Structural Simplification of Withano-
drin B: Discovery of Eribulin J lide A W
3.2. Structural Simplification Leading to Clinical 3.3.2. Structural Simplification by Removing
Candidates K Chiral Centers W
3.2.1. Structural Simplification of Staurospor- 3.3.2.1. Structural Simplification of Podophyl-
ine: Discovery of Ruboxistaurin and lotoxin W
Enzastaurin K 3.3.2.2. Structural Simplification of Taxuspine
3.2.2. Structural Simplification of Asperlicin: X AA
Discovery of Devazepide L 3.3.2.3. Structural Simplification of Cortistatin
3.3. Structural Simplification Leading to Preclin- A AA
ical Candidates or Lead Compounds M 3.3.2.4. Structural Simplification of Bryostatin
3.3.1. Structural Simplification by Reducing 1 AA
the Ring Number N 3.3.2.5. Structural Simplification of Cyclop-
3.3.1.1. Structural Simplification of Trabecte- amine 1 AA
din N 3.3.2.6. Structural Simplification of Carolacton AA
3.3.1.2. Structural Simplification of Erythri- 3.3.2.7. Structural Simplification of Promysalin AB
nane P
3.3.1.3. Structural Simplification of Sampan-
gine P
3.3.1.4. Structural Simplification of Gambogic Received: August 9, 2018
Acid Q

© XXXX American Chemical Society A DOI: 10.1021/acs.chemrev.8b00504


Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3.3.3. Structural Simplification by Truncating biological prevalidation and structural diversity have made NPs
Unnecessary Substructures AB valuable lead compounds for drug discovery, particularly in the
3.3.3.1. Structural Simplification of Chaetocin AB field of cancerous and infectious diseases.19
3.3.3.2. Structural Simplification of Largazole AB Currently, three strategies have been widely used for NP-
3.3.3.3. Structural Simplification of Lobata- based drug discovery and development (Figure 1). The first
mide C AB strategy is the developing and marketing the NP itself.20 On
3.3.3.4. Structural Simplification of Sanglifeh- the basis of the small-molecule drugs approved by the FDA
rin A AB from 1981 to 2014, the percentage of unaltered NPs is 6% of
3.3.3.5. Structural Simplification of Anguino- the drugs.7 However, most NP-derived drugs are NP
mycin AC derivatives, which account for 26% of FDA approved small-
3.3.3.6. Structural Simplification of Capraza- molecule drugs.7 The shortcomings of unmodified NPs mainly
mycin B AC include limited compound availability, poor solubility, and
3.3.3.7. Structural Simplification of Militari- metabolic instability.21 More importantly, the structural
none D AC complexity of NPs results in synthetic difficulties and
3.3.3.8. Structural Simplification of Bistramide unfavorable absorption, distribution, metabolism, excretion,
A AD and toxicity (ADMET) properties.18 Due to their complex
3.3.3.9. Structural Simplification of Withaferin chemical structures, NPs rarely meet the demands for drug-
A AD likeness (e.g., Lipinski’s rule-of-five, RO5), and thus their drug
3.3.3.10. Structural Simplification of Rakicidin development is always hampered by low absorption from the
A AD gut into the blood and poor oral bioavailability.22 In addition,
3.3.3.11. Structural Simplification of Spongis- clarifying the mode of action and molecular target of NPs also
tatins AD remains a major bottleneck.
3.3.3.12. Structural Simplification of Migrasta- Therefore, structural modification of NPs is generally
tin AE required before they can be developed into clinical drugs
3.3.4. Recent Examples of Structural Simplifi- (the second strategy in Figure 1).21−23 For structural
cation AE optimization of NPs, a general procedure includes the design,
4. Conclusions and Perspectives AE synthesis, and evaluation of analogues of NPs, and clarifying
Author Information AE the structure−activity relationship (SAR).21 Selective mod-
Corresponding Author AE ifications can be designed to improve the pharmacokinetic
ORCID AE (PK)/pharmacodynamic (PD) properties, such as potency,
Author Contributions AE selectivity, water solubility, metabolic stability, and oral
Notes AE bioavailability. Moreover, reducing side effects is also an
Biographies AE important goal in structural optimization. This strategy has
Acknowledgments AF been successfully applied in the development of a number of
Abbreviations AF clinical therapeutics to treat various human diseases.
References AF NPs generally have a structurally complex framework and a
high molecular weight (MW), leading to synthetic difficulty
and an unfavorable ADMET profile. In some cases, standard
1. INTRODUCTION structural modification and SAR approaches fail to generate
Natural products (NPs) have historically been the primary drug-like NP derivatives because the PK limitations are always
source of medicines for the treatment of a wide range of caused by the “heavy” or “fat” structures of NPs. Moreover, a
human disease.1,2 In modern drug discovery and development, number of bioactive NPs (e.g., marine NPs) are still
NPs continue to play a vital role and inspire innovative synthetically intractable due to the highly complex structures.
research in chemistry, biology, and medicine.3−6 Detailed In these circumstances, simplifying the complex structures of
analysis of the new drugs approved by the U.S. Food and Drug NPs without interfering with the desired biological activity
Administration (FDA) between 1981 and 2014 revealed that provides an alternative approach for both NP lead optimization
more than half of the clinical drugs were derived from NPs or and the development of new generation of NP-based drugs
their synthetic derivatives.7 Moreover, NPs are also rich (the third strategy in Figure 1). Different from standard NP
sources of chemical probes for chemical biology studies SAR, structural simplification could lead to new, simpler, and
investigating biological mechanisms.8−11 synthetically more accessible compounds that show improved
NPs are biologically prevalidated because they are produced drug-likeness. However, the structural simplification of NPs is
by evolutionary selection.12 During their biosynthesis, NPs a highly challenging task because maintaining potency and
have already interacted with various enzymes and proteins, selectivity may be daunting and the molecular targets (or
which are similar to the environment of the drug targets.13 binding modes) of the NPs are typically unknown at the early
Thus, NPs inherently fall into biologically relevant chemical stage of lead optimization. Different from standard NP
space as their molecular scaffolds (or pharmacophores) have modifications focusing on derivatization, simplification design
evolved as preferred ligand-target binding motifs, which are is generally based on the SAR obtained from step-by-step
often considered privileged structures in drug discovery.14 truncation of NP structures.4,24 After the pharmacophores
Structurally, NPs possess unique features in terms of diversity essential to the desired activity are identified, substructures
and complexity.15 Compared with synthetic compounds, NPs that are unnecessary for the biological activity can be removed
generally have more stereogenic centers, fewer rotatable bonds, or modified. This strategy has been successfully used in the
more oxygen atoms, fewer nitrogen, sulfur, and halogen atoms, lead optimization of NPs and yielded a number of marketed
more fused, bridged, or spiro rings, and so on.16−18 This drugs and drug candidates.
B DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 1. Strategies for NP-based drug discovery and development.

Figure 2. General process for structural simplification of NPs.

Due to its importance in NP-based drug discovery, we simplification of NPs. The focus will be on design strategies,
herein present a comprehensive review of the structural representative examples, and new technologies. A number of
C DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 3. Process of the structural simplification of morphine.

case studies of particularly successful examples leading to unknown during the early stage of structural simplification.
marketed drugs or drug candidates will be discussed in detail to The simplification strategies mainly include step by step
illustrate how structural simplification is applied in lead dissection of the complex structure, the elimination of
optimization of NPs. redundant chiral centers, the reduction of the number of
rings, and scaffold hopping.32−34 During the simplification
2. STRATEGIES FOR THE STRUCTURAL process, structural units should be systematically removed to
SIMPLIFICATION OF NPS determine the relative importance of the different components.
“Molecular obesity” is considered an important factor in the On the basis of the SAR and pharmacophores, the essential
high attrition rates of drug candidates in pharmaceutical substructures are retained in the simplified analogues.
industries.25−27 According to Polanski’s analysis, less complex Recently, important progress has been made in identifying
drugs were more likely to achieve better market success.28 To the targets of NPs.35,36 If the target of the NP is known, the
reduce molecular obesity, structural simplification by the binding mode should be clarified by structural biology,
appropriate removal of nonessential groups represents a chemical biology, or molecular modeling studies. Then, the
practical and powerful strategy in the process of lead key structural motifs responsible for the ligand−target
optimization.25−27 Structural simplification is defined as the interactions can be obtained, which will guide the rational
reduction of molecular complexity of a target compound. design of simplified derivatives via the removal of unnecessary
Thus, the first step of structural simplification design is to fragments.
analyze the molecular complexity of NPs (Figure 2). Although Several issues should be considered during the structural
there is not a global definition for molecular complexity, it is simplification process. First, the biological activity is not the
generally associated with MW, the number and connectivity only criteria in simplification. The main purpose of structural
(linked, fused or bridged) of rings, the number and simplification is to overcome the drawbacks of the lead
configuration of chiral centers, and so on.29 Molecular compound and improve the drug-likeness. In some cases, the
complexity is a crucial concept in drug discovery because it activity of the simplified compounds may be lower than that of
is associated with target selectivity, ADMET, and safety the lead, but it is more important to achieve a balance between
profiles.30,31 Recently, progress has been made in the the pharmacological potency and the ADMET properties. In
quantification of the molecular complexity and complexity− addition to the activity, drug-likeness filters, such as RO5 and
property relationships.31 ligand efficiency (LE), should be considered. However, it is
The strategy for designing simplified analogues of NPs also should be noted that several classes of drugs, such as
depends on whether the binding targets are known. In most antibiotics and antitumor agents, tend to violate RO5
cases, the molecular target or binding mode of the NP is guidelines.37−40 Second, as the size of a molecule decreases,
D DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 4. Structural simplification of schisandrin C.

Figure 5. Structural simplification of ISP-1.

Figure 6. Structural simplification of trichostatin A (A) and the binding modes of trichostatin A (B, PDB code: 1C3R) and vorinostat (C, PDB
code: 1C3S) with HDAC.

the simplified derivatives may act by a new mode of action.41 3. STRUCTURAL SIMPLIFICATION OF NPs: CASE
NPs gain some of their high binding affinity and exquisite STUDIES
selectivity from their structural complexity and overly simple 3.1. Structural Simplification Leading to Clinical Drugs
analogues may bind well to other targets. Therefore, the The effectiveness of structural simplification in NP-based drug
structural simplification must be moderate. Overly simplified development has been validated by several clinical drugs.
molecules tend to show reduced biological activity and target Structural simplification of morphine has successfully gen-
selectivity. Third, notably, not all lead compounds can be erated a series of semisynthetic and synthetic analgesics that
simplified (e.g., digitalis).42 NPs are often metabolites that are widely used in the clinic. Bicyclol, fingolimod, vorinostat,
and eribulin also represent successful examples of structural
evolved for their unique structural features, and form specific simplification of NPs. From the five case studies discussed in
interactions with the targets. As a result, structural the following sections, the initial impetus for structural
simplification of NPs does not always work. simplification is to reduce the synthetic difficulty. The most
E DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 7. Structural simplification of complex NP halichondrin B and the binding mode of eribulin with tubulin (PDB code: 5JH7).

Figure 8. Structural simplification of staurosporine.

Table 1. Inhibitory Activity of Simplified Analogues of Staurosporine towards PKC Isoforms


PKC subtypes (IC50, μM)
α βI βII γ δ ε ζ η
enzastaurin 0.8 0.03 0.03 2 1 0.3 8 0.4
ruboxistaurin 0.36 0.0047 0.0059 0.3 0.25 0.6 >100 0.052
staurosporine 0.045 0.023 0.019 0.11 0.028 0.018 >1.5 0.005

frequently used simplification strategies include reducing the activity, the simplified derivatives generally showed comparable
ring number and removing chiral centers. For the biological activity or even slightly lower activity at the molecular and
F DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 9. Structural simplification of asperlicin.

Table 2. In Vitro Antitumor Activity of Trabectedin and and eribulin), MW is only one of the factors in reducing the
Phthalascidin (IC50, nM) structural complexity. The MW of several morphine derivatives
(e.g., compounds 3−6, 9, and 11) and bicyclol is even slightly
higher than their parent molecules because a reoptimization
process is often required to achieve favorable potency,
selectivity and other drug-like properties. Even though, better
synthetic accessibility can be observed in all the five cases.
3.1.1. Structural Simplification of Morphine: A Classic
Story in Medicinal Chemistry. The development of
simplified morphine derivatives as analgesics is a classic
example of the structural simplification of a NP (Figure 3).43
The chemical structure of morphine (1)44 features a complex
five-ring system. It is mainly a μ-opioid receptor agonist (Ki =
compounds A549 HCT116 A375 PC3 1.8 nM).45 The elimination of the bridging furan ring (E ring)
trabectedin 0.95 0.38 0.17 0.55
of morphine afforded morphinans (Figure 3, compounds 2 and
phthalascidin 1.0 0.50 0.15 0.70
3), which have a similar stereo configuration to that of
morphine. For example, levorphanol (2)46 was obtained after
several chemical transformations of morphine, including the
cellular level compared with the original NP. The reduced elimination of the E ring, reductive hydrogenation of the C
structural complexity is helpful to improve the synthetic ring, and removal of the hydroxyl group. Levorphanol showed
accessibility, and PK/PD profiles and contributes to the clinical an excellent analgesic effect (4-fold more potent than
development. Although MW is always significantly decreased morphine). Mechanism study revealed that levorphanol was
during the simplification process (e.g., fingolimod, vorinostat, 6−8 times as potent as morphine at the μ-opioid receptor (Ki

Table 3. Structural Simplification and Binding Affinity of Simplified Erythrinane Analogues towards nAChRs

binding affinity toward nAChRs (Ki, μM)


compounds R R’ α4β2 α4β4 α3β4 α7/5-HT3A
38 CH3 CH3 ∼100 ∼100 ∼100 ∼30
39 CH3 CH3 5.5 ∼100 ∼300 >500
40 CH3 CH3 0.87 ∼300 ∼300 >500

G DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 4. Structural Simplification of Antifungal NP Sampangine and In Vitro Antifungal Activity (MIC80, μg/mL)a

compounds C. alb. C. par. C. neo. C. gla. A. f um. T. rub. M. gyp.


42 4 8 0.25 4 1 2 2
43 0.5 0.5 0.25 0.25 2 1 0.125
44 0.5 1 0.5 2 4 1 1
sampangine 0.5 4 2 ≤0.125 16 16 >64
a
Abbreviations: A. fum., Aspergillus f umigatus; C. neo., Cryptococcus neoformans; C. alb., Candida albicans; C. par., Candida parapsilosis; C. gla.,
Candida glabrata; T. rub., Trichophyton rubrum; M. gyp., Microsporum gypseum.

Table 5. Structural Simplification and Antitumor Activity of GA and Its Simplified Analogues (IC50, μM)a

compounds MCF-7 BGC-823 SMMC-7721 HepG2 MDA-MB-231


46 9.16 3.59 8.06 2.37 0.32
47 68.3 81.1 147 95
gambogic acid 2.19 2.35 3.20 2.40 1.5
a
Abbreviations: MCF-7, mammary carcinoma; BGC-823, gastric carcinoma; SMMC-7721, hepatocellular carcinoma; HepG2, liver carcinoma;
MDA-MB-231, breast cancer.

Table 6. In Vitro Antibacterial Activities of Simplified agonist (K i = 2.5 nM) and μ receptor competitive
Analogues of Sanguinarinea antagonist,49,50 and is widely used for the treatment of
moderate to severe pain, such as postoperative pain, trauma,
cancer pain, and renal or biliary colic.
The elimination of both the C and E rings of morphine led
to benzomorphans (4−6, Figure 3), whose conformations
were similar to that of morphine. The introduction of
hydrophobic substituents on the nitrogen atom could adjust
the analgesic and addictive effects. The analgesic effect of
phenazocine (4)51 was 10-fold more potent than that of
S.
S.
aureus S. S. B. B.
morphine. Although pentazocine (5)52 was 3-fold less active
compounds aureus PR epidermidis pyogenes subtilis pumilus than morphine, it displayed lower addictive side effects and
49 0.06 0.25 0.125 1 0.06 0.06 was approved as the first nonaddictive opioid analgesic. Similar
50 0.06 0.25 0.125 1 0.125 0.125 to butorphanol, pentazocine is a κ-opioid receptor agonist and
51 0.06 1 0.25 2 0.125 0.125 μ receptor antagonist.50
sanguinarine 8 4 8 4 4 4 Although pethidine (7),53 also known as meperidine, was
a
Abbreviations: S. aureus, Staphylococcus aureus ATCC25923 not discovered by the direct structural simplification of
(penicillin-susceptible strain); S. aureus PR, Staphylococcus aureus morphine, it could be seen as a simplified analogue of
PR (penicillin-resistant strain); S. epidermidis, Staphylococcus morphine by the elimination of the B, C, and E rings.
epidermidis; S. pyogenes, Staphylococcus pyogenes; B. subtilis, Bacillus Compared with morphine, pethidine displayed a better
subtilis; B. pumilus, Bacillus pumilus. analgesic effect and oral potency. Further structural mod-
ifications led to the development of a series of synthetic
= 0.21 nM), and was also an agonist of the δ-opioid receptor analgesics such as α-prodine (8)54 and fentanyl (9).55 Like
(Ki = 4.2 nM) and κ-opioid receptor (Ki = 2.3 nM).46 The morphine, fentanyl exerts its analgesic effects by acting as an
addition of a hydroxyl group on the BC ring and replacement agonist of μ-opioid receptor. Synthetic aminoketone analgesic
of the N-methyl group with a cyclobutylmethyl group afforded methadone (11)56 was derived from the optimization of the
butorphanol (3),47,48 which was 10-fold more potent than derivatives of fluorene-9-carboxylate analgesic 10.57 Meth-
morphine, and more importantly, the addiction side effect was adone can be regarded as the ring-opened analogue of
greatly reduced. Butorphanol acts as a κ-opioid receptor piperidine analgesics without the B, C, D, and E rings of
H DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 10. Structural simplification of dynemicin and the mechanism of dynemicin DNA damage.

aromatic ring, a basic tertiary amine, and a piperidine or


piperidine-mimic group.
3.1.2. Structural Simplification of Schisandrin C:
Discovery of Bicyclol. Schisandrin C (12, MW = 384), a
NP isolated from the fruit of the traditional Chinese medicine
Fructus Schizandrae, displays antihepatitis B virus (HBV)
activity and significantly decreases transaminase levels.58
Initially, the chemical structure of schisandrin C was
incorrectly assigned (13, Figure 4). Furthermore, a total
synthesis of the wrong structure was performed. Interestingly,
the intermediate bifendate (14) was found to possess potent
activity in the decrease of aminotransferase levels.59 Compared
with schisandrin C, bifendate lacks the seven-membered ring,
and thus its synthetic accessibility is greatly improved.
Figure 11. Structural simplification of neurotrophic sesquiterpenes. Importantly, bifendate retained the potent pharmacological
activity of its parent compound. Bifendate has been widely
used in the clinical treatment of patients with elevated
aminotransferase levels, which is caused by viral hepatitis or
drug-induced liver injury. However, its water solubility is poor
due to its high melting point (180 °C),60 which is possibly a
result of its symmetric structure and high lattice energy. To
reduce the molecular symmetry, one carboxylic ester group was
reduced to a hydroxymethyl group, which afforded bicyclol
(15, MW = 390).61 As a result, the melting point of bicyclol
was decreased to 138 °C.62 Compared with bifendate, bicyclol
had improved solubility, greater in vivo absorption, better
bioavailability, and biological activity.63 Pharmacologically,
Figure 12. Structural simplification of pyripyropene A. bicyclol displayed obvious antifibrotic and hepatoprotective
effects against liver injury induced by CCl4 or other
hepatotoxins in mice and rats.64 It also exhibited antihepatitis
morphine. Although methadone is a simple and highly flexible virus activity in duck viral hepatitis.64 Bicyclol has been
molecule, its stereo configuration is similar to that of pethidine approved for treating patients with elevated aminotransferase
(Figure 3). Similar to morphine, methadone is a potent μ levels caused by chronic viral hepatitis in China since 2004.58,65
receptor agonist.50 Moreover, methadone also acts as a N- 3.1.3. Structural Simplification of Myriocin: Discovery
methyl-D-aspartate (NMDA) antagonist and serotonin reup- of Fingolimod. Myriocin (16, MW = 401), also known as
take inhibitor, which could enhance its analgesic properties.50 antibiotic ISP-1, is a metabolite of the fungus Isaria sinclairii,
Currently, methadone is widely used for the treatment of and it has immunosuppressive activity.66 Although myriocin
morphine addiction because of its relatively low risk of showed better in vitro and in vivo immunosuppressive activity
addiction. than ciclosporin A, it was 100-fold more toxic and poorly
The structural simplification of morphine highlights the soluble. Myriocin is a structural analogue of endogenous
importance of retaining the proper conformation and the key sphingosine-1-phosphate (S1P) and acts as an agonist of S1P
pharmacophore. Although simplified morphine derivatives receptors. It has a complex chemical structure that includes
have diverse chemical structures, they share similar con- three chiral centers, one trans-double bond, and five polar
formations and common pharmacophores including an groups (three hydroxyl groups, one carboxyl, and one amino
I DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 13. Structural simplification of pladienolide B.

the basis of a systematic SAR study, derivatives with m and n


values (in the side chain) of 2 and 8, respectively, displayed the
most potent activity. Among these derivatives, fingolimod (20,
MW = 307) was successfully developed in 2010 for treating
patients with multiple sclerosis.69
3.1.4. Structural Simplification of Trichostatin A:
Discovery of Vorinostat. Trichostatin A (TSA, 21, MW =
302), isolated from the actinomycete Streptomyces hygroscopi-
cus, is a potent noncompetitive histone deacetylase (HDAC)
inhibitor (Ki = 3.4 nM).70 It reversibly inhibited HDAC at
nanomolar concentrations by chelating the zinc in the active
site of HDAC with its hydroxamic acid group. SAR analysis
revealed that two conjugated trans double bonds and the chiral
center are not essential for the activity. Breslow et al.
performed a structural simplification study of TSA by
eliminating the conjugated double bonds and chiral center
and retaining the hydroxamic acid group as the key
Figure 14. Structural simplification of melicobisquinolinone B based pharmacophore (Figure 6A).71 The simplified analogue,
on MCRs.
vorinostat (22, MW = 264), can be easily synthesized, and it
retained excellent HDAC inhibitory activity (HDAC1, IC50 =
group). A SAR study indicated that the C-14 ketone group, C- 0.04 μM).72 Crystal structures revealed that TSA and
6 double bond, and C-4 hydroxyl group of myriocin were not vorinostat adopted a similar binding mode in the active site
essential for the activity, and the stereo configuration of the C- of HDAC (Figure 6B,C).73 Vorinostat was approved in 2006
3 hydroxyl group had little effect on the activity.66,67 On the for the treatment of advanced primary cutaneous T-cell
basis of the SAR, Fujita et al. designed and synthesized lymphoma.74
simplified analogues (18) with a symmetric 2-alkyl-2-amino- 3.1.5. Structural Simplification of Halichondrin B:
propane-1,3-diol side chain to mimic the structure of the Discovery of Eribulin. Halichondrin B (23, Figure 7) is a
terminal group of sphingosine (17, Figure 5).68 Derivatives large polyether macrolide (MW = 1110) that was isolated from
containing a side chain with 13−15 carbon atoms displayed the marine sponge Halichondria okadai Kadota.75 The marine
higher activities than that of ciclosporin A. However, the macrolide showed excellent activity against B16 melanoma
molecular conformation became more complicated, which was cells (IC50 = 0.093 ng/mL) and potent in vivo activity in
caused by the flexibility of the alkyl side chain. To solve this inhibiting tumor growth in mice.76 Mechanism studies revealed
problem, Fujita et al. replaced part of the side chain with a that it acted as a tubulin-destabilizing agent and inhibited
phenyl group to restrict the molecular conformations.66 tubulin polymerization and microtubule assembly.77 Despite its
Therefore, derivatives 19 were designed and synthesized. On promising biological activity, the limited supply of the NP from
J DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 15. Structural simplification of campotothecin based on MCRs.

Table 7. In Vitro Antitumor Activity of Campotothecin and Simplified Analogue 81 (IC50, μM)a
compounds Hela Jurkat MCF-7 A549 LoVo U373 SKMEL PC3
81 0.17 0.05 0.45 0.61 0.29 0.10 0.23 0.19
campotothecin 1.3 0.09 6.0 0.20 0.008 0.033 0.047 0.48
a
Abbreviations: Hela, cervical cancer; Jurkat, human T-cell leukemia; LoVo, colon cancer; U373, human glioblastoma; SKMEL, melanoma; PC3,
prostate cancer.

its marine source and its extremely complex structure are 3.2.1. Structural Simplification of Staurosporine:
serious obstacles for its total synthesis and drug development. Discovery of Ruboxistaurin and Enzastaurin. Staurospor-
In 1992, Kishi et al. reported the first total synthesis of ine (25, Figure 8), an alkaloid originally isolated from cultures
halichondrin B, but the synthetic route needed nearly 120 of Streptomyces sp. AM-2282, shows a wide range of biological
steps.78 Despite the length, the total synthesis provided enough activities including antitumor and neurotrophic potency.84,85
material to support preclinical development. Interestingly, in Staurosporine was validated as an inhibitor of protein kinase C
an investigation of the bioactivities of the synthetic (PKC), which is a family of serine/threonine-specific kinases
intermediates and analogues, the C1−C38 macrolide was that is composed of at least 12 isozymes or subtypes.86
found to show antitumor activity similar to that of the parent Therapeutically, a selective PKC antagonist could treat a
compound.79 Further investigations revealed that the unstable variety of diseases including diabetes and cancer.87,88 However,
lactone moiety could be replaced by a non-hydrolyzable ester staurosporine had limited selectivity for both ATP-dependent
bioisostere, and the keto analogues displayed the best activity. kinases and individual PKC isozymes. The complex structure
As a result, a simplified analogue, named eribulin (24), was of staurosporine includes eight rings and four chiral centers,
developed.80,81 Compared with halichondrin B, the structure of making its total synthesis challenging. Lilly’s research group
eribulin is greatly simplified (MW = 729). Additionally, the reported the structural simplification of staurosporine and
synthetic route was significantly shortened. In the crystal identified a novel class of macrocyclic bisindolylmaleimides
structure of tubulin in complex with eribulin, eribulin forms (represented by ruboxistaurin, 26).89 The simplification
directly or water-mediated hydrogen bonds with residues process included opening the central benzene and tetrahy-
Tyr224, Val177, and Asp179 within the binding pocket (Figure dropyrane rings and oxidizing the pyrrolidone into maleimide.
7).82 Eribulin displayed superior efficacy over other anti- After the ring opening, ruboxistaurin contains only one chiral
mitotics such as paclitaxel,81 and its mesylate derivative was center and displayed good selectivity for PKC β. For example,
approved by the FDA in 2010 for the treatment of patients compared with staurosporine, 26 exhibited better activity
with refractory metastatic breast cancer.83 against PKC βI and βII with IC50 values of 0.0047 μM and
0.0059 μM, respectively. Moreover, its selectivity for inhibiting
3.2. Structural Simplification Leading to Clinical
PKC α is significantly increased (Table 1). Ruboxistaurin
Candidates
mesylate was evaluated in the clinic for the treatment of
In contrast to successful clinical drugs, candidate drugs derived diabetic retinopathy, but the phase III clinical trial of
from structural simplification of NPs are relatively limited. ruboxistaurin was terminated because of limited efficacy. In
Herein only two examples of clinical candidates are introduced, 2016, a phase I and II study for ruboxistaurin was conducted
in which reducing the number of rings and chiral centers is also for treating adult patients with heart failure.90 Furthermore,
used as the simplification strategy. The first example Faul et al. identified a series of acyclic (N-azacycloalkyl)
(simplification of staurosporine) illustrates structural simplifi- bisindolylmaleimides (represented by enzastaurin, 27).91
cation guided by different biological activity, while the second Enzastaurin displayed nanomolar activity against PKC βI and
example (simplification of asperlicin) discusses pharmacophore βII with an IC50 value of 0.03 μM. Meanwhile, it had good
inspired structural simplification. selectivity for PKC α. Enzastaurin was advanced into clinical
K DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 16. Structural simplification of yohimbine by the SCONP scaffold tree.

trials; however, its phase III clinical trials failed due to the poor gastrointestinal system.96,97 However, several limitations, such
efficacy for treating glioblastoma, breast and lung cancer.92 as its complicated structure, difficulties in total synthesis,
Now, clinical trials of enzastaurin are focused on treating limited SAR information, and poor oral bioavailability, have
tumor patients with specific biomarkers.93 hampered the direct clinical application of asperlicin.98,99
In addition to its PKC inhibitory activity, staurosporine was Through a systematic structural analysis of asperlicin, Evans et
also confirmed to be an effective Janus kinase 3 (JAK3) al. found that the 1,4-benzodiazepine (BZD, 31) and L-
inhibitor with an IC50 value of 6 nM.84,94,95 The simplification tryptophan (L-Trp, 32) groups were the essential structural
of staurosporine led to a series of new simplified analogues motifs (Figure 9).100 Moreover, the 5-phenyl-1,4-benzodiaze-
with highly potent JAK3 inhibitory activity.95 Initially, Yang et pine scaffold is a privileged structure for the inhibition of G-
al. replaced the amino-sugar group with an alkyl ring in which protein coupled receptors (GPCRs),101−103 which might be
an attached hydroxymethyl group mimicked the methylamino the pharmacophore that binds with CCK. Therefore, the 5-
fragment of the amino-sugar. SAR studies revealed that a phenyl-1,4-benzodiazepine scaffold was used as the lead
suitable substituent at the C3 and/or C9 position could structure for further modification. The right-hand side of
increase the JAK3 inhibitory activity. Therefore, further asperlicin was an analogue of L-Trp, which is the key residue
optimization focused on regioselective functionalization of for the interaction of cholecystokinin with CCK. Therefore, a
the indolocarbazole unit. As a result, simplified analogues 28 series of simplified derivatives (33) were designed by
(IC50 = 3 nM) and 29 (IC50 = 3 nM) were synthesized and combining the BZD and L-Trp groups. After a systematic
demonstrated excellent inhibitory activity against JAK3.95 SAR study, devazepide (MK-329, 34) was discovered,104,105
3.2.2. Structural Simplification of Asperlicin: Discov- and it displayed potent antagonistic activity against CCK with
ery of Devazepide. Asperlicin (30), a mycotoxin isolated an IC50 value of 0.8 nM. Moreover, 34 showed good selectivity
from Aspergillus alliaceus,96 is a selective antagonist (IC50 = 1.4 for the benzodiazepine receptor.100 Although a pilot clinical
μM) of pancreatic cholecystokinin receptors (CCK) with trial of devazepide in patients with advanced pancreatic cancer
potential therapeutic effects in CCK-related disorders of the failed to demonstrate any impact on tumor progression,106 this
L DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 17. BIOS libraries inspired by withanolide A and bioactive simplified analogues.

Figure 18. Structural simplification of podophyllotoxin (A) and the binding modes of podophyllotoxin (B, PDB: 1SA1), analogues 97 (C) and 98
(D) with tubulin. Molecular docking was used to reproduce the binding modes of analogues 97 and 98.

compound is widely used as a tool compound for the analysis of the case studies described in the following sections,
investigation of CCK receptors. reducing the number of rings and chiral centers is also the
3.3. Structural Simplification Leading to Preclinical most commonly used simplification strategy, after which
Candidates or Lead Compounds synthetic difficulty can be significantly reduced. Notably,
Currently, most examples in structural simplification of NPs most of the simplified compounds were only evaluated at the
are still at the stage of identifying new lead compounds. After molecular and/or cellular level, and only two examples (i.e.,
M DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 19. Structural simplification of taxuspine X.

Figure 20. Structural simplification of cortistatin A.

simplified erythrinane and largazole derivatives) reported the from the extracts of the marine organism Ecteinascidia
in vivo potency. The lack of in vivo PK/PD evaluations limits turbinate.114 Trabectedin is a highly potent antitumor agent
further development of these simplified NP derivatives. and has been approved as the first marine-derived antitumor
Moreover, the in vivo evaluations are also helpful to identify drug for the treatment of soft tissue sarcoma and ovarian
the problems to be overcome and guide further structural cancer. Mechanism study revealed trabectedin could bind to
optimizations. Target identification is a bottleneck in NP-based the minor groove of DNA, thus bending the DNA helix toward
drug discovery.10,107 In the following structural simplification the major groove.115 Moreover, it interfered with cellular
examples, although targets of a large portion of NPs have been transcription-coupled nucleotide excision repair to induce cell
validated, the lack of binding mode information limited the death and cytotoxicity.115 With an aim to discover new
efficiency to design simplified analogues. Thus, structural trabectedin analogues with simpler chemical structures, better
biology studies are highly required to clarify the key structural synthetic accessibility, and higher stability, Corey et al.
motifs for binding NPs with their targets and guide the rational performed a structural simplification of trabectedin in which
simplification design. Also, the change of molecular targets has
the third isoquinoline ring was substituted with a phthalimide
been observed in some cases (e.g., simplified bryostatin and
moiety (Table 2).116 Compared with trabectedin, structurally
campotothecin derivatives). Therefore, validation of the
mechanisms is necessary to confirm whether the simplified more simpler phthalascidin (36) displayed comparable or
compounds act on the same target as the parent NP. Recently, higher antitumor activity against A549 (lung cancer), HCT116
highly efficient and practical synthetic methods, such as (colon cancer), and A375 (malignant melanoma) and PC
multicomponent reactions (MCRs),108−110 FOS,111 DTS,112 (prostate cancer) cell lines. Moreover, the antiproliferative
and BIOS,113 have been applied to synthesize simplified NP activity of phthalascidin (IC50 range: 0.1−1 nM) is better than
analogues. Although successful examples are still limited, these those of several well-known natural antitumor agents (e.g.,
new methods will contribute to improve the efficiency of Taxol, camptothecin, adriamycin, and etoposide) by 1−3
structural simplification of NPs. orders of magnitude. Moreover, compound 36 could be easily
3.3.1. Structural Simplification by Reducing the Ring synthesized from known synthetic intermediates (in six
Number. 3.3.1.1. Structural Simplification of Trabecte- chemical steps) and displayed better stability than trabectedin
din. Trabectedin (35, Table 2), also known as ecteinascidin- which contains the unstable spiro 10-member lactonic ring. A
743, is a bis-tetrahydroisoquinoline-fused alkaloid isolated mechanism study revealed that phthalascidin shared a similar
N DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 21. Structural simplification of bryostatin 1.

Figure 22. Structural simplification of cyclopamine 1.

Figure 23. Diverted total synthesis and simplification of carolacton-inspired analogues.

O DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 24. Diverted total synthesis of promysalin-inspired analogues.

Figure 25. Structural simplification of chaetocin.

mode of action to trabectedin and acted by inducing DNA− 3.3.1.3. Structural Simplification of Sampangine.
protein cross-linking.116 Invasive fungal infections (IFIs) are associated with high
3.3.1.2. Structural Simplification of Erythrinane. mortality in immune-compromised hosts. Despite the high
Erythrinane (37, Table 3), a member of the Erythrina alkaloid mortality of IFIs, effective and safe antifungal agents are very
family, is characterized by a unique tetracyclic spiroamine rare. More importantly, increasing resistance is being
framework, and it was found to be a potent antagonist of developed to almost all the clinically available antifungal
nicotinic acetylcholine receptors (nAChRs).117 However, the agents. Sampangine (41, Table 4), an azaoxoaporphine
limited availability of erythrinane and its analogues from alkaloid extracted from the stem bark of Canangaodorat,
natural sources hampers their broad application. In the pursuit displayed broad-spectrum in vitro antifungal activity.119
of simpler and more synthetically accessible analogues, Crestey Further lead optimization and drug development were
et al. performed a structural simplification of erythrinane in hampered because the aromatic tetracyclic derivatives had
which the tetracyclic scaffold was simplified step by step to poor water solubility (kinetic solubility 12.6 μg/mL) and were
inactive in vivo.120 To address these bottlenecks, our group
identify the groups essential to the nAChRs inhibitory activity
performed a series of scaffold hopping and structural
(Table 3).118 First, the A-ring methoxyl group and the two
simplification studies, and identified a thiophene analogue,
double bonds in the A and B rings were sequentially deleted, ZG-20 (42), with improved activity and water solubility
resulting in tetracyclic derivatives 38, which showed retained (kinetic solubility 48 μg/mL, Table 4).121 After rings A and B
affinity for a variety of nAChR subtypes. When the A ring was were removed, simplified tricyclic analogue ZG-20-07 (43) and
removed, tricyclic derivative 39 showed improved affinity for bicyclic analogue ZG-20−41 (44) displayed improved
α4β2 (Ki = 5.5 μM). Notably, the selectivity of 39 for other antifungal activities (Table 4) and water solubility. Moreover,
subtypes was also significantly increased. Further deletion of they showed good in vivo antifungal potencies with low
the B ring gave bicyclic tetrahydroisoquinoline analogue 40, toxicities in a C. elegans−C. albicans infection model.122,123
which displayed a submicromolar binding affinity toward α4β2 Compounds 43 and 44, which were developed via a structural
(Ki = 0.87 μM) with more than 300-fold selectivity over the simplification, had several advantages over fluconazole, such as
other subtypes. In a mouse forced swim test, compound 38 superior fungicidal activity, higher potency against fluconazole-
showed an in vivo antidepressant-like effect at a dose of 30 resistant strains, and better inhibition of biofilm formation and
mg/kg. yeast-to-hypha morphological transitions.122,123
P DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 26. Structural simplification of largazole (A) and the binding mode of largazole thiol (B, PDB code: 3RQD) with HDAC8.

Figure 27. Structural simplification of lobatamide C.

3.3.1.4. Structural Simplification of Gambogic Acid. with low toxicity to normal cells.124 It prevented cancer
Gambogic acid (GA, 45), the major component of gamboges metastasis and angiogenesis, and was evaluated in phase II
resin, displayed potent in vitro and in vivo antitumor activity clinical trials in China.125,126 However, further clinical trial of
Q DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 28. Structural simplification of sanglifehrin A (A) and the binding modes of sanglifehrin A (B, PDB code: 1YND), analogues 142 (C, PDB
code: 1NMK), 143 (D, PDB code: 5T9U), 146 (E, PDB code: 5T9Z), 147 (F, PDB code: 5TA2), and 148 (G, PDB code: 5TA4) with cyclophilin
A.

GA was terminated because of limited antitumor efficacy. The by step simplification of GA to obtain a series of analogues
structure of GA can be divided into two fragments: a planar with a simplified planar region (Table 5).125 Compound 46, a
region that contains rings A, B, and C and the caged core D simplified analogue obtained by removal of the A ring,
ring (4-oxa-tricyclo[4.3.1.03,7]dec-2-one) system. A prelimi- displayed in vitro antitumor activity comparable or superior
nary SAR study revealed that the 4-oxa-tricyclo[4.3.1.03,7]dec- to that of GA.125 Mechanistic studies indicated that compound
2-one ring system was essential for the antitumor activ- 46 was able to induce programmed cell death and arrest cell
ity.125,127 To get more SAR information and improve the cycle growth in the G2/M phase as well as regulate apoptotic
antitumor potency, You’s group performed a systematic step related proteins and cellular caspase-3 activity. The mode of
R DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 8. Biological Activities of Simplified Analogues of


Sanglifehrin A
PPIase
TR-FRET Cyp A functional assay HCV genotype 1b
compounds binding Kd (nM) Ki (nM) replicon EC50 (nM)
143 25 16 600
144 65 65 240
145 2600 795 4500
146 64 14 320
147 11 4 36
148 24 7 87
149 48 4.3 620
cyclophilin A 17 6.7 170

Figure 30. Structural simplification of caprazamycin B.

Figure 31. Structural simplification of militarinone D.

bacteria strains than sanguinarine (Table 6). Mechanism


Figure 29. Structural simplification of anguinomycin. studies revealed that these analogues exhibited strong cell
division inhibitory activity and obvious inhibition on B. subtilis
FtsZ polymerization, indicating that they inhibited the
action of compound 46 was consistent with the apoptotic bacterial proliferation by interfering the function of bacterial
induction effects of GA. In contrast, compound 47, which was FtsZ.
obtained both the A and B rings were removed, was 3.3.1.6. Structural Simplification of Dynemicin. The
substantially less effective. violet dynemicin (52, Figure 10), a cyclic 10-membered
3.3.1.5. Structural Simplification of Sanguinarine. enediyne NP isolated from the fermentation broth of
Sanguinarine (48, Table 6) is a benzo[c]phenanthridine Micromonospora chersina. M956-1,131,132 exhibited strong in
alkaloid which displayed moderate antibacterial activity against vitro activity against Gram-positive bacteria and significant in
a broad range of Gram-positive bacteria.128 It was validated as vivo potency against Staphylococcus aureus Smith infection in
a filamenting temperature-sensitive protein Z (FtsZ, an mice (i.p., PD50 = 0.13 mg/kg) with low toxicity. In addition,
antibacterial target)129 inhibitor that can alter the Z-ring dynemicin demonstrated highly potent antitumor activity
formation and function of FtsZ.130 To further improve the against a variety of cancer cell lines and significantly prolonged
antibacterial activity, Ma et al. simplified the skeleton of the life span of mice with P388 leukemia or B16
sanguinarine to synthesize a series of 5-methyl-2-phenyl- melanoma.131,133 Structurally, dynemicin contains an anthra-
phenanthridium derivatives (Table 6).128 Compounds 49, 50, quinone and a 10-membered ring with a 1,5-diyn-3-ene bridge.
and 51 displayed better activity against sensitive and resistant The mechanism of its antitumor activity involves intercalation
S DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 32. Structural simplification of bistramide A (A) and the binding mode of bistramide A (B, PDB code: 2FXU) with monomeric actin.

rings of dynemicin and incorporated a phenylsulfone ethylene


carbamate moiety on the nitrogen atom. The resulting
simplified analogue, 54, was highly active with an IC50 value
of 20 fM against Molt-4 T-cell leukemia cells.134
3.3.1.7. Structural Simplification of Merrilactone A.
Sesquiterpenoid NPs, such as merrilactone A (55, Figure
11)137 and anislactone A/B (56),138 possess a variety of
biological activities.139 Structurally, merrilactone A is a
sesquiterpene dilactone consisting of two γ-lactones and an
oxetane ring, and it displayed intriguing neurotrophic activity
in primary cultures of fetal rat cortical neurons.137 Although
Figure 33. Structural simplification of withaferin A by FOS. nonpeptidic small molecules with neurotrophic activity are
considered to have potential for the development of
therapeutic agents against neurodegenerative diseases such as
Alzheimer’s and Parkinson’s diseases, the drug development of
merrilactone A was hampered because of its structural
complexity as well as the poor yields and high costs of
obtaining it from its natural source.139 Thus, the structural
simplification of merrilactone A was performed by Tiefen-
bacher’s group to reduce its complexity while maintaining its
neurotrophic activity. By simplifying the scaffold of merrilac-
tone A, analogues 57 and 58 showed neurite outgrowth activity
comparable to that of the reference drug jiadifenolide.140
Figure 34. Structural simplification of rakicidin A. Compared with the 17−26 steps required for the total
synthesis of merrilactone A, compounds 57 and 58 can be
of the anthraquinone core into DNA and triggering of a synthesized in just 6−8 steps with high yields from commercial
Bergman reaction, generating highly reactive benzenoid materials.141,142
diradicals, causing DNA damage (Figure 10).134,135 Guided 3.3.1.8. Structural Simplification of Pyripyropene A.
by the mechanism, structural simplification was focused on Pyripyropene A (59, Figure 12), isolated from the culture
mimics of dihydroquinoline epoxides spanned by an enediyne broth of Aspergillus f umigatus strain FO-1289, is a strong and
bridge (red part in compound 52) while retaining the selective inhibitor of sterol O-acyltransferase 2 (SOAT2) (IC50
antitumor activity and reducing the molecular complexity. = 0.07 μM),143−145 which is an important target for the
The simplest analogue, 53, was synthesized by Wender et al. in treatment of hypercholesterolemia and atherosclerosis.146,147
only seven steps and maintained the antitumor activity of the In the pursuit of new cholesterol-lowering or antiatheroscler-
parent compound.135,136 Nicolaou et al. removed the A and B otic agents, Nagamitsu’s group reported the first total synthesis
T DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 35. Structural simplification of spongistatins.

of pyripyropene A in 2011. 148 Then, the structural IC50 = 1.5 μM).23,153 Furthermore, Burkart et al. reported
simplification was performed to give new A ring-simplified C18−C19 cyclopropyl derivative 65, which displayed
pyripyropene A analogues based on the results of the SAR improved stability and effectively mimicked the apoptotic
studies. Among these simplified analogues, 60 showed and splicing activity of PB.154 It showed potent inhibitory
excellent SOAT2 inhibitory activity (IC50 = 0.07 μM) with activity against HCT116 cell line (IC50 = 42.9 nM), which was
high isozyme selectivity over SOAT1, and its activity was slightly less active than PB (IC50 = 24.8 nM). The results
comparable to that of natural pyripyropene A.147 60 was indicated that the C18−C19 epoxide moiety was not essential
considered to be the most potent and selective synthetic for the spliceosome activity.
inhibitor of SOAT2 and an attractive candidate for further Compound 66 lacking the 6-OH group demonstrated
development. potent inhibitory activity against the mutant SF3B1 pancreatic
3.3.1.9. Structural Simplification of Pladienolide B. cancer cell line (GI50 = 8.1 nM), high binding affinity to the
Pladienolide B (PB, 61, Figure 13) contains a 12-member SF3b complex (IC50 = 4.0 nM), and inhibition of pre-mRNA
macrolide ring with an extended epoxide-containing side splicing.155 Moreover, NP herboxidiene (67, Figure 13) could
chain.149 It acted by targeting the SF3B1 subunit of be seen as a simplified analogue of PB, whose 12-member
spliceosome and displayed potent antiproliferative and tumor macrolide ring was replaced by the 6-member tetrahydropyran
suppressive activity when assayed in both cell culture and ring.156 Herboxidiene displayed potent inhibitory activity
xenograft models.149,150 To investigate the importance of the against several human cancer cell lines (e.g., Hela cell line,
epoxide-containing side chain, Maier et al. reported a IC50 = 14.7 nM). However, analogue 68, a hybrid molecule
simplified analogue bearing a phenyl substituted side chain from herboxidiene and PB, was totally inactive (IC50 > 20
(compound 62).151 However, it was inactive up to 4 μg/mL in μM), which highlighted the central role of the methyl
a cellular proliferation assay against L929 mouse fibroblasts. substituent in the tetrahydropyran ring and the hydroxyl
The results highlighted the central role of the epoxide- group in the side chain.156
containing side chain for binding. After removal of the 3- 3.3.1.10. Structural Simplification of Melicobisquino-
hydroxy group and the methyl groups at positions 10, 16, 20, linone B. Melicobisquinolinone B (69, Figure 14) containing
and 22 in PB, simplified derivative 63 was inactive against the pyranoquinolone structural motif exhibited potent
several human cancer cell lines at the concentration up to 20 antitumor activities and was investigated as a potential
μM.152 In contrast, the aryl analogue 64 with a truncated side anticancer agent.157−159 The MCRs have the advantages of
chain displayed moderate antitumor activity (A549 cell line, environmental friendliness, atom economy, and efficient
U DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 36. Structural simplification of spongistatins.

Table 9. Antitumor Activity of Simplified Spongistatin acts by the inhibition of topoisomerase I (Top1).160 The total
Derivatives (IC50, nM) synthesis of campotothecin requires more than 15 steps due to
its five conjugated planar rings. The investigation of
human cancer cell line paclitaxel 166 174 175
camptothecin analogues identified two clinically useful agents,
MIP101 colon 200 0.1 0.08 587
topotecan161 and irinotecan,162 that can be employed for the
HCT116 colon 0.3 0.05 0.02 407
treatment of colon and ovarian cancer, respectively.163
1A9PTX22 47 0.03 0.007 >632
Magedov et al. developed a one-step multicomponent
1A9 1 0.03 0.007 >632
A549 6 0.07 0.04 >632
condensation of various aminoheterocycles (77), aldehydes
(76), and 1,3-indanedione (78) that facilitated the synthesis of
various heterofused indenopyridines (Figure 15).110 These
construction of molecules in only one or two synthetic steps. analogues (79) mimicked the A−D rings of campotothecin,
Magedov et al. developed a three-component reaction of
and the dihydropyridine moiety could be transformed to the
pyridine (70) with malononitrile (72) and various aromatic
planar indenopyridine skeleton (80) via an intracellular
aldehydes (71) to efficiently construct a library of compounds
with pyrano[3,2-c]pyridone and pyrano[3,2-c]quinolone scaf- oxidation. These simplified derivatives displayed broad-
folds (73, Figure 14).109 Most derivatives displayed sub- spectrum antitumor activity. For example, compound 81
micromolar or low micromolar inhibitory activities against showed better inhibitory activity toward HeLa, Jurkat, and
HeLa cells. Generally, pyrano[3,2-c]quinolones were more MCF-7 cells than campotothecin (Table 7).110 Interestingly,
potent than pyrano[3,2-c]pyridines.109 Compound 74 showed mechanism studies revealed that these analogues inhibited
potent inhibitory activity against HeLa cells with a GI50 value topoisomerase II (Top2) instead of Top1. These results
of 47 nM and significantly induced the apoptosis of Jurkat cells indicated that the structural simplification of NPs may result in
at a concentration of 5 μM. Mechanistic studies revealed that a change in the molecular target.
these compounds induced cell cycle arrest in the G2/M phase 3.3.1.12. Structural Simplification of Yohimbine.
and blocked in vitro tubulin polymerization. Yohimbine (82, Figure 16) is an indole alkaloid derived
3.3.1.11. Structural Simplification of Campotothecin. from the bark of the Pausinystalia yohimbe tree with diverse
Campotothecin (75, Figure 15) is a classic antitumor NP that biological activities.164 It was identified as an inhibitor of
V DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 37. Structural simplification of migrastatin (A) and the binding mode of analogue 185 with fascin (B, PDB code: 3O8K).

phosphatase Cdc25A by Waldmann’s group.41 The pentacyclic Hedgehog (Hh) signaling. Among these analogues, compound
scaffold of yohimbine was sequentially truncated into simpler 92 was proven to bind to the SMO protein (Ki = 57 nM) with
compounds by a chemistry-based rule165,166 called “structural low cytotoxicity. The same group further constructed a
classification of natural products” (SCONP).41 As a result, a compound library of substituted δ-lactones 93.173 Several
scaffold tree of yohimbine was obtained in which complex withanolides, particularly vinylogous urethane analogues,
scaffolds become smaller step by step (Figure 16A). The displayed significant potency in inhibiting the Wnt/β-catenin
simpler scaffolds in the scaffold tree may retain the bioactivity pathway. Compound 94 was the most potent withanolide
of the more complex compounds, which can be used for (IC50 = 0.11 μM) and interfered with the Wnt pathway by a
structural simplification. For example, the simplified tetracyclic novel tankyrase-independent mechanism. Interestingly, its
scaffold 85 showed similar phosphatase Cdc25A inhibitory simplified analogue, 95, retained the ability to inhibit the
activity (IC50 = 20.3 μM) to yohimbane (IC50 = 22.3 μM).167 Wnt reporter gene (IC50 = 4.5 μM), although it was 40-fold
In contrast, analogue 86 was inactive against Cdc25A and was less active. The novel and potent inhibitors of the Hedgehog
identified as a mycobacterial protein tyrosine phosphatase B signaling pathway and the Wnt/β-catenin pathway were
(MptpB) inhibitor.167 Moreover, analogue 87 showed potent promising lead compounds for drug discovery.
activity against the HeLa cervical cancer cell line (IC50 = 0.8 3.3.2. Structural Simplification by Removing Chiral
μM) and arrested the cell cycle at the G2/M phase.168 In Centers. 3.3.2.1. Structural Simplification of Podophyl-
addition to the inhibitory activity against MptpB,41 additional lotoxin. Podophyllotoxin (96, Figure 18), a well-known NP
simpler tetrahydrocarbazole (88) and indole (89) derivatives isolated from the roots of Podophyllotoxin peltatum, displays
were inhibitors of vascular endothelial growth factor receptors good in vitro antitumor activity.174 However, the total
2 (VEGFR-2).169 The results also indicated that over synthesis of podophyllotoxin is challenging due to its complex
simplification might lead to the change of the binding target. structure. The C ring of podophyllotoxin contains four chiral
3.3.1.13. Structural Simplification of Withanolide A. centers, and the facile epimerization of the C-2 position
Withanolide A (90, Figure 17) is steroid-like NP with diverse converts the compound into its inactive cis-acetone isomer in
bioactivities.170 The trans-hydrindane/dehydro-δ-lactone frag- vivo.175 The structural modification of podophyllotoxin was
ment was thought to be the essential pharmacophore. focused on its semisynthetic derivatives. Hitotsuyanagi et al.
Waldmann’s group developed the concept of BIOS, which designed and synthesized structurally simplified podophyllo-
integrates computational and synthetic tools to design and toxin analogues containing the 4-aza-2,3-dehydro-4-deoxypo-
synthesize bioactive simplified NP analogues.113,171 To reduce dophyllotoxin scaffold, which could be easily synthesized
the structural complexity of withanolide A, an efficient BIOS because of the elimination of the chiral centers.176 Interest-
method was developed for the synthesis of trans-hydrindane ingly, these derivatives retained the good antitumor activity of
dehydro-δ-lactone derivatives (91).172 Cell-based assays their parent compound.176 For example, compound 97
indicated that four withanolide analogues act by inhibiting displayed potent antitumor activity against the P-388 leukemia
W DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 10. Selected Examples of Structural Simplification of Complex NPs

X DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 10. continued

Y DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 10. continued

cell line with an IC50 value of 0.00182 μg/mL, which makes it mL). Using multicomponent synthesis, Magedov et al.
2-fold more potent than podophyllotoxin (IC50 = 0.0043 μg/ synthesized analogues bearing a dihydropyridine pyrazole
Z DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

scaffold.177 Compared with podophyllotoxin, the A ring and 3.3.2.4. Structural Simplification of Bryostatin 1.
the chiral centers were removed. A one-step multicomponent Chikungunya virus (CHIKV) is an arbovirus of the Alphavirus
reaction enabled a highly efficient construction of the library of genus that has been spreading rapidly.191 Currently, there is no
analogues and elucidation of the SARs. These simplified effective therapy to treat CHIKV infections, and the only
derivatives displayed good in vitro antitumor activities. For medications (i.e., the anti-inflammatory drugs) can be used to
example, compound 98 had an IC50 value of 0.02 μM and 0.01 alleviate the symptoms. PKC modulators were reported to
μM against the HeLa and MCF-7 cancer cell line, respectively, inhibit CHIKV virus-induced cell death, representing a
which were comparable to that of podophyllotoxin. Moreover, promising therapeutic choice to treat CHIKV infection.192−194
it showed an antiapoptotic effect comparable to that of Bryostatin 1 (106, Figure 21) is a marine NP bearing a
podophyllotoxin against the Jurkat leukemia cell line. polyacetate backbone which was originally isolated from the
Mechanism studies revealed that compound 98 inhibited bryozoan Bugula neritina.195 It is a great challenge for the total
tubulin polymerization and disrupted the formation of mitotic synthesis of bryostatin 1 because of the complex scaffold
spindles in dividing cells at concentrations as low as 5 nM, (nearly 55 synthetic steps).196 Bryostatin 1 was proven to be a
which was similar to podophyllotoxin. Molecular docking potent pan-conventional and novel PKC isoform modulator
results revealed analogues 97, 98, and podophyllotoxin (PKC δ binding affinity Ki = 1.1 nM). However, it was inactive
adopted a similar binding mode with tubulin (Figure in inhibiting CHIKV virus-induced cell death.191 Wender’s
18).177,178 group designed the simplified analogues of bryostatin 1 by
3.3.2.2. Structural Simplification of Taxuspine X. removing several chiral centers (Figure 21).191 Compared with
Paclitaxel (Taxol) and its derivative docetaxel are widely bryostatin 1, synthetic routes of simplified analogues 107 and
used anticancer agents for the treatment of ovarian, breast, and 108 were significantly shortened (∼30 steps). Interestingly,
lung cancer. These compounds have potent antitumor activity compounds 107 and 108 displayed potent activity in the
via binding to β-tubulin. However, broad clinical application CHIKV cell protective assay with an EC50 value of 0.13 μM
resulted in significant multidrug resistance (MDR), which was and 0.8 μM, respectively.191 Moreover, compound 108
a significant obstacle for the success of chemotherapy in displayed no observable toxicity (CC50 > 50 μM). In 2014,
different types of cancers.179,180 One major mechanism of Wender et al. reported a new class of simplified analogues in
MDR is the overexpression of P-glycoprotein (P-gp), a well- which the complex A/B-ring system of bryostatin 1 was
known trans-membrane ABC transporter, which effluxes a replaced with simple salicylate-derived fragments (109, Figure
variety of chemotherapeutic drugs out of tumor cells.180−182 21).197,198 These analogues retained potent affinity (nano-
molar) for PKC isoforms, whereas the synthetic route was
Taxuspine X (99), a Taxol analogue that was isolated from the
greatly shortened to 23−25 steps. Analogue 109 displayed
Japanese yew Taxus cuspidate, was proven to be a potent MDR
good protective activity (EC50 = 1.4 μM) and lower toxicity
reversing agent.183−185 However, 99 was isolated in very poor
(CC50 > 50 μM). SAR study of various substitutions attached
yields from the natural source, and the total synthesis was
to the hydroxyl group revealed that methoxyl analogue 110
limited by intrinsic difficulties and high costs. To explore the
displayed good protective activity and lower toxicity (CHIKV
biological potential of taxuspine analogues, Botta et al.
EC50 = 2.2 μM, CC50 > 50 μM). However, compound 110
performed structural simplification studies (Figure 19).186,187 displayed a lower affinity to PKC (PKCδ Ki = 1.0 μM),
Simplified diterpenoid 100 was synthesized through a ring- indicating that there might be a PKC-independent mechanism
closing metathesis approach. However, its significant structural for evading CHIKV-mediated cell death. This structural
complexity still represents a synthetic challenge, partly due to simplification study provided a good starting point for the
the presence of six stereocenters. Therefore, further structural design of new drugs for the treatment of CHIKV infections.
simplification was conducted to obtain additional simplified 3.3.2.5. Structural Simplification of Cyclopamine 1.
analogues 101 and 102. Compound 101, bearing a benzoyloxy The steroidal alkaloid cyclopamine 1 (111, Figure 22) was
moiety on C13 of the taxane skeleton, was proven to be an isolated from Veratrum californicum and proven to be
efficient P-gp inhibitor (IC50 = 7.2 μM). responsible for craniofacial birth defects.199,200 Subsequently,
3.3.2.3. Structural Simplification of Cortistatin A. mechanism studies revealed that cyclopamine 1 blocked the
Cortistatin A (103, Figure 20), a novel antiangiogenic steroidal Sonic Hedgehog (SHH) signaling pathway by binding to 7-
alkaloid isolated from an Indonesian marine organism, showed transmembrane protein Smoothened (SMO).201,202 Despite its
potent and selective antitumor activity against human umbilical teratogenicity, cyclopamine 1 showed highly potent activity
vein endothelial cells (HUVECs).188 Compound 103 also against a number of human cancers and holds great promise
significantly inhibited in vitro migration and tubular formation for cancer chemotherapy.203−205 However, its poor aqueous
of HUVECs induced by vascular endothelial growth factor solubility and acid lability hampered its further develop-
(VEGF) or basic fibroblast growth factor (bFGF).188 There- ment.206,207 To address these limitations, Dahmane et al.
fore, this compound is expected to be a promising explored the SAR of cyclopamine 1 by structural simplifica-
antiangiogenic lead compound. Kobayashi et al. designed and tion.207 First, replacement of the C-nor-D-homo steroidal
synthesized simplified analogues by reducing the number of system with the androstane ring and aromatization of rings A
chiral centers and removing a bridged ring (Figure 20).189,190 and F generated simplified analogue 112 as well as the C-17
Although compound 104 (IC50 = 0.035 μM) was less active epimer 114. The C-3 deoxy compound 113 was also prepared.
than cortistatin A in the inhibition of HUVECs, it showed Biological evaluations demonstrated that simplified analogues
potent in vivo antiangiogenic and antitumor activities by oral 112−114 showed comparable SHH inhibitory activities with
administration. After introduction of an amide side chain on improved metabolically stability.208
the scaffold, acetamide analogue 105 displayed a highly potent 3.3.2.6. Structural Simplification of Carolacton.
activity against HUVECs (IC50 = 0.0026 μM) as well as an Carolacton (115, Figure 23) is a NP isolated from
excellent selectivity for KB3−1 cells (IC50 = 8.2 μM).190 myxobacterium Sorangium cellulosum.209 It was unable to
AA DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

affect the viability of planktonic bacteria but effectively kills group.221 The macrocyclic depsipeptide subunit locates
pathogen Streptococcus mutans biofilm cells at low nanomolar outside the binding pocket and forms mainly hydrophobic
concentration.210,211 Earlier SAR work revealed that mod- interaction with surrounding residues.221 Nan’s group
ifications of the 12-member macrolide ring were not performed a structural simplification of largazole in which
tolerated.212,213 Thus, Wuest et al. applied DTS to synthesize the −CH2CONHCH2− group and the C7-methyl group were
16 simplified carolacton-inspired analogues (Figure 23).214 1,3- eliminated to afford ring-opened analogue 131 (Figure 26).222
Disubstituted aryl motifs were rationally selected to mimic the Compared with largazole, analogue 131 displayed decreased
trisubstituted olefin side chain. Among them, compound 120 HDAC1inhibitory activity, but it was more potent against
inhibited the formation of biofilms at concentrations at the HDAC3 and HDAC6. A typical HDAC inhibitor has three
concentration of 63 μM, and compound 121 exhibited a essential pharmacophoric features, namely, a recognition cap
similar effect on S. mutans biofilms at concentrations as low as group, a hydrophobic linker, and a zinc-binding group (ZBG).
500 nM. The results highlighted that 1,3-disubstituted aryl Therefore, the long side chain of compound 131 was replaced
motifs could be used as an bioisosteric substitution of the with the classic hydroxamic acid ZBG to afford a series of
trisubstituted olefins, which significantly reduced the synthetic analogues 132.222 Disappointingly, these hydroxamic acid
difficulty while maintaining the structural integrity. analogues displayed significantly reduced HDAC1 inhibitory
3.3.2.7. Structural Simplification of Promysalin. activity (IC50 values of 2.45−9.69 μM). Further structural
Promysalin (122, Figure 24) is a NP produced by Pseudomonas simplification, including the replacement of the chiral L-valine
putida, and displays potent inhibitory activity against moiety with a carbon chain linker, was conducted to afford
Pseudomonas aeruginosa at nanomolar concentrations,215 analogue 133, which displayed significantly improved inhib-
while Gram-positive bacteria show no susceptibility to it.216 itory activity.222 Furthermore, analogue 134, containing a 2,2′-
Promysalin acts by a unique mode of action that it could bisthiazole moiety instead of the 2,4′-bisthiazole moiety,
inhibit the production of pyoverdine (a siderophore) in P. showed a 2-fold increase in activity compared to that of
putida KT2440, which can actively chelate Fe3+ for its end use compound 133. Further structure modification of 134 afforded
in enzymatic processes and is essential for the growth of analogue 135 via the introduction of a hydrophobic cyclo-
Pseudomonads. Wuest et al. synthesized 16 promysalin propyl group to the thiazole ring, and 135 displayed HDAC1
analogues utilizing DTS (Figure 24).217 Several analogues inhibitory activity (IC50 = 0.07 μM) comparable to that of
displayed comparable inhibitory activity against P. aeruginosa lagazole.222 Moreover, analogue 135 displayed potent and
to promysalin. For example, the simplified analogue 126, broad-spectrum inhibitor activity against various HDAC
obtained by removing the hydroxyl group in the side chain, isoforms except HDAC7 and HDAC9. The simplified structure
displayed potent inhibitory activity against P. aeruginosa strains of 135 has a smaller molecular weight and showed favorable
PAO1 and PA14 with an IC50 value of 5.8 μM and 0.035 μM, pharmacokinetic profiles characterized by high Cmax (7.04 μg/
respectively. The results determined the key structural features mL) and AUC0‑t (12.07 μg·h/mL) values and excellent
responsible for bioactivity and highlighted the central role of absolute bioavailability (F% = 118.7%) in mice (20 mg/kg,
the intramolecular hydrogen-bonding network for binding with P.O.).222 Moreover, this compound showed in vivo efficacy in
iron. ameliorating clinical symptoms of experimental autoimmune
3.3.3. Structural Simplification by Truncating Un- encephalomyelitis (EAE) mice when administered orally. The
necessary Substructures. 3.3.3.1. Structural Simplifica- simplified largazole derivatives are currently under preclinical
tion of Chaetocin. Chaetocin (127, Figure 25), a NP isolated evaluation for the treatment of various tumors.
from the genus Chaetomium, is the first reported inhibitor of 3.3.3.3. Structural Simplification of Lobatamide C.
protein lysine methyltransferase G9a (PKMT G9a).218 It is a Lobatamide C (137, Figure 27) belongs to a unique class of
dimer composed of symmetric structures that contain eight NPs with a common benzolactone core structure and a highly
chiral centers. Due to the structural complexity and functional unsaturated enamide side chain.223 This compound displayed
group density, the total synthesis of chaetocin is highly highly potent inhibitory activities against the NCI’s 60 cell-line
challenging. In the structural simplification of chaetocin, human tumor screen (mean panel GI50 values = 1.6 nM) along
Fujishiro et al. synthesized monomeric derivatives to with unique cellular response profiles.223 Mechanism studies
investigate the importance of the dimeric structure for the revealed that the salicylate enamide macrolides potently
activity (Figure 25).219 Interestingly, monomeric derivative inhibited mammalian vacuolar-type (H+)-ATPases (V-AT-
128 was a more potent PKMT G9a inhibitor (IC50 = 3.3 μM) Pases), which are proton-translocating pumps ubiquitous in
than chaetocin (IC50 = 7.2 μM). Further removal of the indole eukaryotic cells.224−226 To clarify the minimal pharmacophores
moiety of analogue 128 afforded structurally simplified bicyclic required for V-ATPase inhibition, Porco et al. designed and
compound 129, which retained the PKMT G9a inhibitory synthesized acyclic lobatamide analogues via opening the
activity (IC50 = 5.2 μM) of the parent compound. Compared lactone ring (Figure 27).227,228 Biological assays indicated that
with chaetocin, the structure of compound 129 was greatly simplified salicylate enamide analogue 138 displayed potent
simplified and its toxicity was also significantly lower. inhibitory activity toward V-ATPase with an IC50 value of 0.1
3.3.3.2. Structural Simplification of Largazole. Larga- μM.227 Analogue 139, with a steric prenyl substituent, showed
zole (130, Figure 26) is a depsipeptide NP that was isolated slightly higher activity (IC50 = 60 nM).228 The results
from cyanobacterium Symploca sp. by Luesch and co-workers indicated that the simplification of lobatamide C was feasible
in 2008.220 It displayed potent HDAC1 inhibitory activity with and that acyclic salicylate enamide V-ATPase inhibitors
an IC 50 value of 0.0137 μM and exhibited potent required further optimization.
antiproliferative activity and selectivity for cancer cells. In the 3.3.3.4. Structural Simplification of Sanglifehrin A.
crystal structure of HDAC8 in complex with largazole thiol Sanglifehrin A (SFA, 140, Figure 28), an immunosuppressive
(136, Figure 26B), the thiol side chain inserts into the binding NP isolated from Streptomyces sp. A92-308110,229,230 exhibited
pocket and chelates the zinc in the active site with its thiol high affinity (Kd = 6.9 nM) to cyclophilin A and inhibited the
AB DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

peptidyl-prolyl cis−trans isomerase (PPIase) activity of cyclo- 3.3.3.5. Structural Simplification of Anguinomycin.
philin A (Ki = 12.8 nM) in a cell-free assay.231 It was The polyketide NPs of the leptomycin family have attracted
investigated as a potential treatment for hepatitis C virus substantial interest from chemists and biologists because of
(HCV). However, SFA has a complex molecular structure their highly complex structure and excellent antitumor
consisting of a 22-membered macrocycle bearing a nine-carbon activity.235,236 Unfortunately, drug development of leptomycin
tether at the 23 position terminated by a highly substituted B (150, Figure 29) failed in phase I clinical trials due to
spirobicyclic moiety, which poses a tremendous challenge for toxicity to patients.237 In contrast, two related compounds,
the synthesis of new analogues. The crystal structure of SFA- anguinomycins C and D (151 and 152), showed excellent
cyclophilin A complex revealed that the 22-membered antitumor activities at low nanomolar concentrations and
macrocycle scaffold formed a hydrogen bond network with induced apoptosis in pRB-inactivated tumor cells while only
residues Gln63, Arg55, His126, and Asn102 (Figure 28B).232 inducing growth arrest in normal cells.235,236 A mechanism
In contrast, the terminal substituted spirobicyclic moiety was study revealed that these antitumor antibiotics inhibited the
oriented out of the binding domain, indicating that it was nuclear export receptor chromosome region maintenance 1
unnecessary for the binding activity.232 For the structural (CRM1).238−240 Gademann et al. investigated the importance
simplification of SFA, Wagner et al. performed the selective of the polyketide side chain for the antitumor potency. After
oxidative cleavage of the C26C27 exocyclic double bond to the side chain was replaced with a simpler linear terpene,
afford the spirolactam fragment and macrolide 141 (Figure compound 153 displayed similar activity in blocking
28).233 Biological assays indicated that macrolide 141 nucleocytoplasmic transport.241 Further reducing the length
displayed potent binding affinity for cyclophilin (Kd = 29 of the polyketide side chain led to the very simple analogue
nM), which was slightly less active than SFA, indicating the 154, which effectively blocked nuclear export at 25 nM.241 The
macrocyclic portion of SFA was the key fragment for the simplicity, good synthetic accessibility, and excellent activity of
activity. Moreover, the removal of the C24 polyketide side compound 154 make it a particularly attractive candidate for
chain eliminated the immunosuppressive activity. Further further development.
degradation studies afforded analogue 142, which has an 3.3.3.6. Structural Simplification of Caprazamycin B.
allylic alcohol side chain, and it displayed improved binding Caprazamycin B (CPZ-B, 155, Figure 30), a nucleoside NP
affinity for cyclophilin (Kd = 5.7 nM). extracted from the culture broth of the actinomycete
On the basis of analogue 142, Schultz et al. performed a Streptomyces sp. MK730−62F2,242 demonstrated excellent in
further step-by-step structural simplification study (Figure 28, vitro antimicrobial activity against both drug-susceptible and
Table 8).234 After analyzing the cocrystal structure of 142 with multidrug-resistant Mycobacterium tuberculosis strains with low
cyclophilin A, they envisioned that the C23 and C14 side toxicity to mice at single doses up to 200 mg/kg.242 CPZ-B was
chains and the two hydroxyl groups might be unnecessary a new member of the 6′-N-alkyl-5′-β-O-aminoribosoyl-C-
(Figure 28C).233 Therefore, the C23 and C14 side chains were glycluridine class of antibiotics, which act by targeting
removed, and the two hydroxyls were methylated to afford phospho-MurNAc-pentapeptide translocase (MraY) to block
simplified analogue 143. This compound displayed potent the formation of lipid I and peptidoglycan biosynthesis.243−247
binding affinity for cyclophilin A (Kd = 25 nM) and sub- On the basis of its excellent biological properties, CPZ-B was
micromolar antiviral activity in the HCV replicon cellular assay expected to be a promising antibacterial drug candidate.
(EC50 = 600 nM). Similar to 142, analogue 143 retained the However, the total synthesis of CPZ-B remains challenging.
hydrogen bonding interaction with residues Arg55, Gln63, Only recently have Takemoto et al. reported a 23-step total
Asn102, and His126 (Figure 28D).234 Furthermore, the C14 synthesis for the related compound caprazamycin A.248 This
to C17 stereocenters were investigated using analogues 144− fact, combined with its extremely poor water solubility,
145, which were synthesized in only a few steps. Biological hampered the clinical development of this compound.249
assays indicated that the chiral centers at C16 and C17 were Therefore, Matsuda et al. reported the rational simplification of
not essential for the activity. Subsequent replacement of the CPZ-B through FOS.247 First, the diazepanone ring of CPZ-B
C18−C21 diene unit with a styryl group led to potent was proven to be nonessential for the antibacterial activity.
analogues 146 and 147. Analogue 147 displayed the most Then, the diazepanone and aminoribose units were replaced
potent binding affinity for cyclophilin A (Kd = 11 nM) and with an oxazolidine moiety to give oxazolidine analogue 156
sub-micromolar antiviral activity in the HCV replicon cellular (Figure 30). Biological evaluation revealed that compound 156
assay (EC50 = 36 nM). The crystal structure of analogue 147 in showed good antibacterial potency against drug-susceptible S.
complex with cyclophilin A displayed a novel binding mode in aureus, E. faecalis, and E. faecium and drug-resistant bacterial
which the m-tyrosine (m-Tyr) residue was displaced into the strains (e.g., MRSA and VRE strains) with MIC values ranging
solvent (Figure 28F).234 On the basis of this observation, they from 2 to 16 μg/mL.247 Compound 156 can be synthesized in
further simplified the scaffold and replaced the m-Tyr with an just 12 steps, which makes it a good candidate for further
alanine to afford analogues 148 and 149. Analogue 148 development.
displayed an excellent binding affinity for cyclophilin A (Kd = 3.3.3.7. Structural Simplification of Militarinone D. A
24 nM) with potent cellular activity (EC50 = 87 nM), number of pyridone alkaloids with diverse bioactivities and
confirming that the m-Tyr residue was unnecessary in these pharmaceutical relevance have been isolated from entomopa-
new analogues. The binding mode of compound 148 was thogenic Deuteromycetes fungi. Among these compounds,
similar to that of SFA, whose macrocycle moiety formed militarinone D (157, Figure 31) was discovered from a
hydrogen bonds with residues Asn102, Arg55, and Gln63 mycelial extract of the entomopathogenic fungus Paecilomyces
(Figure 28G).234 After removal of the C14 to C17 stereo- militaris. (RCEF 0095) and identified through activity-directed
centers of compound 148, analogue 149 only showed a 2-fold fractionation.250,251 Militarinone D showed potent neurito-
decrease in binding affinity, and its cellular activity was genic activity in PC-12 cells and significantly induced neurite
retained. outgrowth in the absence of a nerve growth factor
AC DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(NGF).251−253 In 2011, Gademann et al. reported a unified and metastasis.271,272 Rakicidin A (164, Figure 34) is a
approach for the stereoselective total synthesis of pyridine macrocyclic depsipeptide isolated from culture broth of
alkaloids and evaluated their neuritogenic activities in a Micromonospora sp. strain R385-2.273 In 2011, Ashihara’s
standardized assay with PC-12 cells.254 The results revealed group demonstrated that rakicidin A exhibited selective growth
that neither the side chain length nor the absolute inhibitory activity against hypoxic cancer cells and inhibited
configuration was the essential pharmacophore. Therefore, the proliferation of imatinib-resistant, hypoxia-adapted chronic
Gademann’s group synthesized truncated analogue 158 by myelogenous leukemia (HA-CML) cells by inducing apopto-
deleting the side chain, and this compound showed neurite sis.274 Structurally, rakicidin A is a cyclic depsipeptide
outgrowth activity comparable to that of militarinone D at 20 consisting of 4-amino-penta-2,4-dienoic acid (APD, red in
μM.255 Figure 34), 3-hydroxy-2,4,16-trimethylheptadecanoic acid,
3.3.3.8. Structural Simplification of Bistramide A. sarcosine, and β-hydroxyasparagine (β-HOAsn) moieties.
Bistramide A (159, Figure 32) is a marine-derived NP isolated This compound belongs to the APD-containing cyclolipodep-
from the ascidian Lissoclinum bistratum.256−258 In addition to sipeptide (APD-CLD) class of NPs with five chiral centers.275
highly potent antitumor activity against A549 lung cancer cell Due to its structural complexity and unique biological activity,
lines, bistramide A was also confirmed to alter the voltage rakicidin A is of particular interest for both synthetic and
dependence of muscle twitch tension and inhibit Na+ medicinal studies. Chen et al. reported the first total synthesis
conductance.259,260 Initially, bistramide A was thought to of rakicidin A and determined that the absolute configuration
selectively activate PKC δ.261 However, Kozmin et al. proved of rakicidin A is 2S,3S,14S,15S,16R.276 SAR studies revealed
that actin was the primary target of bistramide A.261,262 that the APD functionality was essential for the biological
Structurally, bistramide A can be divided into three parts, activities of rakicidin A.276−279 Guided by the SAR, Poulsen et
namely, C(19)−C(38) lipophilic subunit A, C(13)−C(18) al. synthesized simplified analogue 165 by pruning the β-
hydrophilic subunit B, and C(1)−C(13) reactive subunit C HOAsn side chain and the C14 and C16 methyl groups, and
(Figure 32A).263 The X-ray crystal structure of the actin- this derivative showed potent and selective growth inhibition
bistramide A complex revealed that subunits A and B spanned under hypoxic conditions with a GI50 value of 352 nM.279
the entire deep binding cleft of actin through an network of Although the potency was 10-fold lower, compound 165
hydrogen bonds and van der Waals interactions (Figure showed better hypoxic selectivity (>67-fold) than rakicidin A.
32B).264 In contrast, subunit C was oriented out of the binding 3.3.3.11. Structural Simplification of Spongistatins.
domain and faced to the solvent. Subunit C was proved to be The spongipyran NPs, such as spongistatin 1 (166, altohyrtin
unnecessary for the binding affinity and could be trun- A, Figure 35) and spongistatin 2 (167, altohyrtin C, Figure
cated.263,265 Thus, simplified analogue 160 was designed, 35), are a unique family of bis(spiroacetal) macrolides which
which simultaneously retained the G-actin affinity and reduced were isolated from marine sponges in the early 1990s.280−283
the structural complexity of the parent compound.263 They displayed exceptional cytotoxicity against a wide range of
3.3.3.9. Structural Simplification of Withaferin A. The human cancer cell lines, acting by interfering with tubulin
withanolides constitute a class of NPs characterized by an polymerization.284 Spongistatin 1 had an average IC50 of 0.13
ergostane skeleton with a modified side chain in which carbons nM against the panel of 60 cancer cell lines, which was thought
22 and 26 are oxidized to form a δ-lactone.170 To date, over to be the most potent antiproliferative agent ever discov-
600 members of this family have been isolated and ered.285 Although it shows great promise in cancer chemo-
characterized, and they have shown a broad range of biological therapy, the extreme scarcity of natural supply and synthetic
activities including anti-inflammatory, antitumor, antimicrobial, difficulty have halted its further preclinical development. In
and immunoregulatory activities as well as antifeedant and order to obtain SAR and identify the spongipyran pharmaco-
insecticidal activity.170,266 Withaferin A (161, Figure 33) was phore, several groups devoted their efforts to simplify this
the first withanolide analogue isolated from Withania somnifera highly complex NP. Smith et al. reported two simplified
Dun (Solanaceae) by Lavie in 1965.267 The compound was analogues (compounds 168 and 169) bearing the F ring and
reported to trigger apoptosis and largely suppress cell the C44−C51 triene side-chain.286 However, these analogues
migration/invasion in MDA-MB-231 human breast cancer only displayed micromolar inhibitory activity against several
cells depending on the inhibition of STAT3 phosphoryla- human cancer cell lines. Furthermore, the same group
tion.268 Withacnistin (162, Figure 33), a C18-functionalized synthesized the AB-spiroacetal spongistatin derivatives 170
derivative of withaferin A, was also identified as a potent and 171.287 Neither of them had significant cytotoxic or
STAT3 inhibitor and induced regression in breast tumor antitubulin activity. Smith et al. synthesized two diminutive
cells.268,269 Considering their potent inhibitory activities derivatives of spongistatin 1 (172 and 173, Figure 35).285
against STAT3, withaferin A and withacnistin were inves- Analogue 172 displayed no significant cytotoxic activity (IC50
tigated as potential lead compounds for antitumor drug > 1 μM). In contrast analogue 173 displayed potent cytotoxic
discovery. The FOS approach was applied to synthesize activity against U937 (IC50 = 60 nM) and MDA-MB-435 (IC50
simplified withaferin A analogues with maintained or improved = 83 nM).
biological activity.270 After truncating of the δ-lactone side A less drastic simplification of spongistatin 1 was performed
chain, compound 163 showed STAT3 inhibitory activity by Paterson et al.288 After dehydration at E-ring C35 position,
(more than 50% inhibition of STAT3 phosphorylation at 1 compound 174 (Figure 36) showed enhanced cytotoxicity at
μM) comparable to that of compound 161 in MDA-MB-231 the low picomolar level (Table 9), whereas truncation of the
breast cancer cell lines. side-chain at C46 (compound 175, Figure 36) resulted in an
3.3.3.10. Structural Simplification of Rakicidin A. obvious decrease of the activity (Table 9).288 The SAR
Tumor hypoxia is a common feature of most tumors and is revealed that E-ring modification could be tolerated instead of
highly correlated to poor prognosis of cancer patients owing to F-ring side-chain truncation. Furthermore, Wagner et al.
its multiple contributions to chemo-resistance, radio-resistance, synthesized several structurally simplified analogues containing
AD DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

only part of the spongistatin structure.289 However, truncation Besides DOS, DTS, and FOS, the development of new
of C1−C28 (176 and 177) and replacement of the two highly efficient synthesis methods,111,309 such as chemical or
spiroketals and their associated functionality by a simple alkane molecular editing310 and dynamic strategic bond analysis,311
linker (lactone 178) led to loss of the activity. Similarly, will provide a deep understanding of SAR and biological
replacing C, D rings or E, F rings by the alkane linker (179 and functions of complex NPs, and lay an important foundation for
180) also resulted in weak cytotoxicity (Figure 36). structural simplification. These types of approaches are
3.3.3.12. Structural Simplification of Migrastatin. particularly useful in the development of antibacterial NPs,
Tumor metastasis is one of the main causes of death in cancer which require three-dimensionality to retain the activity.312,313
patients. Migrastatin (181, Figure 37), a novel 14-membered They offer a new opportunity to solve the difficulty in
lactone isolated from Streptomyces sp. MK-929−43F1, showed synthesizing complex structures and their simplified analogues.
moderate antimetastatic activity (IC50 = 29 μM) against 4T1 Moreover, new technologies are still highly desirable for
mouse breast tumor cells.290,291 Despite the modest potency, improving the efficiency of structural simplification. In
Danishefsky and co-workers envisioned that migrastatin could particular, the BIOS strategy employs biologically relevant
be used as a lead compound to discover more potent structural information encoded in complex NPs to guide the
antimetastatic and antiangiogenic agents. They reported the simplification of NP structures.171 Computational tools for the
first total synthesis of migrastatin and then performed fragmentation of NPs enable hierarchical structuring, visual-
structural simplification studies using the DTS approach ization, and analysis of complex structures and bioactivity data,
(Figure 37).292−295 After removal of the glutarimide side and the identification of simplified scaffolds with desirable
chain, simplified analogues 183 and 184 were 1000-fold more activities. The resulting NP fragments have been proven to be
potent than migrastatin. These compounds significantly a rich source of bioactive molecules.4
inhibited cell migration in 4T1 mouse breast cancer cells Ideally, structural simplification should identify less
with IC50 values of 22 nM and 24 nM, respectively. complicated molecules that retain the biological activity of
Interestingly, these analogues did not show any cytotoxic or the parent compound. However, simplified molecules were
antiproliferative activity up to 20 μM.292 On the basis of often less active than the lead compound at the molecular or
analogue 183, further structural simplification was performed cellular level. Thus, the success of structural simplification
by modifying the ester group. Analogues 185 and 186 retained needs to be evaluated in terms of the balance between the
potent antimetastatic activity in MDA-MB-435 breast cancer synthetic accessibility, in vitro activity, in vivo potency,
cells with an IC50 value of 0.10 μM and 0.37 μM, respectively. physicochemical properties, PK profiles, and so on. For most
Moreover, 186 showed good in vivo potency by significantly simplified molecules, reoptimization is required to improve the
reducing the metastasized MDA231 cells (87%) in a pretreated efficacy and drug-likeness. More importantly, reductions in
mice group. It could not attenuate tumor growth, which structural complexity will lead to simplified analogues with new
indicated the inhibition was not related to cytotoxicity.295 To bioactivities and new targets. Target identification by chemical
elucidate its mechanism of action, Chen et al. demonstrated biology approaches is important for drug development. With
that migrastatin analogues targeted actin-bundling protein the advancement of synthetic, computational, and biological
fascin to block tumor metastasis using an affinity protein approaches, structural simplification will play an increasingly
purification approach (Figure 37B).296 important role in drug discovery and contribute to improving
3.3.4. Recent Examples of Structural Simplification. the efficiency and success rate of drug development.
Besides the case studies described above, examples of
structurally simplified NPs reported in recent years are AUTHOR INFORMATION
summarized in Table 10. Corresponding Author
*Phone/Fax: +86 21 81871239. E-mail: shengcq@smmu.edu.
4. CONCLUSIONS AND PERSPECTIVES cn.
The design of “low-fat” drug-like molecules by structural ORCID
simplification represents a powerful strategy in drug discov-
ery.25−27 Structural simplification is particularly useful in NPs- Shengzheng Wang: 0000-0002-2383-2673
based drug development because NPs are challenging for Chunquan Sheng: 0000-0001-9489-804X
direct drug development because of their complex structures, Author Contributions
abundant chiral centers, difficult total syntheses, and #
S.W. and G.D. contributed equally to this perspective.
unfavorable ADMET profiles. The structural simplification of Notes
NPs has yielded several marketed drugs highlighting its utility
in drug development. Recently, numerous NPs with extremely The authors declare no competing financial interest.
complex structures, particularly marine NPs, have been shown Biographies
to possess highly potent biological activities. These NPs usually
have a MW greater than 1000 and more than 5 chiral centers, Shengzheng Wang received his bachelor’s degree in pharmacy (2009)
which pose great challenges in total synthesis and drug and Ph.D. in medicinal chemistry (2014) from Second Military
development. Structural simplification of these extremely Medical University. Since 2014, he has been a lecturer in the
complex NPs is highly desirable. Another important area for Department of Medicinal Chemistry, School of Pharmacy, Fourth
structural simplification is the development of new anti- Military Medical University. His research interests include antitumor
bacterial agents, which require three-dimensionality, but drug discovery and organocatalysis.
complex structures are difficult to synthesize and commerci- Guoqiang Dong received his bachelor’s degree in pharmacy (2008)
alize. Although progress has been made, successful examples of and Ph.D. in medicinal chemistry (2013) from Second Military
structural simplification are rather limited, and more medicinal Medical University. Since 2013, he has been a lecturer in the
chemistry efforts need to be devoted in this important field. Department of Medicinal Chemistry, School of Pharmacy, Second

AE DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Military Medical University. His current research is focused on the bFGF basic fibroblast growth factor
modification and scaffold hopping of natural products and multi- V-ATPases vacuolar-type (H+)-ATPases
targeting antitumor drugs. SFA Sanglifehrin A
Chunquan Sheng received his bachelor’s degree in pharmacy (2000) HCV hepatitis C virus
and Ph.D. in medicinal chemistry (2005) from Second Military FtsZ filamenting temperature-sensitive protein Z
Medical University, after which he worked as a faculty member in the CHIKV Chikungunya virus
same university. From 2014 to 2015, he worked as a visiting scholar at CRM1 chromosome region maintenance 1
Department of Chemistry and Chemical Biology in University of New CPZ-B caprazamycin B
Mexico, USA. He is currently Professor and Director of Department MraY phospho-MurNAc-pentapeptide translocase
of Medicinal Chemistry and the Vice Dean of School of Pharmacy, NGF nerve growth factor
Second Military Medical University. Professor Sheng’s research HA-CML hypoxia-adapted chronic myelogenous leukemia
interests mainly include rational drug design and natural products- APD 4-amino-penta-2,4-dienoic acid
inspired drug discovery focusing on antifungal and antitumor agents. β-HOAsn β-hydroxyasparagine
He was supported by several talent programs in China including APD-CLD APD-containing cyclolipodepsipeptide
National Outstanding Young Scholarship (NSFC) and Young Chang SOAT sterol O-acyltransferase
Jiang Scholar (MOE). SHH sonic hedgehog
SMO smoothened
PB pladienolide
ACKNOWLEDGMENTS Top topoisomerase
This work was supported by the National Natural Science MptpB Mycobacterium tuberculosis protein tyrosine phos-
Foundation of China (Grants 81725020 to C.S. and 21602252 phatases B
to S.W.), the National Key R&D Program of China Hh hedgehog
(2017YFA0506000 to C.S.), and the Hong Kong Scholars
Program (XJ201713 to S.W.). REFERENCES
(1) Cragg, G. M.; Newman, D. J. Natural products: a continuing
ABBREVIATIONS source of novel drug leads. Biochim. Biophys. Acta, Gen. Subj. 2013,
NP natural product 1830, 3670−3695.
(2) Butler, M. S.; Robertson, A. A.; Cooper, M. A. Natural product
FDA U.S. Food and Drug Administration and natural product derived drugs in clinical trials. Nat. Prod. Rep.
ADMET absorption, distribution, metabolism, excretion, 2014, 31, 1612−1661.
metabolism, and toxicity (3) Harvey, A. L.; Edrada-Ebel, R.; Quinn, R. J. The re-emergence of
RO5 Lipinski’s rule of five natural products for drug discovery in the genomics era. Nat. Rev.
PK pharmacokinetic Drug Discovery 2015, 14, 111−129.
PD pharmacodynamic (4) Crane, E. A.; Gademann, K. Capturing Biological Activity in
SAR structure−activity relationship Natural Product Fragments by Chemical Synthesis. Angew. Chem., Int.
DOS diversity-oriented synthesis Ed. 2016, 55, 3882−3902.
DTS diverted total synthesis (5) Xiao, Z.; Morris-Natschke, S. L.; Lee, K. H. Strategies for the
BIOS biology-oriented synthesis Optimization of Natural Leads to Anticancer Drugs or Drug
Candidates. Med. Res. Rev. 2016, 36, 32−91.
FOS function-oriented synthesis (6) Shen, B. A. New Golden Age of Natural Products Drug
BGC biosynthetic gene cluster Discovery. Cell 2015, 163, 1297−1300.
MW molecular weight (7) Newman, D. J.; Cragg, G. M. Natural Products as Sources of
LE ligand efficiency New Drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629−661.
NMDA N-methyl-D-aspartate (8) Kato, N.; Takahashi, S.; Nogawa, T.; Saito, T.; Osada, H.
HBV hepatitis B virus Construction of a microbial natural product library for chemical
S1P sphingosine-1-phosphate biology studies. Curr. Opin. Chem. Biol. 2012, 16, 101−108.
TSA trichostatin A (9) Hong, J. Role of natural product diversity in chemical biology.
HDAC histone deacetylase Curr. Opin. Chem. Biol. 2011, 15, 350−354.
PKC protein kinase C (10) Kakeya, H. Natural products-prompted chemical biology:
phenotypic screening and a new platform for target identification. Nat.
JAK Janus kinase Prod. Rep. 2016, 33, 648−654.
CCK cholecystokinin receptors (11) Pascolutti, M.; Quinn, R. J. Natural products as lead structures:
BZD 1,4-benzodiazepine chemical transformations to create lead-like libraries. Drug Discovery
L-Trp L-tryptophan Today 2014, 19, 215−221.
GPCR G-protein coupled receptor (12) Eberhardt, L.; Kumar, K.; Waldmann, H. Exploring and
MCR multicomponent reaction exploiting biologically relevant chemical space. Curr. Drug Targets
nAChR nicotinic acetylcholine receptor 2011, 12, 1531−1546.
PKMT protein lysine methyltransferase (13) Breinbauer, R.; Vetter, I. R.; Waldmann, H. From protein
ZBG zinc-binding group domains to drug candidates-natural products as guiding principles in
EAE experimental autoimmune encephalomyelitis the design and synthesis of compound libraries. Angew. Chem., Int. Ed.
2002, 41, 2878−2890.
IFI invasive fungal infection (14) Kellenberger, E.; Hofmann, A.; Quinn, R. J. Similar interactions
GA gambogic acid of natural products with biosynthetic enzymes and therapeutic targets
MDR multidrug resistance could explain why nature produces such a large proportion of existing
P-gp P-glycoprotein drugs. Nat. Prod. Rep. 2011, 28, 1483−1492.
HUVEC human umbilical vein endothelial cell (15) Nicolaou, K. C.; Hale, C. R.; Nilewski, C.; Ioannidou, H. A.
VEGF vascular endothelial growth factor Constructing molecular complexity and diversity: total synthesis of

AF DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

natural products of biological and medicinal importance. Chem. Soc. (40) Davis, T. D.; Gerry, C. J.; Tan, D. S. General platform for
Rev. 2012, 41, 5185−5238. systematic quantitative evaluation of small-molecule permeability in
(16) Feher, M.; Schmidt, J. M. Property distributions: differences bacteria. ACS Chem. Biol. 2014, 9, 2535−2544.
between drugs, natural products, and molecules from combinatorial (41) Noren-Muller, A.; Reis-Correa, I., Jr.; Prinz, H.; Rosenbaum,
chemistry. J. Chem. Inf. Comput. Sci. 2003, 43, 218−227. C.; Saxena, K.; Schwalbe, H. J.; Vestweber, D.; Cagna, G.; Schunk, S.;
(17) Henkel, T.; Brunne, R. M.; Muller, H.; Reichel, F. Statistical Schwarz, O.; Schiewe, H.; Waldmann, H. Discovery of protein
Investigation into the Structural Complementarity of Natural phosphatase inhibitor classes by biology-oriented synthesis. Proc. Natl.
Products and Synthetic Compounds. Angew. Chem., Int. Ed. 1999, Acad. Sci. U. S. A. 2006, 103, 10606−10611.
38, 643−647. (42) Cerri, A.; Gobbini, M. Simplified digitalis-like compounds
(18) Rodrigues, T.; Reker, D.; Schneider, P.; Schneider, G. Counting acting on Na(+), K(+)-ATPase. J. Enzyme Inhib. Med. Chem. 2003,
on natural products for drug design. Nat. Chem. 2016, 8, 531−541. 18, 289−295.
(19) Koehn, F. E.; Carter, G. T. The evolving role of natural (43) Eguchi, M. Recent advances in selective opioid receptor
products in drug discovery. Nat. Rev. Drug Discovery 2005, 4, 206− agonists and antagonists. Med. Res. Rev. 2004, 24, 182−212.
220. (44) Blakemore, P. R.; White, J. D. Morphine, the Proteus of organic
(20) Patridge, E.; Gareiss, P.; Kinch, M. S.; Hoyer, D. An analysis of molecules. Chem. Commun. 2002, 11, 1159−1168.
FDA-approved drugs: natural products and their derivatives. Drug (45) Corbett, A. D.; Paterson, S. J.; Kosterlitz, H. W. Selectivity of
Discovery Today 2016, 21, 204−207. Ligands for Opioid Receptors. In Opioids; Herz, A.; Akil, H.; Simon,
(21) Yao, H.; Liu, J.; Xu, S.; Zhu, Z.; Xu, J. The structural E. J., Eds.; Springer: Berlin, 1993; pp 645−679.
modification of natural products for novel drug discovery. Expert (46) Gudin, J.; Fudin, J.; Nalamachu, S. Levorphanol use: past,
Opin. Drug Discovery 2017, 12, 121−140. present and future. Postgrad. Med. 2016, 128, 46−53.
(22) Chen, J.; Li, W.; Yao, H.; Xu, J. Insights into drug discovery (47) Heel, R. C.; Brogden, R. N.; Speight, T. M.; Avery, G. S.
from natural products through structural modification. Fitoterapia Butorphanol: a review of its pharmacological properties and
2015, 103, 231−241. therapeutic efficacy. Drugs 1978, 16, 473−505.
(23) Maier, M. E. Design and synthesis of analogues of natural (48) Commiskey, S.; Fan, L. W.; Ho, I. K.; Rockhold, R. W.
products. Org. Biomol. Chem. 2015, 13, 5302−5343. Butorphanol: effects of a prototypical agonist-antagonist analgesic on
(24) Guo, Z. The modification of natural products for medical use. kappa-opioid receptors. J. Pharmacol. Sci. 2005, 98, 109−116.
Acta Pharm. Sin. B 2017, 7, 119−136. (49) Gharagozlou, P.; Hashemi, E.; DeLorey, T. M.; Clark, J. D.;
(25) Hann, M. M. Molecular obesity, potency and other addictions Lameh, J. Pharmacological profiles of opioid ligands at kappa opioid
in drug discovery. MedChemComm 2011, 2, 349−355. receptors. BMC Pharmacol. 2006, 6, 3.
(26) Hann, M. M.; Keseru, G. M. Finding the sweet spot: the role of (50) Aarnes, T. K.; Muir, W. W. Chapter 26 - Pain Assessment and
nature and nurture in medicinal chemistry. Nat. Rev. Drug Discovery Management. In Small Animal Pediatrics; Peterson, M. E.; Kutzler, M.
A., Eds.; W.B. Saunders: Saint Louis, 2011; pp 220−232.
2012, 11, 355−365.
(51) Prezzavento, O.; Arena, E.; Sánchez-Fernández, C.; Turnaturi,
(27) Mignani, S.; Huber, S.; Tomas, H.; Rodrigues, J.; Majoral, J. P.
R.; Parenti, C.; Marrazzo, A.; Catalano, R.; Amata, E.; Pasquinucci, L.;
Compound high-quality criteria: a new vision to guide the
Cobos, E. J. (+)-and (−)-Phenazocine enantiomers: Evaluation of
development of drugs, current situation. Drug Discovery Today
their dual opioid agonist/σ1 antagonist properties and antinociceptive
2016, 21, 573−584.
effects. Eur. J. Med. Chem. 2017, 125, 603−610.
(28) Polanski, J.; Bogocz, J.; Tkocz, A. The analysis of the market
(52) Levine, J. D.; Gordon, N. C. Synergism between the analgesic
success of FDA approvals by probing top 100 bestselling drugs. J.
actions of morphine and pentazocine. Pain 1988, 33, 369−372.
Comput.-Aided Mol. Des. 2016, 30, 381−389. (53) Woolfe, G.; Macdonald, A. D. The evaluation of the analgesic
(29) Bottcher, T. An Additive Definition of Molecular Complexity. J. action of pethidine hydrochloride (demerol). J. Pharmacol. Exp. Ther.
Chem. Inf. Model. 2016, 56, 462−470. 1944, 80, 300−307.
(30) Hann, M. M.; Leach, A. R.; Harper, G. Molecular complexity (54) Ziering, A.; Lee, J. Piperidine derivatives; 1,3-dialkyl-4-aryl-4-
and its impact on the probability of finding leads for drug discovery. J. acyloxypiperidines. J. Org. Chem. 1947, 12, 911−914.
Chem. Inf. Comput. Sci. 2001, 41, 856−864. (55) Van Bever, W. F. M.; Niemegeers, C. J. E.; Janssen, P. A. J.
(31) Mendez-Lucio, O.; Medina-Franco, J. L. The many roles of Synthetic analgesics. Synthesis and pharmacology of the diaster-
molecular complexity in drug discovery. Drug Discovery Today 2017, eoisomers of N-[3-methyl-1-(2-phenylethyl)-4-piperidyl]-N-phenyl-
22, 120−126. propanamide and N-[3-methyl-1-(1-methyl-2-phenylethyl)-4-piperid-
(32) Schneider, G.; Neidhart, W.; Giller, T.; Schmid, G. ’Scaffold- yl]-N-phenylpropanamide. J. Med. Chem. 1974, 17, 1047−1051.
Hopping’ by topological pharmacophore search: A contribution to (56) Barnett, C. J. Modification of methadone synthesis process step.
virtual screening. Angew. Chem., Int. Ed. 1999, 38, 2894−2896. U.S. Patent US4048211A, 1977.
(33) Bohm, H. J.; Flohr, A.; Stahl, M. Scaffold hopping. Drug (57) King, J. A.; Meltzer, R. I.; Doczi, J. The Synthesis of Some
Discovery Today: Technol. 2004, 1, 217−224. Fluorene Derivatives1. J. Am. Chem. Soc. 1955, 77, 2217−2223.
(34) Sun, H.; Tawa, G.; Wallqvist, A. Classification of scaffold- (58) Liu, G. T. Bicyclol: a novel drug for treating chronic viral
hopping approaches. Drug Discovery Today 2012, 17, 310−324. hepatitis B and C. Med. Chem. 2009, 5, 29−43.
(35) Pan, S.; Zhang, H.; Wang, C.; Yao, S. C.; Yao, S. Q. Target (59) Liu, G. T. Therapeutic effects of biphenyl dimethyl
identification of natural products and bioactive compounds using dicarboxylate (DDB) on chronic viral hepatitis B. Proc. Chin. Acad.
affinity-based probes. Nat. Prod. Rep. 2016, 33, 612−620. Med. Sci. Peking Union Med. Coll. 1987, 2, 228−233.
(36) Farha, M. A.; Brown, E. D. Strategies for target identification of (60) Nie, J.; Yang, D.; Hu, K.; Lu, Y. Study on four polymorphs of
antimicrobial natural products. Nat. Prod. Rep. 2016, 33, 668−680. bifendate based on X-ray crystallography. Acta Pharm. Sin. B 2016, 6,
(37) Frantz, S. Better antibiotics through chemistry. Nat. Rev. Drug 234−242.
Discovery 2004, 3, 900−901. (61) Liu, G. T. The anti-virus and hepatoprotective effect of bicyclol
(38) O’Shea, R.; Moser, H. E. Physicochemical properties of and its mechanism of action. Chin. J. New Drugs 2001, 10, 325−327.
antibacterial compounds: implications for drug discovery. J. Med. (62) Zhou, P.; Huang, L.; Zhang, Z.; Wang, L.; Huo, S.; Lei, Z.
Chem. 2008, 51, 2871−2878. Simple method for the preparation of bicyclol from bifendate. Chin.
(39) Lipinski, C. A.; Lombardo, F.; Dominy, B. W.; Feeney, P. J. Patent CN103724317A, 2014.
Experimental and computational approaches to estimate solubility and (63) Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The Traditional Medicine
permeability in drug discovery and development settings. Adv. Drug and Modern Medicine from Natural Products. Molecules 2016, 21,
Delivery Rev. 2001, 46, 3−26. 559.

AG DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(64) Sun, H.; Yu, L.; Wei, H.; Liu, G. A novel antihepatitis drug, Littlefield, B. A. In vitro and in vivo anticancer activities of synthetic
bicyclol, prevents liver carcinogenesis in diethylnitrosamine-initiated macrocyclic ketone analogues of halichondrin B. Cancer Res. 2001, 61,
and phenobarbital-promoted mice tumor model. J. Biomed. Biotechnol. 1013−1021.
2012, 2012, 584728. (82) Doodhi, H.; Prota, A. E.; Rodriguez-Garcia, R.; Xiao, H.;
(65) Bao, X. Q.; Liu, G. T. Bicyclol: a novel antihepatitis drug with Custar, D. W.; Bargsten, K.; Katrukha, E. A.; Hilbert, M.; Hua, S.;
hepatic heat shock protein 27/70-inducing activity and cytoprotective Jiang, K.; Grigoriev, I.; Yang, C. P.; Cox, D.; Horwitz, S. B.; Kapitein,
effects in mice. Cell Stress Chaperones 2008, 13, 347−355. L. C.; Akhmanova, A.; Steinmetz, M. O. Termination of Protofilament
(66) Fujita, T.; Hirose, R.; Hamamichi, N.; Kitao, Y.; Sasaki, S.; Elongation by Eribulin Induces Lattice Defects that Promote
Yoneta, M.; Chiba, K. 2-Substituted 2-aminoethanol: Minimum Microtubule Catastrophes. Curr. Biol. 2016, 26, 1713−1721.
essential structure for immunosuppressive activity of ISP-I (Myr- (83) Donoghue, M.; Lemery, S. J.; Yuan, W.; He, K.; Sridhara, R.;
iocin). Bioorg. Med. Chem. Lett. 1995, 5, 1857−1860. Shord, S.; Zhao, H.; Marathe, A.; Kotch, L.; Jee, J.; Wang, Y.; Zhou,
(67) Sasaki, S.; Hashimoto, R.; Kiuchi, M.; Inoue, K.; Ikumoto, T.; L.; Adams, W. M.; Jarral, V.; Pilaro, A.; Lostritto, R.; Gootenberg, J.
Hirose, R.; Chiba, K.; Hoshino, Y.; Okumoto, T.; Fujita, T. Fungal E.; Keegan, P.; Pazdur, R. Eribulin mesylate for the treatment of
metabolites. Part 14. Novel potent immunosuppressants, mycester- patients with refractory metastatic breast cancer: use of a ″physician’s
icins, produced by Mycelia sterilia. J. Antibiot. 1994, 47, 420−433. choice″ control arm in a randomized approval trial. Clin. Cancer Res.
(68) Fujita, T.; Yoneta, M.; Hirose, R.; Sasaki, S.; Inoue, K.; Kiuchi, 2012, 18, 1496−1505.
M.; Hirase, S.; Adachi, K.; Arita, M.; Chiba, K. Simple compounds, 2- (84) Omura, S.; Sasaki, Y.; Iwai, Y.; Takeshima, H. Staurosporine, a
alkyl-2-amino-1,3-propanediols have potent immunosuppressive potentially important gift from a microorganism. J. Antibiot. 1995, 48,
activity. Bioorg. Med. Chem. Lett. 1995, 5, 847−852. 535−548.
(69) Thomas, K.; Proschmann, U.; Ziemssen, T. Fingolimod (85) Omura, S.; Iwai, Y.; Hirano, A.; Nakagawa, A.; Awaya, J.;
hydrochloride for the treatment of relapsing remitting multiple Tsuchiya, H.; Takahashi, Y.; Asuma, R. A new alkaloid AM-2282 OF
sclerosis. Expert Opin. Pharmacother. 2017, 18, 1649−1660. Streptomyces origin. Taxonomy, fermentation, isolation and prelimi-
(70) Yoshida, M.; Horinouchi, S.; Beppu, T. Trichostatin A and nary characterization. J. Antibiot. 1977, 30, 275−282.
trapoxin: novel chemical probes for the role of histone acetylation in (86) Hug, H.; Sarre, T. F. Protein kinase C isoenzymes: divergence
chromatin structure and function. BioEssays 1995, 17, 423−430. in signal transduction? Biochem. J. 1993, 291, 329−343.
(71) Stowell, J. C.; Huot, R. I.; Van Voast, L. The synthesis of N- (87) Jirousek, M. R.; Goekjian, P. G. Protein kinase C inhibitors as
hydroxy-N’-phenyloctanediamide and its inhibitory effect on pro- novel anticancer drugs. Expert Opin. Invest. Drugs 2001, 10, 2117−
liferation of AXC rat prostate cancer cells. J. Med. Chem. 1995, 38, 2140.
1411−1413. (88) van Gijn, R.; Zuidema, X.; Bult, A.; Beijnen, J. H. Protein kinase
(72) Huang, Y.; Dong, G.; Li, H.; Liu, N.; Zhang, W.; Sheng, C. C as a target for new anti-cancer agents. J. Oncol. Pharm. Pract. 1999,
Discovery of Janus Kinase 2 (JAK2) and Histone Deacetylase 5, 166−178.
(HDAC) Dual Inhibitors as a Novel Strategy for the Combinational (89) Jirousek, M. R.; Gillig, J. R.; Gonzalez, C. M.; Heath, W. F.;
Treatment of Leukemia and Invasive Fungal Infections. J. Med. Chem.
McDonald, J. H., 3rd; Neel, D. A.; Rito, C. J.; Singh, U.; Stramm, L.
2018, 61, 6056−6074.
E.; Melikian-Badalian, A.; Baevsky, M.; Ballas, L. M.; Hall, S. E.;
(73) Finnin, M. S.; Donigian, J. R.; Cohen, A.; Richon, V. M.;
Winneroski, L. L.; Faul, M. M. (S)-13-[(dimethylamino)methyl]-
Rifkind, R. A.; Marks, P. A.; Breslow, R.; Pavletich, N. P. Structures of
10,11,14,15-tetrahydro-4,9:16, 21-dimetheno-1H, 13H-dibenzo[e,k]-
a histone deacetylase homologue bound to the TSA and SAHA
pyrrolo[3,4-h][1,4,13]oxadiazacyclohexadecene-1,3(2H)-d ione
inhibitors. Nature 1999, 401, 188−193.
(74) Mann, B. S.; Johnson, J. R.; Cohen, M. H.; Justice, R.; Pazdur, (LY333531) and related analogues: isozyme selective inhibitors of
R. FDA approval summary: vorinostat for treatment of advanced protein kinase C beta. J. Med. Chem. 1996, 39, 2664−2671.
primary cutaneous T-cell lymphoma. Oncologist 2007, 12, 1247− (90) https://clinicaltrials.gov/, clinical trial identifier:
1252. NCT02769611.
(75) Uemura, D.; Takahashi, K.; Yamamoto, T.; Katayama, C.; (91) Faul, M. M.; Gillig, J. R.; Jirousek, M. R.; Ballas, L. M.;
Tanaka, J.; Okumura, Y.; Hirata, Y. Norhalichondrin A: an antitumor Schotten, T.; Kahl, A.; Mohr, M. Acyclic N-(azacycloalkyl)-
polyether macrolide from a marine sponge. J. Am. Chem. Soc. 1985, bisindolylmaleimides: isozyme selective inhibitors of PKCbeta. Bioorg.
107, 4796−4798. Med. Chem. Lett. 2003, 13, 1857−1859.
(76) Hirata, Y.; Uemura, D. Halichondrins - antitumor polyether (92) Crump, M.; Leppa, S.; Fayad, L.; Lee, J. J.; Di Rocco, A.;
macrolides from a marine sponge. Pure Appl. Chem. 1986, 58, 701− Ogura, M.; Hagberg, H.; Schnell, F.; Rifkin, R.; Mackensen, A.;
710. Offner, F.; Pinter-Brown, L.; Smith, S.; Tobinai, K.; Yeh, S. P.; Hsi, E.
(77) Bai, R. L.; Paull, K. D.; Herald, C. L.; Malspeis, L.; Pettit, G. R.; D.; Nguyen, T.; Shi, P.; Hahka-Kemppinen, M.; Thornton, D.; Lin,
Hamel, E. Halichondrin B and homohalichondrin B, marine natural B.; Kahl, B.; Schmitz, N.; Savage, K. J.; Habermann, T. Randomized,
products binding in the vinca domain of tubulin. Discovery of tubulin- Double-Blind, Phase III Trial of Enzastaurin Versus Placebo in
based mechanism of action by analysis of differential cytotoxicity data. Patients Achieving Remission After First-Line Therapy for High-Risk
J. Biol. Chem. 1991, 266, 15882−15889. Diffuse Large B-Cell Lymphoma. J. Clin. Oncol. 2016, 34, 2484−2492.
(78) Aicher, T. D.; Buszek, K. R.; Fang, F. G.; Forsyth, C. J.; Jung, S. (93) https://clinicaltrials.gov/, clinical trial identifier:
H.; Kishi, Y.; Matelich, M. C.; Scola, P. M.; Spero, D. M.; Yoon, S. K. NCT03263026.
Total synthesis of halichondrin B and norhalichondrin B. J. Am. Chem. (94) Fiorucci, G.; Percario, Z. A.; Marcolin, C.; Coccia, E. M.;
Soc. 1992, 114, 3162−3164. Affabris, E.; Romeo, G. Inhibition of protein phosphorylation
(79) Jackson, K. L.; Henderson, J. A.; Phillips, A. J. The modulates expression of the Jak family protein tyrosine kinases. J.
halichondrins and E7389. Chem. Rev. 2009, 109, 3044−3079. Virol. 1995, 69, 5833−5837.
(80) Zheng, W.; Seletsky, B. M.; Palme, M. H.; Lydon, P. J.; Singer, (95) Yang, S. M.; Malaviya, R.; Wilson, L. J.; Argentieri, R.; Chen,
L. A.; Chase, C. E.; Lemelin, C. A.; Shen, Y.; Davis, H.; Tremblay, L.; X.; Yang, C.; Wang, B.; Cavender, D.; Murray, W. V. Simplified
Towle, M. J.; Salvato, K. A.; Wels, B. F.; Aalfs, K. K.; Kishi, Y.; staurosporine analogs as potent JAK3 inhibitors. Bioorg. Med. Chem.
Littlefield, B. A.; Yu, M. J. Macrocyclic ketone analogues of Lett. 2007, 17, 326−331.
halichondrin B. Bioorg. Med. Chem. Lett. 2004, 14, 5551−5554. (96) Chang, R.; Lotti, V.; Monaghan, R.; Birnbaum, J.; Stapley, E.;
(81) Towle, M. J.; Salvato, K. A.; Budrow, J.; Wels, B. F.; Kuznetsov, Goetz, M.; Albers-Schonberg, G.; Patchett, A.; Liesch, J.; Hensens, O.;
G.; Aalfs, K. K.; Welsh, S.; Zheng, W.; Seletsky, B. M.; Palme, M. H.; et al. A potent nonpeptide cholecystokinin antagonist selective for
Habgood, G. J.; Singer, L. A.; Dipietro, L. V.; Wang, Y.; Chen, J. J.; peripheral tissues isolated from Aspergillus alliaceus. Science 1985,
Quincy, D. A.; Davis, A.; Yoshimatsu, K.; Kishi, Y.; Yu, M. J.; 230, 177−179.

AH DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(97) Zucker, K. A.; Adrian, T. E.; Zdon, M. J.; Ballantyne, G. H.; (114) Rinehart, K. L.; Holt, T. G.; Fregeau, N. L.; Stroh, J. G.;
Modlin, I. M. Asperlicin: a unique nonpeptide cholecystokinin Keifer, P. A.; Sun, F.; Li, L. H.; Martin, D. G. Ecteinascidins 729, 743,
antagonist. Surgery 1987, 102, 163−170. 745, 759A, 759B, and 770: potent antitumor agents from the
(98) Lotti, V. J.; Cerino, D. J.; Kling, P. J.; Chang, R. S. A new simple Caribbean tunicate Ecteinascidin turbinata. J. Org. Chem. 1990, 55,
mouse model for the in vivo evaluation of cholecystokinin (CCK) 4512−4515.
antagonists: comparative potencies and durations of action of (115) Abraham, I.; El Sayed, K.; Chen, Z. S.; Guo, H. Current status
nonpeptide antagonists. Life Sci. 1986, 39, 1631−1638. on marine products with reversal effect on cancer multidrug
(99) Bock, M. G.; DiPardo, R. M.; Rittle, K. E.; Evans, B. E.; resistance. Mar. Drugs 2012, 10, 2312−2321.
Freidinger, R. M.; Veber, D. F.; Chang, R. S.; Chen, T. B.; Keegan, M. (116) Martinez, E. J.; Owa, T.; Schreiber, S. L.; Corey, E. J.
E.; Lotti, V. J. Cholecystokinin antagonists. Synthesis of asperlicin Phthalascidin, a synthetic antitumor agent with potency and mode of
analogues with improved potency and water solubility. J. Med. Chem. action comparable to ecteinascidin 743. Proc. Natl. Acad. Sci. U. S. A.
1986, 29, 1941−1945. 1999, 96, 3496−3501.
(100) Evans, B. E.; Bock, M. G.; Rittle, K. E.; DiPardo, R. M.; (117) Iturriagavásquez, P.; Carbone, A.; Garcíabeltrán, O.;
Whitter, W. L.; Veber, D. F.; Anderson, P. S.; Freidinger, R. M. Livingstone, P. D.; Biggin, P. C.; Cassels, B. K.; Wonnacott, S.;
Design of potent, orally effective, nonpeptidal antagonists of the Zapatatorres, G.; Bermudez, I. Molecular determinants for com-
peptide hormone cholecystokinin. Proc. Natl. Acad. Sci. U. S. A. 1986, petitive inhibition of alpha4beta2 nicotinic acetylcholine receptors.
83, 4918−4922. Mol. Pharmacol. 2010, 78, 366−375.
(101) Huang, X. P.; Karpiak, J.; Kroeze, W. K.; Zhu, H.; Chen, X.; (118) Crestey, F.; Jensen, A. A.; Borch, M.; Andreasen, J. T.;
Moy, S. S.; Saddoris, K. A.; Nikolova, V. D.; Farrell, M. S.; Wang, S.; Andersen, J.; Balle, T.; Kristensen, J. L. Design, synthesis, and
et al. Allosteric ligands for the pharmacologically dark receptors biological evaluation of erythrina alkaloid analogues as neuronal
GPR68 and GPR65. Nature 2015, 527, 477−483. nicotinic acetylcholine receptor antagonists. J. Med. Chem. 2013, 56,
(102) Spencer, J.; Rathnam, R. P.; Chowdhry, B. Z. 1,4- 9673−9682.
Benzodiazepin-2-ones in medicinal chemistry. Future Med. Chem. (119) Rao, J. U. M.; Giri, G. S.; Hanumaiah, T.; Rao, K. V. J.
2010, 2, 1441−1449. Sampangine, a New Alkaloid from Cananga odorata. J. Nat. Prod.
(103) Gao, F.; Sexton, P. M.; Christopoulos, A.; Miller, L. J. 1986, 49, 346−347.
Benzodiazepine ligands can act as allosteric modulators of the Type 1 (120) Peterson, J. R.; Zjawiony, J. K.; Liu, S.; Hufford, C. D.; Clark,
cholecystokinin receptor. Bioorg. Med. Chem. Lett. 2008, 18, 4401− A. M.; Rogers, R. D. Copyrine alkaloids: synthesis, spectroscopic
4404. characterization, and antimycotic/antimycobacterial activity of A- and
(104) Evans, B. E.; Rittle, K. E.; Bock, M. G.; DiPardo, R. M.; B-ring-functionalized sampangines. J. Med. Chem. 1992, 35, 4069−
Freidinger, R. M.; Whitter, W. L.; Gould, N. P.; Lundell, G. F.; 4077.
(121) Jiang, Z.; Liu, N.; Dong, G.; Jiang, Y.; Liu, Y.; He, X.; Huang,
Homnick, C. F. Design of nonpeptidal ligands for a peptide receptor:
Y.; He, S.; Chen, W.; Li, Z.; Yao, J.; Miao, Z.; Zhang, W.; Sheng, C.
cholecystokinin antagonists. J. Med. Chem. 1987, 30, 1229−1239.
Scaffold hopping of sampangine: discovery of potent antifungal lead
(105) Evans, B. E.; Rittle, K. E.; Bock, M. G.; DiPardo, R. M.;
compound against Aspergillus fumigatus and Cryptococcus neofor-
Freidinger, R. M.; Whitter, W. L.; Lundell, G. F.; Veber, D. F.;
mans. Bioorg. Med. Chem. Lett. 2014, 24, 4090−4094.
Anderson, P. S.; Chang, R. S.; et al. Methods for drug discovery:
(122) Jiang, Z.; Liu, N.; Hu, D.; Dong, G.; Miao, Z.; Yao, J.; He, H.;
development of potent, selective, orally effective cholecystokinin
Jiang, Y.; Zhang, W.; Wang, Y.; Sheng, C. The discovery of novel
antagonists. J. Med. Chem. 1988, 31, 2235−2246.
antifungal scaffolds by structural simplification of the natural product
(106) Abbruzzese, J. L.; Gholson, C. F.; Daugherty, K.; Larson, E.;
sampangine. Chem. Commun. 2015, 51, 14648−14651.
DuBrow, R.; Berlin, R.; Levin, B. A pilot clinical trial of the (123) Liu, N.; Zhong, H.; Tu, J.; Jiang, Z.; Jiang, Y.; Li, J.; Zhang,
cholecystokinin receptor antagonist MK-329 in patients with W.; Wang, Y.; Sheng, C.; Jiang, Y.; Jiang, Y. Discovery of simplified
advanced pancreatic cancer. Pancreas 1992, 7, 165−171. sampangine derivatives as novel fungal biofilm inhibitors. Eur. J. Med.
(107) Su, Y.; Ge, J.; Zhu, B.; Zheng, Y. G.; Zhu, Q.; Yao, S. Q. Chem. 2018, 143, 1510−1523.
Target identification of biologically active small molecules via in situ (124) Han, Q. B.; Xu, H. X. Caged Garcinia xanthones: development
methods. Curr. Opin. Chem. Biol. 2013, 17, 768−775. since 1937. Curr. Med. Chem. 2009, 16, 3775−3796.
(108) Weber, L. The application of multi-component reactions in (125) Wang, X.; Lu, N.; Yang, Q.; Gong, D.; Lin, C.; Zhang, S.; Xi,
drug discovery. Curr. Med. Chem. 2002, 9, 2085−2093. M.; Gao, Y.; Wei, L.; Guo, Q.; You, Q. Studies on chemical
(109) Magedov, I. V.; Manpadi, M.; Ogasawara, M. A.; Dhawan, A. modification and biology of a natural product, gambogic acid (III):
S.; Rogelj, S.; Van Slambrouck, S.; Steelant, W. F.; Evdokimov, N. M.; determination of the essential pharmacophore for biological activity.
Uglinskii, P. Y.; Elias, E. M.; Knee, E. J.; Tongwa, P.; Antipin, M. Y.; Eur. J. Med. Chem. 2011, 46, 1280−1290.
Kornienko, A. Structural simplification of bioactive natural products (126) Chi, Y.; Zhan, X. K.; Yu, H.; Xie, G. R.; Wang, Z. Z.; Xiao, W.;
with multicomponent synthesis. 2. antiproliferative and antitubulin Wang, Y. G.; Xiong, F. X.; Hu, J. F.; Yang, L.; Cui, C. X.; Wang, J. W.
activities of pyrano[3,2-c]pyridones and pyrano[3,2-c]quinolones. J. An open-labeled, randomized, multicenter phase IIa study of
Med. Chem. 2008, 51, 2561−2570. gambogic acid injection for advanced malignant tumors. Chin. Med.
(110) Evdokimov, N. M.; Van Slambrouck, S.; Heffeter, P.; Tu, L.; J. 2013, 126, 1642−1646.
Le Calve, B.; Lamoral-Theys, D.; Hooten, C. J.; Uglinskii, P. Y.; (127) Sun, H.; Chen, F.; Wang, X.; Liu, Z.; Yang, Q.; Zhang, X.;
Rogelj, S.; Kiss, R.; Steelant, W. F.; Berger, W.; Yang, J. J.; Bologa, C. Zhu, J.; Qiang, L.; Guo, Q.; You, Q. Studies on gambogic acid (IV):
G.; Kornienko, A.; Magedov, I. V. Structural simplification of Exploring structure−activity relationship with IκB kinase-beta
bioactive natural products with multicomponent synthesis. 3. Fused (IKKβ). Eur. J. Med. Chem. 2012, 51, 110−123.
uracil-containing heterocycles as novel topoisomerase-targeting (128) Liu, J.; Ma, R.; Bi, F.; Zhang, F.; Hu, C.; Venter, H.; Semple,
agents. J. Med. Chem. 2011, 54, 2012−2021. S. J.; Ma, S. Novel 5-methyl-2-phenylphenanthridium derivatives as
(111) Wender, P. A.; Verma, V. A.; Paxton, T. J.; Pillow, T. H. FtsZ-targeting antibacterial agents from structural simplification of
Function-oriented synthesis, step economy, and drug design. Acc. natural product sanguinarine. Bioorg. Med. Chem. Lett. 2018, 28,
Chem. Res. 2008, 41, 40−49. 1825−1831.
(112) Szpilman, A. M.; Carreira, E. M. Probing the biology of (129) Lock, R. L.; Harry, E. J. Cell-division inhibitors: new insights
natural products: molecular editing by diverted total synthesis. Angew. for future antibiotics. Nat. Rev. Drug Discovery 2008, 7, 324−338.
Chem., Int. Ed. 2010, 49, 9592−9628. (130) Beuria, T. K.; Santra, M. K.; Panda, D. Sanguinarine blocks
(113) Wetzel, S.; Bon, R. S.; Kumar, K.; Waldmann, H. Biology- cytokinesis in bacteria by inhibiting FtsZ assembly and bundling.
oriented synthesis. Angew. Chem., Int. Ed. 2011, 50, 10800−10826. Biochemistry 2005, 44, 16584−16593.

AI DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(131) Konishi, M.; Ohkuma, H.; Matsumoto, K.; Tsuno, T.; Kamei, as a target of the antitumor natural product pladienolide. Nat. Chem.
H.; Miyaki, T.; Oki, T.; Kawaguchi, H.; VanDuyne, G. D.; Clardy, J. Biol. 2007, 3, 570−575.
Dynemicin A, a novel antibiotic with the anthraquinone and 1,5-diyn- (151) Müller, S.; Mayer, T.; Sasse, F.; Maier, M. E. Synthesis of a
3-ene subunit. J. Antibiot. 1989, 42, 1449−1452. Pladienolide B Analogue with the Fully Functionalized Core
(132) Lynch, V. M.; Fairhurst, R. A.; Iliadis, T. N.; Magnus, P.; Structure. Org. Lett. 2011, 13, 3940−3943.
Davis, B. E. Two [7.3.1]Azabicyclo-z-3-ene-1,5-diyne Analogues of (152) Gundluru, M. K.; Pourpak, A.; Cui, X.; Morris, S. W.; Webb,
Dynemicin A. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1995, T. R. Design, synthesis and initial biological evaluation of a novel
51, 782−786. pladienolide analog scaffold. MedChemComm 2011, 2, 904−908.
(133) Nicolaou, K. C.; Smith, A. L.; Yue, E. W. Chemistry and (153) Kumar, V. P.; Chandrasekhar, S. Enantioselective Synthesis of
biology of natural and designed enediynes. Proc. Natl. Acad. Sci. U. S. Pladienolide B and Truncated Analogues as New Anticancer Agents.
A. 1993, 90, 5881−5888. Org. Lett. 2013, 15, 3610−3613.
(134) Nicolaou, K. C.; Dai, W. M.; Tsay, S. C.; Estevez, V. A.; (154) Villa, R.; Kashyap, M. K.; Kumar, D.; Kipps, T. J.; Castro, J. E.;
Wrasidlo, W. Designed enediynes: a new class of DNA-cleaving La Clair, J. J.; Burkart, M. D. Stabilized cyclopropane analogs of the
molecules with potent and selective anticancer activity. Science 1992, splicing inhibitor FD-895. J. Med. Chem. 2013, 56, 6576−6582.
256, 1172−1178. (155) Arai, K.; Buonamici, S.; Chan, B.; Corson, L.; Endo, A.;
(135) Wender, P. A.; Zercher, C. K. Studies on DNA-cleaving Gerard, B.; Hao, M. H.; Karr, C.; Kira, K.; Lee, L.; Liu, X.; Lowe, J. T.;
agents: synthesis of a functional dynemicin analogue. J. Am. Chem. Luo, T.; Marcaurelle, L. A.; Mizui, Y.; Nevalainen, M.; O’Shea, M. W.;
Soc. 1991, 113, 2311−2313. Park, E. S.; Perino, S. A.; Prajapati, S.; Shan, M.; Smith, P. G.;
(136) Nicolaou, K. C.; Smith, A. L. Molecular design, chemical Tivitmahaisoon, P.; Wang, J. Y.; Warmuth, M.; Wu, K. M.; Yu, L.;
synthesis, and biological action of enediynes. Acc. Chem. Res. 1992, 25, Zhang, H.; Zheng, G. Z.; Keaney, G. F. Total synthesis of 6-
497−503. deoxypladienolide D and Assessment of Splicing Inhibitory Activity in
(137) Huang, J.-m.; Yokoyama, R.; Yang, C.-s.; Fukuyama, Y. a Mutant SF3B1 cancer cell line. Org. Lett. 2014, 16, 5560−5563.
Merrilactone A, a novel neurotrophic sesquiterpene dilactone from (156) Lagisetti, C.; Yermolina, M. V.; Sharma, L. K.; Palacios, G.;
Illicium merrillianum. Tetrahedron Lett. 2000, 41, 6111−6114. Prigaro, B. J.; Webb, T. R. Pre-mRNA splicing-modulatory
(138) Kouno, I.; Mori, K.; Kawano, N.; Sato, S. Structure of pharmacophores: the total synthesis of herboxidiene, a pladienolide-
anislactone A; a new skeletal type of sesquiterpene from the pericarps herboxidiene hybrid analog and related derivatives. ACS Chem. Biol.
of Illicium anisatum. Tetrahedron Lett. 1989, 30, 7451−7452. 2014, 9, 643−648.
(139) Shi, L.; Meyer, K.; Greaney, M. F. Synthesis of (157) Mcbrien, K. D.; Gao, Q.; Huang, S.; Klohr, S. E.; Wang, R. R.;
(±)-merrilactone A and (±)-anislactone A. Angew. Chem., Int. Ed. Pirnik, D. M.; Neddermann, K. M.; Bursuker, I.; Kadow, K. F.; Leet, J.
2010, 49, 9250−9253. E. Fusaricide, a new cytotoxic N-hydroxypyridone from Fusarium sp.
(140) Richers, J.; Pothig, A.; Herdtweck, E.; Sippel, C.; Hausch, F.; J. Nat. Prod. 1996, 59, 1151−1153.
(158) Kamperdick, C.; Van, N. H.; Sung, T. V.; Adam, G.
Tiefenbacher, K. Synthesis and Neurotrophic Activity Studies of
Bisquinolinone alkaloids from Melicope ptelefolia. Phytochemistry
Illicium Sesquiterpene Natural Product Analogues. Chem. - Eur. J.
1999, 50, 177−181.
2017, 23, 3178−3183.
(159) Chen, I. S.; Wu, S. J.; Tsai, I. L. Chemical and bioactive
(141) Chen, J.; Gao, P.; Yu, F.; Yang, Y.; Zhu, S.; Zhai, H. Total
constituents from Zanthoxylum simulans. J. Nat. Prod. 1994, 57,
synthesis of (±)-merrilactone A. Angew. Chem., Int. Ed. 2012, 51,
1206−1211.
5897−5899.
(160) Hsiang, Y. H.; Hertzberg, R.; Hecht, S.; Liu, L. F.
(142) Birman, V. B.; Danishefsky, S. J. The total synthesis of
Camptothecin induces protein-linked DNA breaks via mammalian
(±)-merrilactone A. J. Am. Chem. Soc. 2002, 124, 2080−2081.
DNA topoisomerase I. J. Biol. Chem. 1985, 260, 14873−14878.
(143) Omura, S.; Tomoda, H.; Kim, Y. K.; Nishida, H.
(161) Pommier, Y. DNA topoisomerase I inhibitors: chemistry,
Pyripyropenes, highly potent inhibitors of acyl-CoA:cholesterol biology, and interfacial inhibition. Chem. Rev. 2009, 109, 2894−2902.
acyltransferase produced by Aspergillus fumigatus. J. Antibiot. 1993, (162) Xu, Y.; Her, C. Inhibition of Topoisomerase (DNA) I
46, 1168−1169. (TOP1): DNA Damage Repair and Anticancer Therapy. Biomolecules
(144) Lada, A. T.; Davis, M.; Kent, C.; Chapman, J.; Tomoda, H.; 2015, 5, 1652−1670.
Omura, S.; Rudel, L. L. Identification of ACAT1- and ACAT2-specific (163) Pommier, Y. Topoisomerase I inhibitors: camptothecins and
inhibitors using a novel, cell-based fluorescence assay: individual beyond. Nat. Rev. Cancer 2006, 6, 789−802.
ACAT uniqueness. J. Lipid Res. 2004, 45, 378−386. (164) Tam, S. W.; Worcel, M.; Wyllie, M. Yohimbine: a clinical
(145) Ohshiro, T.; Rudel, L. L.; Omura, S.; Tomoda, H. Selectivity review. Pharmacol. Ther. 2001, 91, 215−243.
of microbial acyl-CoA: cholesterol acyltransferase inhibitors toward (165) Koch, M. A.; Schuffenhauer, A.; Scheck, M.; Wetzel, S.;
isozymes. J. Antibiot. 2007, 60, 43−51. Casaulta, M.; Odermatt, A.; Ertl, P.; Waldmann, H. Charting
(146) Roth, B. D. ACAT inhibitors: evolution from cholesterol- biologically relevant chemical space: a structural classification of
absorption inhibitors to antiatherosclerotic agents. Drug Discovery natural products (SCONP). Proc. Natl. Acad. Sci. U. S. A. 2005, 102,
Today 1998, 3, 19−25. 17272−17277.
(147) Ohtawa, M.; Arima, S.; Ichida, N.; Terayama, T.; Ohno, H.; (166) Schuffenhauer, A.; Ertl, P.; Roggo, S.; Wetzel, S.; Koch, M. A.;
Yamazaki, T.; Ohshiro, T.; Sato, N.; Omura, S.; Tomoda, H.; Waldmann, H. The scaffold tree–visualization of the scaffold universe
Nagamitsu, T. Design and Synthesis of A-Ring Simplified by hierarchical scaffold classification. J. Chem. Inf. Model. 2007, 47,
Pyripyropene A Analogues as Potent and Selective Synthetic 47−58.
SOAT2 Inhibitors. ChemMedChem 2018, 13, 411−421. (167) Correa, I. R., Jr.; Noren-Muller, A.; Ambrosi, H. D.; Jakupovic,
(148) Odani, A.; Ishihara, K.; Ohtawa, M.; Tomoda, H.; Omura, S.; S.; Saxena, K.; Schwalbe, H.; Kaiser, M.; Waldmann, H. Identification
Nagamitsu, T. Total synthesis of pyripyropene A. Tetrahedron 2011, of inhibitors for mycobacterial protein tyrosine phosphatase B
67, 8195−8203. (MptpB) by biology-oriented synthesis (BIOS). Chem. - Asian J.
(149) Mizui, Y.; Sakai, T.; Iwata, M.; Uenaka, T.; Okamoto, K.; 2007, 2, 1109−1126.
Shimizu, H.; Yamori, T.; Yoshimatsu, K.; Asada, M.; Uenaka, T. (168) Wehner, F.; Noren-Muller, A.; Muller, O.; Reis-Correa, I., Jr.;
Pladienolides, new substances from culture of Streptomyces platensis Giannis, A.; Waldmann, H. Indoloquinolizidine derivatives as novel
Mer-11107. III. In vitro and in vivo antitumor activities. J. Antibiot. and potent apoptosis inducers and cell-cycle blockers. ChemBioChem
2004, 57, 188−196. 2008, 9, 401−405.
(150) Kotake, Y.; Sagane, K.; Owa, T.; Mimori-Kiyosue, Y.; Shimizu, (169) Rosenbaum, C.; Baumhof, P.; Mazitschek, R.; Muller, O.;
H.; Uesugi, M.; Ishihama, Y.; Iwata, M.; Mizui, Y. Splicing factor SF3b Giannis, A.; Waldmann, H. Synthesis and biological evaluation of an

AJ DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

indomethacin library reveals a new class of angiogenesis-related kinase simplified taxanes with remarkable p-glycoprotein inhibitory activity.
inhibitors. Angew. Chem., Int. Ed. 2004, 43, 224−228. ACS Med. Chem. Lett. 2010, 1, 416−421.
(170) Chen, L. X.; He, H.; Qiu, F. Natural withanolides: an (187) Avramova, S. I.; Galletti, E.; Renzulli, M. L.; Giorgi, G.;
overview. Nat. Prod. Rep. 2011, 28, 705−740. Sgaragli, G.; Alderighi, D.; Ghiron, C.; Corelli, F.; Radi, M.; Botta, M.
(171) van Hattum, H.; Waldmann, H. Biology-oriented synthesis: Synthesis of an original oxygenated taxuspine X analogue: a versatile
harnessing the power of evolution. J. Am. Chem. Soc. 2014, 136, ″non-natural″ natural product with remarkable P-gp modulating
11853−11859. activity. ChemMedChem 2008, 3, 745−748.
(172) Svenda, J.; Sheremet, M.; Kremer, L.; Maier, L.; Bauer, J. O.; (188) Aoki, S.; Watanabe, Y.; Sanagawa, M.; Setiawan, A.; Kotoku,
Strohmann, C.; Ziegler, S.; Kumar, K.; Waldmann, H. Biology- N.; Kobayashi, M. Cortistatins A, B, C, and D, anti-angiogenic
oriented synthesis of a withanolide-inspired compound collection steroidal alkaloids, from the marine sponge Corticium simplex. J. Am.
reveals novel modulators of hedgehog signaling. Angew. Chem., Int. Ed. Chem. Soc. 2006, 128, 3148−3149.
2015, 54, 5596−5602. (189) Kotoku, N.; Sumii, Y.; Hayashi, T.; Tamura, S.; Kawachi, T.;
(173) Sheremet, M.; Kapoor, S.; Schroder, P.; Kumar, K.; Ziegler, S.; Shiomura, S.; Arai, M.; Kobayashi, M. Creation of readily accessible
Waldmann, H. Small Molecules Inspired by the Natural Product and orally active analogue of cortistatin a. ACS Med. Chem. Lett. 2012,
Withanolides as Potent Inhibitors of Wnt Signaling. ChemBioChem 3, 673−677.
2017, 18, 1797−1806. (190) Kotoku, N.; Ito, A.; Shibuya, S.; Mizuno, K.; Takeshima, A.;
(174) Giri, A.; Narasu, M. L. Production of podophyllotoxin from Nogata, M.; Kobayashi, M. Short-step synthesis and structure-activity
Podophyllum hexandrum: a potential natural product for clinically relationship of cortistatin A analogs. Tetrahedron 2017, 73, 1342−
useful anticancer drugs. Cytotechnology 2000, 34, 17−26. 1349.
(175) Rönquist-Nii, Y.; Eksborg, S.; Axelson, M.; Harmenberg, J.; (191) Staveness, D.; Abdelnabi, R.; Near, K. E.; Nakagawa, Y.;
Beck, O. Determination of picropodophyllin and its isomer Neyts, J.; Delang, L.; Leyssen, P.; Wender, P. A. Inhibition of
podophyllotoxin in human serum samples with electrospray ionization Chikungunya Virus-Induced Cell Death by Salicylate-Derived
of hexylamine adducts by liquid chromatography−tandem mass Bryostatin Analogues Provides Additional Evidence for a PKC-
spectrometry. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2011, Independent Pathway. J. Nat. Prod. 2016, 79, 680−684.
879, 326−334. (192) Bourjot, M.; Delang, L.; Nguyen, V. H.; Neyts, J.; Gueritte, F.;
(176) Hitotsuyanagi, Y.; Fukuyo, M.; Tsuda, K.; Kobayashi, M.; Leyssen, P.; Litaudon, M. Prostratin and 12-O-tetradecanoylphorbol
Ozeki, A.; Itokawa, H.; Takeya, K. 4-Aza-2,3-dehydro-4-deoxypodo- 13-acetate are potent and selective inhibitors of Chikungunya virus
phyllotoxins: simple aza-podophyllotoxin analogues possessing potent replication. J. Nat. Prod. 2012, 75, 2183−2187.
cytotoxicity. Bioorg. Med. Chem. Lett. 2000, 10, 315−317. (193) Nothias-Scaglia, L. F.; Pannecouque, C.; Renucci, F.; Delang,
(177) Magedov, I. V.; Frolova, L.; Manpadi, M.; Bhoga, U.; Tang, L.; Neyts, J.; Roussi, F.; Costa, J.; Leyssen, P.; Litaudon, M.; Paolini, J.
H.; Evdokimov, N. M.; George, O.; Georgiou, K. H.; Renner, S.; Antiviral Activity of Diterpene Esters on Chikungunya Virus and HIV
Getlik, M.; Kinnibrugh, T. L.; Fernandes, M. A.; Van slambrouck, S.; Replication. J. Nat. Prod. 2015, 78, 1277−1283.
Steelant, W. F.; Shuster, C. B.; Rogelj, S.; van Otterlo, W. A.; (194) Gupta, D. K.; Kaur, P.; Leong, S. T.; Tan, L. T.; Prinsep, M.
Kornienko, A. Anticancer properties of an important drug lead R.; Chu, J. J. Anti-Chikungunya viral activities of aplysiatoxin-related
podophyllotoxin can be efficiently mimicked by diverse heterocyclic compounds from the marine cyanobacterium Trichodesmium
scaffolds accessible via one-step synthesis. J. Med. Chem. 2011, 54, erythraeum. Mar. Drugs 2014, 12, 115−127.
4234−4246. (195) Pettit, G. R.; Herald, C. L.; Doubek, D. L.; Herald, D. L.;
(178) Ravelli, R. B.; Gigant, B.; Curmi, P. A.; Jourdain, I.; Lachkar, Arnold, E.; Clardy, J. Isolation and structure of bryostatin 1. J. Am.
S.; Sobel, A.; Knossow, M. Insight into tubulin regulation from a Chem. Soc. 1982, 104, 6846−6848.
complex with colchicine and a stathmin-like domain. Nature 2004, (196) Keck, G. E.; Poudel, Y. B.; Cummins, T. J.; Rudra, A.; Covel, J.
428, 198−202. A. Total synthesis of bryostatin 1. J. Am. Chem. Soc. 2011, 133, 744−
(179) Galletti, E.; Magnani, M.; Renzulli, M. L.; Botta, M. Paclitaxel 747.
and docetaxel resistance: molecular mechanisms and development of (197) Wender, P. A.; Nakagawa, Y.; Near, K. E.; Staveness, D.
new generation taxanes. ChemMedChem 2007, 2, 920−942. Computer-guided design, synthesis, and protein kinase C affinity of a
(180) Teodori, E.; Dei, S.; Martelli, C.; Scapecchi, S.; Gualtieri, F. new salicylate-based class of bryostatin analogs. Org. Lett. 2014, 16,
The functions and structure of ABC transporters: implications for the 5136−5139.
design of new inhibitors of Pgp and MRP1 to control multidrug (198) Wender, P. A.; Staveness, D. Improved protein kinase C
resistance (MDR). Curr. Drug Targets 2006, 7, 893−909. affinity through final step diversification of a simplified salicylate-
(181) Leslie, E. M.; Deeley, R. G.; Cole, S. P. C. Multidrug derived bryostatin analog scaffold. Org. Lett. 2014, 16, 5140−5143.
resistance proteins: role of P-glycoprotein, MRP1, MRP2, and BCRP (199) Binns, W.; James, L. F.; Keeler, R. F.; Balls, L. D. Effects of
(ABCG2) in tissue defense. Toxicol. Appl. Pharmacol. 2005, 204, teratogenic agents in range plants. Cancer Res. 1968, 28, 2323−2326.
216−237. (200) Keeler, R. F.; Binns, W. Teratogenic compounds of Veratrum
(182) Perez-Tomas, R. Multidrug resistance: retrospect and californicum (Durand). II. Production of ovine fetal cyclopia by
prospects in anti-cancer drug treatment. Curr. Med. Chem. 2006, 13, fractions and alkaloid preparations. Can. J. Biochem. 1966, 44, 829−
1859−1876. 838.
(183) Zamir, L. O.; Zhou, Z. H.; Caron, G.; Nedea, M. E.; Sauriol, (201) Cooper, M. K.; Porter, J. A.; Young, K. E.; Beachy, P. A.
F.; Mamer, O. Isolation of a putative biogenetic taxane precursor from Teratogen-mediated inhibition of target tissue response to Shh
Taxus canadensis needles. J. Chem. Soc., Chem. Commun. 1995, 26, signaling. Science 1998, 280, 1603−1607.
529−530. (202) Chen, J. K.; Taipale, J.; Cooper, M. K.; Beachy, P. A.
(184) Begley, M. J.; Jackson, C. B.; Pattenden, G. Investigation of Inhibition of Hedgehog signaling by direct binding of cyclopamine to
transannular cyclisations of verticillanes to the taxane ring system. Smoothened. Genes Dev. 2002, 16, 2743−2748.
Tetrahedron Lett. 1985, 26, 3397−3400. (203) Berman, D. M.; Karhadkar, S. S.; Hallahan, A. R.; Pritchard, J.
(185) Hosoyama, H.; Inubushi, A.; Katsui, T.; Shigemori, H.; I.; Eberhart, C. G.; Watkins, D. N.; Chen, J. K.; Cooper, M. K.;
Kobayashi, J. i. Taxuspines U, V, and W, new taxane and related Taipale, J.; Olson, J. M.; Beachy, P. A. Medulloblastoma growth
diterpenoids from the Japanese yew Taxus cuspidata. Tetrahedron inhibition by hedgehog pathway blockade. Science 2002, 297, 1559−
1996, 52, 13145−13150. 1561.
(186) Castagnolo, D.; Contemori, L.; Maccari, G.; Avramova, S. I.; (204) Dahmane, N.; Sanchez, P.; Gitton, Y.; Palma, V.; Sun, T.;
Neri, A.; Sgaragli, G.; Botta, M. From taxuspine x to structurally Beyna, M.; Weiner, H.; Ruiz i Altaba, A. The Sonic Hedgehog-Gli

AK DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

pathway regulates dorsal brain growth and tumorigenesis. Development cyanobacterium Symploca sp. J. Am. Chem. Soc. 2008, 130, 1806−
2001, 128, 5201−5212. 1807.
(205) Sanchez, P.; Ruiz i Altaba, A. In vivo inhibition of endogenous (221) Cole, K. E.; Dowling, D. P.; Boone, M. A.; Phillips, A. J.;
brain tumors through systemic interference of Hedgehog signaling in Christianson, D. W. Structural basis of the antiproliferative activity of
mice. Mech. Dev. 2005, 122, 223−230. largazole, a depsipeptide inhibitor of the histone deacetylases. J. Am.
(206) Keeler, R. F. Teratogenic compounds in Veratrum Chem. Soc. 2011, 133, 12474−12477.
californicum (Durand) IX. Structure-activity relation. Teratology (222) Chen, F.; Chai, H.; Su, M. B.; Zhang, Y. M.; Li, J.; Xie, X.;
1970, 3, 169−173. Nan, F. J. Potent and orally efficacious bisthiazole-based histone
(207) Winkler, J. D.; Isaacs, A.; Holderbaum, L.; Tatard, V.; deacetylase inhibitors. ACS Med. Chem. Lett. 2014, 5, 628−633.
Dahmane, N. Design and synthesis of inhibitors of Hedgehog (223) McKee, T. C.; Galinis, D. L.; Pannell, L. K.; Cardellina, J. H.;
signaling based on the alkaloid cyclopamine. Org. Lett. 2009, 11, Laakso, J.; Ireland, C. M.; Murray, L.; Capon, R. J.; Boyd, M. R. The
2824−2827. Lobatamides, Novel Cytotoxic Macrolides from Southwestern Pacific
(208) Isaacs, A. K.; Xiang, C.; Baubet, V.; Dahmane, N.; Winkler, J. Tunicates. J. Org. Chem. 1998, 63, 7805−7810.
D. Studies directed toward the elucidation of the pharmacophore of (224) Nishi, T.; Forgac, M. The vacuolar (H+)-ATPases  nature’s
steroid-based Sonic Hedgehog signaling inhibitors. Org. Lett. 2011, most versatile proton pumps. Nat. Rev. Mol. Cell Biol. 2002, 3, 94−
13, 5140−5143. 103.
(209) Jansen, R.; Irschik, H.; Huch, V.; Schummer, D.; Steinmetz, (225) Changelian, P. S.; Flanagan, M. E.; Ball, D. J.; Kent, C. R.;
H.; Bock, M.; Schmidt, T.; Kirschning, A.; Müller, R. Carolacton − A Magnuson, K. S.; Martin, W. H.; Rizzuti, B. J.; Sawyer, P. S.; Perry, B.
Macrolide Ketocarbonic Acid that Reduces Biofilm Formation by the D.; Brissette, W. H.; McCurdy, S. P.; Kudlacz, E. M.; Conklyn, M. J.;
Caries- and Endocarditis-Associated Bacterium Streptococcus mutans. Elliott, E. A.; Koslov, E. R.; Fisher, M. B.; Strelevitz, T. J.; Yoon, K.;
Eur. J. Org. Chem. 2010, 2010, 1284−1289. Whipple, D. A.; Sun, J.; Munchhof, M. J.; Doty, J. L.; Casavant, J. M.;
(210) Kunze, B.; Reck, M.; Dötsch, A.; Lemme, A.; Schummer, D.; Blumenkopf, T. A.; Hines, M.; Brown, M. F.; Lillie, B. M.;
Irschik, H.; Steinmetz, H.; Wagner-Döbler, I. Damage of Strepto- Subramanyam, C.; Shang-Poa, C.; Milici, A. J.; Beckius, G. E.;
coccus mutans biofilms by carolacton, a secondary metabolite from Moyer, J. D.; Su, C.; Woodworth, T. G.; Gaweco, A. S.; Beals, C. R.;
the myxobacterium Sorangium cellulosum. BMC Microbiol. 2010, 10, Littman, B. H.; Fisher, D. A.; Smith, J. F.; Zagouras, P.; Magna, H. A.;
199. Saltarelli, M. J.; Johnson, K. S.; Nelms, L. F.; Des Etages, S. G.; Hayes,
(211) Reck, M.; Rutz, K.; Kunze, B.; Tomasch, J.; Surapaneni, S. K.; L. S.; Kawabata, T. T.; Finco-Kent, D.; Baker, D. L.; Larson, M.; Si,
Schulz, S.; Wagner-Dobler, I. The biofilm inhibitor carolacton M. S.; Paniagua, R.; Higgins, J.; Holm, B.; Reitz, B.; Zhou, Y. J.;
disturbs membrane integrity and cell division of Streptococcus Morris, R. E.; O’Shea, J. J.; Borie, D. C. Prevention of organ allograft
mutans through the serine/threonine protein kinase PknB. J. Bacteriol. rejection by a specific Janus kinase 3 inhibitor. Science 2003, 302,
2011, 193, 5692−5706. 875−878.
(212) Schmidt, T.; Kirschning, A. Total Synthesis of Carolacton, a (226) Boyd, M. R.; Farina, C.; Belfiore, P.; Gagliardi, S.; Kim, J. W.;
Hayakawa, Y.; Beutler, J. A.; McKee, T. C.; Bowman, B. J.; Bowman,
Highly Potent Biofilm Inhibitor. Angew. Chem., Int. Ed. 2012, 51,
E. J. Discovery of a novel antitumor benzolactone enamide class that
1063−1066.
selectively inhibits mammalian vacuolar-type (H+)-atpases. J.
(213) Stumpp, N.; Premnath, P.; Schmidt, T.; Ammermann, J.;
Pharmacol. Exp. Ther. 2001, 297, 114−120.
Dräger, G.; Reck, M.; Jansen, R.; Stiesch, M.; Wagner-Döbler, I.;
(227) Shen, R.; Lin, C. T.; Bowman, E. J.; Bowman, B. J.; Porco, J.
Kirschning, A. Synthesis of new carolacton derivatives and their
A., Jr. Synthesis and V-ATPase inhibition of simplified lobatamide
activity against biofilms of oral bacteria. Org. Biomol. Chem. 2015, 13,
analogues. Org. Lett. 2002, 4, 3103−3106.
5765−5774. (228) Shen, R.; Lin, C. T.; Bowman, E. J.; Bowman, B. J.; Porco, J.
(214) Solinski, A. E.; Koval, A. B.; Brzozowski, R. S.; Morrison, K. A. Total synthesis, stereochemical assignment, preparation of
R.; Fraboni, A. J.; Carson, C. E.; Eshraghi, A. R.; Zhou, G.; Quivey, R. simplified analogues, and V-ATPase inhibition studies. J. Am. Chem.
G.; Voelz, V. A.; Buttaro, B. A.; Wuest, W. M. Diverted Total Soc. 2003, 125, 7889−7901.
Synthesis of Carolacton-Inspired Analogs Yields Three Distinct (229) Fehr, T.; Kallen, J.; Oberer, L.; Sanglier, J. J.; Schilling, W.
Phenotypes in Streptococcus mutans Biofilms. J. Am. Chem. Soc. Sanglifehrins A, B, C and D, novel cyclophilin-binding compounds
2017, 139, 7188−7191. isolated from Streptomyces sp. A92−308110. II. Structure elucidation,
(215) Steele, A. D.; Knouse, K. W.; Keohane, C. E.; Wuest, W. M. stereochemistry and physico-chemical properties. J. Antibiot. 1999, 52,
Total synthesis and biological investigation of (−)-promysalin. J. Am. 474−479.
Chem. Soc. 2015, 137, 7314−7317. (230) Sanglier, J. J.; Quesniaux, V.; Fehr, T.; Hofmann, H.; Mahnke,
(216) Li, W.; Estrada-de los Santos, P.; Matthijs, S.; Xie, G. L.; M.; Memmert, K.; Schuler, W.; Zenke, G.; Gschwind, L.; Maurer, C.;
Busson, R.; Cornelis, P.; Rozenski, J.; De Mot, R. Promysalin, a Schilling, W. Sanglifehrins A, B, C and D, novel cyclophilin-binding
salicylate-containing Pseudomonas putida antibiotic, promotes surface compounds isolated from Streptomyces sp. A92−308110. I.
colonization and selectively targets other Pseudomonas. Chem. Biol. Taxonomy, fermentation, isolation and biological activity. J. Antibiot.
2011, 18, 1320−1330. 1999, 52, 466−473.
(217) Steele, A. D.; Keohane, C. E.; Knouse, K. W.; Rossiter, S. E.; (231) Zenke, G.; Strittmatter, U.; Fuchs, S.; Quesniaux, V. F. J.;
Williams, S. J.; Wuest, W. M. Diverted Total Synthesis of Promysalin Brinkmann, V.; Schuler, W.; Zurini, M.; Enz, A.; Billich, A.; Sanglier, J.
Analogs Demonstrates That an Iron-Binding Motif Is Responsible for J.; Fehr, T. Sanglifehrin A, a novel cyclophilin-binding compound
Its Narrow-Spectrum Antibacterial Activity. J. Am. Chem. Soc. 2016, showing immuno suppressive activity with a new mechanism of
138, 5833−5836. action. J. Immunol. 2001, 166, 7165−7171.
(218) Greiner, D.; Bonaldi, T.; Eskeland, R.; Roemer, E.; Imhof, A. (232) Kallen, J.; Sedrani, R.; Zenke, G.; Wagner, J. Structure of
Identification of a specific inhibitor of the histone methyltransferase human cyclophilin A in complex with the novel immunosuppressant
SU(VAR)3−9. Nat. Chem. Biol. 2005, 1, 143−145. sanglifehrin A at 1.6 A resolution. J. Biol. Chem. 2005, 280, 21965−
(219) Fujishiro, S.; Dodo, K.; Iwasa, E.; Teng, Y.; Sohtome, Y.; 21971.
Hamashima, Y.; Ito, A.; Yoshida, M.; Sodeoka, M. Epidithiodiketo- (233) Sedrani, R.; Kallen, J.; Martin Cabrejas, L. M.; Papageorgiou,
piperazine as a pharmacophore for protein lysine methyltransferase C. D.; Senia, F.; Rohrbach, S.; Wagner, D.; Thai, B.; Jutzi Eme, A. M.;
G9a inhibitors: reducing cytotoxicity by structural simplification. France, J.; Oberer, L.; Rihs, G.; Zenke, G.; Wagner, J. Sanglifehrin-
Bioorg. Med. Chem. Lett. 2013, 23, 733−736. cyclophilin interaction: degradation work, synthetic macrocyclic
(220) Taori, K.; Paul, V. J.; Luesch, H. Structure and activity of analogues, X-ray crystal structure, and binding data. J. Am. Chem.
largazole, a potent antiproliferative agent from the Floridian marine Soc. 2003, 125, 3849−3859.

AL DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(234) Steadman, V. A.; Pettit, S. B.; Poullennec, K. G.; Lazarides, L.; (252) Cheng, Y.; Schneider, B.; Riese, U.; Schubert, B.; Li, Z.;
Keats, A. J.; Dean, D. K.; Stanway, S. J.; Austin, C. A.; Sanvoisin, J. A.; Hamburger, M. Farinosones A-C, neurotrophic alkaloidal metabolites
Watt, G. M.; Fliri, H. G.; Liclican, A. C.; Jin, D. B.; Wong, M. H.; from the entomogenous deuteromycete Paecilomyces farinosus. J.
Leavitt, S. A.; Lee, Y. J.; Tian, Y.; Frey, C. R.; Appleby, T. C.; Schmitz, Nat. Prod. 2004, 67, 1854−1858.
U.; Jansa, P.; Mackman, R. L.; Schultz, B. E. Discovery of Potent (253) Isaka, M.; Chinthanom, P.; Supothina, S.; Tobwor, P.; Hywel-
Cyclophilin Inhibitors Based on the Structural Simplification of Jones, N. L. Pyridone and tetramic acid alkaloids from the spider
Sanglifehrin A. J. Med. Chem. 2017, 60, 1000−1017. pathogenic fungus Torrubiella sp. BCC 2165. J. Nat. Prod. 2010, 73,
(235) Hayakawa, Y.; Adachi, K.; Komeshima, N. New antitumor 2057−2060.
antibiotics, anguinomycins A and B. J. Antibiot. 1987, 40, 1349−1352. (254) Jessen, H. J.; Schumacher, A.; Shaw, T.; Pfaltz, A.; Gademann,
(236) Hayakawa, Y.; Sohda, K. Y.; Shin-Ya, K.; Hidaka, T.; Seto, H. K. A unified approach for the stereoselective total synthesis of
Anguinomycins C and D, new antitumor antibiotics with selective pyridone alkaloids and their neuritogenic activity. Angew. Chem., Int.
cytotoxicity against transformed cells. J. Antibiot. 1995, 48, 954−961. Ed. 2011, 50, 4222−4226.
(237) Newlands, E. S.; Rustin, G. J.; Brampton, M. H. Phase I trial of (255) Schmid, F.; Jessen, H. J.; Burch, P.; Gademann, K. Truncated
elactocin. Br. J. Cancer 1996, 74, 648−649. militarinone fragments identified by total chemical synthesis induce
(238) Sun, Q.; Carrasco, Y. P.; Hu, Y.; Guo, X.; Mirzaei, H.;
neurite outgrowth. MedChemComm 2013, 4, 135−139.
Macmillan, J.; Chook, Y. M. Nuclear export inhibition through
(256) Degnan, B. M.; Hawkins, C. J.; Lavin, M. F.; McCaffrey, E. J.;
covalent conjugation and hydrolysis of Leptomycin B by CRM1. Proc.
Parry, D. L.; Watters, D. J. Novel cytotoxic compounds from the
Natl. Acad. Sci. U. S. A. 2013, 110, 1303−1308.
ascidian Lissoclinum bistratum. J. Med. Chem. 1989, 32, 1354−1359.
(239) Bonazzi, S.; Guttinger, S.; Zemp, I.; Kutay, U.; Gademann, K.
(257) Gouiffès, D.; Moreau, S.; Helbecque, N.; Bernier, J. L.;
Total synthesis, configuration, and biological evaluation of anguino-
mycin C. Angew. Chem., Int. Ed. 2007, 46, 8707−8710. Hénichart, J. P.; Barbin, Y.; Laurent, D.; Verbist, J. F. Proton Nuclear
(240) Kudo, N.; Wolff, B.; Sekimoto, T.; Schreiner, E. P.; Yoneda, Magnetic Study of Bistramide A, a new cytotoxic drug isolated from
Y.; Yanagida, M.; Horinouchi, S.; Yoshida, M. Leptomycin B Lissoclinum Bistratum Sluiter. Tetrahedron 1988, 44, 451−459.
inhibition of signal-mediated nuclear export by direct binding to (258) Biard, J. F.; Roussakis, C.; Kornprobst, J. M.; Gouiffes-Barbin,
CRM1. Exp. Cell Res. 1998, 242, 540−547. D.; Verbist, J. F.; Cotelle, P.; Foster, M. P.; Ireland, C. M.; Debitus, C.
(241) Bonazzi, S.; Eidam, O.; Guttinger, S.; Wach, J. Y.; Zemp, I.; Bistramides A, B, C, D, and K: a new class of bioactive cyclic
Kutay, U.; Gademann, K. Anguinomycins and derivatives: total polyethers from Lissoclinum bistratum. J. Nat. Prod. 1994, 57, 1336−
syntheses, modeling, and biological evaluation of the inhibition of 1345.
nucleocytoplasmic transport. J. Am. Chem. Soc. 2010, 132, 1432− (259) Sauviat, M. P.; Verbist, J. F. Alteration of the voltage-
1442. dependence of the twitch tension in frog skeletal muscle fibres by a
(242) Igarashi, M.; Nakagawa, N.; Doi, N.; Hattori, S.; Naganawa, polyether, Bistramide A. Gen. Physiol. Biophys. 1993, 12, 465−471.
H.; Hamada, M. Caprazamycin B, a novel anti-tuberculosis antibiotic, (260) Sauviat, M. P.; Gouiffes-Barbin, D.; Ecault, E.; Verbist, J. F.
from Streptomyces sp. J. Antibiot. 2003, 56, 580−583. Blockade of sodium channels by Bistramide A in voltage-clamped frog
(243) Bugg, T. D.; Lloyd, A. J.; Roper, D. I. Phospho-MurNAc- skeletal muscle fibres. Biochim. Biophys. Acta, Biomembr. 1992, 1103,
pentapeptide translocase (MraY) as a target for antibacterial agents 109−114.
and antibacterial proteins. Infect. Disord.: Drug Targets 2006, 6, 85− (261) Griffiths, G.; Garrone, B.; Deacon, E.; Owen, P.; Pongracz, J.;
106. Mead, G.; Bradwell, A.; Watters, D.; Lord, J. The polyether bistratene
(244) Bouhss, A.; Mengin-Lecreulx, D.; Le Beller, D.; Van A activates protein kinase C-delta and induces growth arrest in HL60
Heijenoort, J. Topological analysis of the MraY protein catalysing cells. Biochem. Biophys. Res. Commun. 1996, 222, 802−808.
the first membrane step of peptidoglycan synthesis. Mol. Microbiol. (262) Statsuk, A. V.; Bai, R.; Baryza, J. L.; Verma, V. A.; Hamel, E.;
1999, 34, 576−585. Wender, P. A.; Kozmin, S. A. Actin is the primary cellular receptor of
(245) Bouhss, A.; Trunkfield, A. E.; Bugg, T. D.; Mengin-Lecreulx, bistramide A. Nat. Chem. Biol. 2005, 1, 383−388.
D. The biosynthesis of peptidoglycan lipid-linked intermediates. (263) Rizvi, S. A.; Liu, S.; Chen, Z.; Skau, C.; Pytynia, M.; Kovar, D.
FEMS Microbiol. Rev. 2008, 32, 208−233. R.; Chmura, S. J.; Kozmin, S. A. Rationally simplified bistramide
(246) Al-Dabbagh, B.; Henry, X.; El Ghachi, M.; Auger, G.; Blanot, analog reversibly targets actin polymerization and inhibits cancer
D.; Parquet, C.; Mengin-Lecreulx, D.; Bouhss, A. Active site mapping progression in vitro and in vivo. J. Am. Chem. Soc. 2010, 132, 7288−
of MraY, a member of the polyprenyl-phosphate N-acetylhexosamine 7290.
1-phosphate transferase superfamily, catalyzing the first membrane (264) Rizvi, S. A.; Tereshko, V.; Kossiakoff, A. A.; Kozmin, S. A.
step of peptidoglycan biosynthesis. Biochemistry 2008, 47, 8919− Structure of bistramide A-actin complex at a 1.35 angstroms
8928.
resolution. J. Am. Chem. Soc. 2006, 128, 3882−3883.
(247) Ii, K.; Ichikawa, S.; Al-Dabbagh, B.; Bouhss, A.; Matsuda, A.
(265) Rizvi, S. A.; Courson, D. S.; Keller, V. A.; Rock, R. S.; Kozmin,
Function-oriented synthesis of simplified caprazamycins: discovery of
S. A. The dual mode of action of bistramide A entails severing of
oxazolidine-containing uridine derivatives as antibacterial agents
filamentous actin and covalent protein modification. Proc. Natl. Acad.
against drug-resistant bacteria. J. Med. Chem. 2010, 53, 3793−3813.
(248) Nakamura, H.; Tsukano, C.; Yasui, M.; Yokouchi, S.; Igarashi, Sci. U. S. A. 2008, 105, 4088−4092.
M.; Takemoto, Y. Total synthesis of (−)-caprazamycin A. Angew. (266) Misico, R. I.; Nicotra, V. E.; Oberti, J. C.; Barboza, G.; Gil, R.
Chem., Int. Ed. 2015, 54, 3136−3139. R.; Burton, G. Withanolides and related steroids. Prog. Chem. Org.
(249) Takahashi, Y.; Igarashi, M.; Miyake, T.; Soutome, H.; Nat. Prod. 2011, 94, 127−229.
Ishikawa, K.; Komatsuki, Y.; Koyama, Y.; Nakagawa, N.; Hattori, S.; (267) Lavie, D.; Glotter, E.; Shvo, Y. Constituents of Withania
Inoue, K.; Doi, N.; Akamatsu, Y. Novel semisynthetic antibiotics from somnifera Dun. III. The Side Chain of Withaferin A. J. Org. Chem.
caprazamycins A-G: caprazene derivatives and their antibacterial 1965, 30, 1774−1778.
activity. J. Antibiot. 2013, 66, 171−178. (268) Lee, J.; Hahm, E. R.; Singh, S. V. Withaferin A inhibits
(250) Schmidt, K.; Gunther, W.; Stoyanova, S.; Schubert, B.; Li, Z.; activation of signal transducer and activator of transcription 3 in
Hamburger, M. Militarinone A, a neurotrophic pyridone alkaloid from human breast cancer cells. Carcinogenesis 2010, 31, 1991−1998.
Paecilomyces militaris. Org. Lett. 2002, 4, 197−199. (269) Kupchan, S. M.; Anderson, W. K.; Bollinger, P.; Doskotch, R.
(251) Schmidt, K.; Riese, U.; Li, Z.; Hamburger, M. Novel tetramic W.; Smith, R. M.; Saenz-Renauld, J. A.; Schnoes, H. K.; Burlingame,
acids and pyridone alkaloids, militarinones B, C, and D, from the A. L.; Smith, D. H. Tumor inhibitors. XXXIX. Active principles of
insect pathogenic fungus Paecilomyces militaris. J. Nat. Prod. 2003, Acnistus arborescens. Isolation and structural and spectral studies of
66, 378−383. withaferin A and withacnistin. J. Org. Chem. 1969, 34, 3858−3866.

AM DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(270) Tahara, T.; Streit, U.; Pelish, H. E.; Shair, M. D. STAT3 evaluation of a spongistatin AB-spiroketal analogue. Bioorg. Med.
Inhibitory Activity of Structurally Simplified Withaferin A Analogues. Chem. Lett. 2002, 12, 2039−2042.
Org. Lett. 2017, 19, 1538−1541. (288) Paterson, I.; Aceña, J. L.; Bach, J.; Chen, D. Y. K.; Coster, M.
(271) Yamazaki, Y.; Kunimoto, S.; Ikeda, D. Rakicidin A: a hypoxia- J. Synthesis and biological evaluation of spongistatin/altohyrtin
selective cytotoxin. Biol. Pharm. Bull. 2007, 30, 261−265. analogues: E-ring dehydration and C46 side-chain truncation. Chem.
(272) Wilson, W. R.; Hay, M. P. Targeting hypoxia in cancer Commun. 2003, 3, 462−463.
therapy. Nat. Rev. Cancer 2011, 11, 393−410. (289) Wagner, C. E.; Wang, Q.; Melamed, A.; Fairchild, C. R.; Wild,
(273) McBrien, K. D.; Berry, R. L.; Lowe, S. E.; Neddermann, K. M.; R.; Heathcock, C. H. Synthesis and biological evaluation of analogs of
Bursuker, I.; Huang, S.; Klohr, S. E.; Leet, J. E. Rakicidins, new altohyrtin C (spongistatin 2). Tetrahedron 2008, 64, 124−136.
cytotoxic lipopeptides from Micromonospora sp. fermentation, (290) Nakae, K.; Yoshimoto, Y.; Sawa, T.; Homma, Y.; Hamada, M.;
isolation and characterization. J. Antibiot. 1995, 48, 1446−1452. Takeuch, T.; Imoto, M. Migrastatin, a new inhibitor of tumor cell
(274) Takeuchi, M.; Ashihara, E.; Yamazaki, Y.; Kimura, S.; migration from Streptomyces sp. MK929−43F1. Taxonomy,
Nakagawa, Y.; Tanaka, R.; Yao, H.; Nagao, R.; Hayashi, Y.; Hirai, fermentation, isolation and biological activities. J. Antibiot. 2000, 53,
H.; Maekawa, T. Rakicidin A effectively induces apoptosis in hypoxia 1130−1136.
adapted Bcr-Abl positive leukemic cells. Cancer Sci. 2011, 102, 591− (291) Nakae, K.; Yoshimoto, Y.; Ueda, M.; Sawa, T.; Takahashi, Y.;
596. Naganawa, H.; Takeuchi, T.; Imoto, M. Migrastatin, a novel 14-
(275) Tsakos, M.; Jacobsen, K. M.; Yu, W.; Villadsen, N. L.; membered lactone from Streptomyces sp. MK929−43F1. J. Antibiot.
Poulsen, T. B. The Rakicidin Family of Anticancer Natural Products 2000, 53, 1228−1230.
− Synthetic Strategies towards a New Class of Hypoxia-Selective (292) Njardarson, J. T.; Gaul, C.; Shan, D.; Huang, X. Y.;
Cytotoxins. Synlett 2016, 27, 1898−1906. Danishefsky, S. J. Discovery of potent cell migration inhibitors
(276) Sang, F.; Li, D.; Sun, X.; Cao, X.; Wang, L.; Sun, J.; Sun, B.; through total synthesis: lessons from structure-activity studies of
Wu, L.; Yang, G.; Chu, X.; Wang, J.; Dong, C.; Geng, Y.; Jiang, H.; (+)-migrastatin. J. Am. Chem. Soc. 2004, 126, 1038−1040.
Long, H.; Chen, S.; Wang, G.; Zhang, S.; Zhang, Q.; Chen, Y. Total (293) Gaul, C.; Njardarson, J. T.; Danishefsky, S. J. The total
synthesis and determination of the absolute configuration of rakicidin synthesis of (+)-migrastatin. J. Am. Chem. Soc. 2003, 125, 6042−6043.
A. J. Am. Chem. Soc. 2014, 136, 15787−15791. (294) Shan, D.; Chen, L.; Njardarson, J. T.; Gaul, C.; Ma, X.;
(277) Chen, J.; Li, J.; Wu, L.; Geng, Y.; Yu, J.; Chong, C.; Wang, M.; Danishefsky, S. J.; Huang, X. Y. Synthetic analogues of migrastatin
Gao, Y.; Bai, C.; Ding, Y.; Chen, Y.; Zhang, Q. Syntheses and anti- that inhibit mammary tumor metastasis in mice. Proc. Natl. Acad. Sci.
pancreatic cancer activities of rakicidin A analogues. Eur. J. Med. U. S. A. 2005, 102, 3772−3776.
Chem. 2018, 151, 601−627. (295) Oskarsson, T.; Nagorny, P.; Krauss, I. J.; Perez, L.; Mandal,
(278) Oku, N.; Matoba, S.; Yamazaki, Y. M.; Shimasaki, R.; M.; Yang, G.; Ouerfelli, O.; Xiao, D.; Moore, M. A.; Massague, J.;
Miyanaga, S.; Igarashi, Y. Complete stereochemistry and preliminary Danishefsky, S. J. Diverted total synthesis leads to the generation of
structure-activity relationship of rakicidin A, a hypoxia-selective promising cell-migration inhibitors for treatment of tumor metastasis:
cytotoxin from Micromonospora sp. J. Nat. Prod. 2014, 77, 2561− in vivo and mechanistic studies on the migrastatin core ether analog. J.
2565. Am. Chem. Soc. 2010, 132, 3224−3228.
(279) Tsakos, M.; Clement, L. L.; Schaffert, E. S.; Olsen, F. N.; (296) Chen, L.; Yang, S.; Jakoncic, J.; Zhang, J. J.; Huang, X. Y.
Rupiani, S.; Djurhuus, R.; Yu, W.; Jacobsen, K. M.; Villadsen, N. L.; Migrastatin analogues target fascin to block tumour metastasis. Nature
Poulsen, T. B. Total Synthesis and Biological Evaluation of Rakicidin 2010, 464, 1062−1066.
A and Discovery of a Simplified Bioactive Analogue. Angew. Chem., (297) Tichenor, M. S.; MacMillan, K. S.; Stover, J. S.; Wolkenberg,
Int. Ed. 2016, 55, 1030−1035. S. E.; Pavani, M. G.; Zanella, L.; Zaid, A. N.; Spalluto, G.; Rayl, T. J.;
(280) Pettit, G. R.; Chicacz, Z. A.; Gao, F.; Herald, C. L.; Boyd, M. Hwang, I.; Baraldi, P. G.; Boger, D. L. Rational design, synthesis, and
R.; Schmidt, J. M.; Hooper, J. N. A. Antineoplastic agents. 257. evaluation of key analogues of CC-1065 and the duocarmycins. J. Am.
Isolation and structure of spongistatin 1. J. Org. Chem. 1993, 58, Chem. Soc. 2007, 129, 14092−14099.
1302−1304. (298) MacMillan, K. S.; Lajiness, J. P.; Cara, C. L.; Romagnoli, R.;
(281) Pettit, G. R.; Cichacz, Z. A.; Gao, F.; Herald, C. L.; Boyd, M. Robertson, W. M.; Hwang, I.; Baraldi, P. G.; Boger, D. L. Synthesis
R. Isolation and structure of the remarkable human cancer cell growth and evaluation of a thio analogue of duocarmycin SA. Bioorg. Med.
inhibitors spongistatins 2 and 3 from an eastern indian ocean Spongia Chem. Lett. 2009, 19, 6962−6965.
sp. J. Chem. Soc., Chem. Commun. 1993, 1166−1168. (299) Rujirawanich, J.; Kim, S.; Ma, A. J.; Butler, J. R.; Wang, Y.;
(282) Fusetani, N.; Shinoda, K.; Matsunaga, S. Bioactive marine Wang, C.; Rosen, M.; Posner, B.; Nijhawan, D.; Ready, J. M. Synthesis
metabolites. 48. Cinachyrolide A: a potent cytotoxic macrolide and Biological Evaluation of Kibdelone C and Its Simplified
possessing two spiro ketals from marine sponge Cinachyra sp. J. Am. Derivatives. J. Am. Chem. Soc. 2016, 138, 10561−10570.
Chem. Soc. 1993, 115, 3977−3981. (300) Schnermann, M. J.; Beaudry, C. M.; Egorova, A. V.;
(283) Kobayashi, M.; Aoki, S.; Sakai, H.; Kawazoe, K.; Kihara, N.; Polishchuk, R. S.; Sutterlin, C.; Overman, L. E. Golgi-modifying
Sasaki, T.; Kitagawa, I. Altohyrtin A, a potent anti-tumor macrolide properties of macfarlandin E and the synthesis and evaluation of its
from the Okinawan marine sponge Hyrtios altum. Tetrahedron Lett. 2,7-dioxabicyclo[3.2.1]octan-3-one core. Proc. Natl. Acad. Sci. U. S. A.
1993, 34, 2795−2798. 2010, 107, 6158−6163.
(284) Bai, R.; Cichacz, Z. A.; Herald, C. L.; Pettit, G. R.; Hamel, E. (301) Chen, L.; Riaz Ahmed, K. B.; Huang, P.; Jin, Z. Design,
Spongistatin 1, a highly cytotoxic, sponge-derived, marine natural synthesis, and biological evaluation of truncated superstolide A.
product that inhibits mitosis, microtubule assembly, and the binding Angew. Chem., Int. Ed. 2013, 52, 3446−3449.
of vinblastine to tubulin. Mol. Pharmacol. 1993, 44, 757−766. (302) Just-Baringo, X.; Bruno, P.; Pitart, C.; Vila, J.; Albericio, F.;
(285) Smith, A. B., 3rd; Risatti, C. A.; Atasoylu, O.; Bennett, C. S.; Alvarez, M. Dissecting the structure of thiopeptides: assessment of
Liu, J.; Cheng, H.; TenDyke, K.; Xu, Q. Design, synthesis, and thiazoline and tail moieties of baringolin and antibacterial activity
biological evaluation of diminutive forms of (+)-spongistatin 1: optimization. J. Med. Chem. 2014, 57, 4185−4195.
lessons learned. J. Am. Chem. Soc. 2011, 133, 14042−14053. (303) Sohoel, H.; Liljefors, T.; Ley, S. V.; Oliver, S. F.; Antonello, A.;
(286) Smith, A. B.; Lin, Q.; Pettit, G. R.; Chapuis, J.-C.; Schmidt, J. Smith, M. D.; Olsen, C. E.; Isaacs, J. T.; Christensen, S. B. Total
M. Synthesis and in vitro cancer cell growth inhibitory activity of synthesis of two novel subpicomolar sarco/endoplasmatic reticulum
monocyclic model compounds containing spongistatin triene side- Ca2+-ATPase inhibitors designed by an analysis of the binding site of
chains. Bioorg. Med. Chem. Lett. 1998, 8, 567−568. thapsigargin. J. Med. Chem. 2005, 48, 7005−7011.
(287) Smith, A. B.; Corbett, R. M.; Pettit, G. R.; Chapuis, J.-C.; (304) Alonso, E.; Fuwa, H.; Vale, C.; Suga, Y.; Goto, T.; Konno, Y.;
Schmidt, J. M.; Hamel, E.; Jung, M. K. Synthesis and biological Sasaki, M.; LaFerla, F. M.; Vieytes, M. R.; Gimenez-Llort, L.; Botana,

AN DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

L. M. Design and synthesis of skeletal analogues of gambierol:


attenuation of amyloid-beta and tau pathology with voltage-gated
potassium channel and N-methyl-D-aspartate receptor implications. J.
Am. Chem. Soc. 2012, 134, 7467−7479.
(305) Zheng, S.; Laraia, L.; O'Connor, C. J.; Sorrell, D.; Tan, Y. S.;
Xu, Z.; Venkitaraman, A. R.; Wu, W.; Spring, D. R. Synthesis and
biological profiling of tellimagrandin I and analogues reveals that the
medium ring can significantly modulate biological activity. Org.
Biomol. Chem. 2012, 10, 2590−2593.
(306) Peng, W.; Tang, P.; Hu, X.; Liu, J. O.; Yu, B. Synthesis of the
A,B-ring-truncated OSW saponin analogs and their antitumor
activities. Bioorg. Med. Chem. Lett. 2007, 17, 5506−5509.
(307) Bauer, W.; Briner, U.; Doepfner, W.; Haller, R.; Huguenin, R.;
Marbach, P.; Petcher, T. J.; Pless, J. Pless. SMS 201−995: a very
potent and selective octapeptide analogue of somatostatin with
prolonged action. Life Sci. 1982, 31, 1133−1140.
(308) Veber, D. F.; Freidlinger, R. M.; Perlow, D. S.; Paleveda, W. J.,
Jr.; Holly, F. W.; Strachan, R. G.; Nutt, R. F.; Arison, B. H.; Homnick,
C.; Randall, W. C.; Glitzer, M. S.; Saperstein, R.; Hirschmann, R. A
potent cyclic hexapeptide analogue of somatostatin. Nature 1981, 292,
55−58.
(309) Wender, P. A.; Quiroz, R. V.; Stevens, M. C. Function through
synthesis-informed design. Acc. Chem. Res. 2015, 48, 752−760.
(310) Wilson, R. M.; Danishefsky, S. J. Small molecule natural
products in the discovery of therapeutic agents: the synthesis
connection. J. Org. Chem. 2006, 71, 8329−8351.
(311) Roach, J. J.; Sasano, Y.; Schmid, C. L.; Zaidi, S.; Katritch, V.;
Stevens, R. C.; Bohn, L. M.; Shenvi, R. A. Dynamic Strategic Bond
Analysis Yields a Ten-Step Synthesis of 20-nor-Salvinorin A, a Potent
kappa-OR Agonist. ACS Cent. Sci. 2017, 3, 1329−1336.
(312) von Nussbaum, F.; Brands, M.; Hinzen, B.; Weigand, S.;
Habich, D. Antibacterial natural products in medicinal chemistry–
exodus or revival? Angew. Chem., Int. Ed. 2006, 45, 5072−5129.
(313) Brown, D. G.; Lister, T.; May-Dracka, T. L. New natural
products as new leads for antibacterial drug discovery. Bioorg. Med.
Chem. Lett. 2014, 24, 413−418.

AO DOI: 10.1021/acs.chemrev.8b00504
Chem. Rev. XXXX, XXX, XXX−XXX

You might also like