Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Pharmacology & Therapeutics 162 (2016) 35–57

Contents lists available at ScienceDirect

Pharmacology & Therapeutics

journal homepage: www.elsevier.com/locate/pharmthera

Associate Editor: Y. Zhang

Pharmacological modulation of oncogenic Ras by natural products and


their derivatives: Renewed hope in the discovery of novel anti-Ras drugs
Shun Ying Quah a,1, Michelle Siying Tan a,1, Yuan Han Teh a,1, Johnson Stanslas a,b,⁎
a
Pharmacotherapeutics Unit, Department of Medicine, Faculty of Medicine and Health Sciences, Universiti Putra Malaysia, 43400 UPM, Serdang, Selangor, Malaysia
b
Laboratory of Natural Products, Institute of Bioscience, Universiti Putra Malaysia, 43400 UPM, Serdang, Selangor, Malaysia

a r t i c l e i n f o a b s t r a c t

Available online 22 March 2016 Oncogenic rat sarcoma (Ras) is linked to the most fatal cancers such as those of the pancreas, colon, and lung. De-
cades of research to discover an efficacious drug that can block oncogenic Ras signaling have yielded disappoint-
Keywords: ing results; thus, Ras was considered “undruggable” until recently. Inhibitors that directly target Ras by binding to
Oncogenic Ras previously undiscovered pockets have been recently identified. Some of these molecules are either isolated from
Anti-Ras natural products or derived from natural compounds. In this review, we described the potential of these com-
Cancer
pounds and other inhibitors of Ras signaling in drugging Ras. We highlighted the modes of action of these com-
Natural products
Derivatives
pounds in suppressing signaling pathways activated by oncogenic Ras, such as mitogen-activated protein kinase
Strategy (MAPK) signaling and the phosphoinositide-3-kinase (PI3K) pathways. The anti-Ras strategy of these com-
pounds can be categorized into four main types: inhibition of Ras–effector interaction, interference of Ras mem-
brane association, prevention of Ras–guanosine triphosphate (GTP) formation, and downregulation of Ras
proteins. Another promising strategy that must be validated experimentally is enhancement of the intrinsic
Ras–guanosine triphosphatase (GTPase) activity by small chemical entities. Among the inhibitors of Ras signaling
that were reported thus far, salirasib and TLN-4601 have been tested for their clinical efficacy. Although both
compounds passed phase I trials, they failed in their respective phase II trials. Therefore, new compounds of nat-
ural origin with relevant clinical activity against Ras-driven malignancies are urgently needed. Apart from
salirasib and TLN-4601, some other compounds with a proven inhibitory effect on Ras signaling include deriva-
tives of salirasib, sulindac, polyamine, andrographolide, lipstatin, levoglucosenone, rasfonin, and quercetin.
© 2016 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2. Approaches of developing lead compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3. Inhibition of Ras–effector interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4. Interference of Ras membrane association . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
41
5. Prevention of Ras–GTP formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
44
6. Downregulation of Ras proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
48
7. Discussion and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
51
Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
55
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
55
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
55

Abbreviations: AGP, Andrographolide; APT1, Acyl protein thioesterase 1; COX-2, cyclooxygenase-2; DPI, Dual prenylation inhibitor; ERK, Extracellular signal-regulated kinase; FBDD,
Fragment-based drug discovery; FTase, Farnesyltransferase; FTIs, Farnesyltransferase inhibitors; G12, Glycine-12; G13, Glycine-13; Q61, Glutamine-61; GAP, GTPase-activating protein;
GDP, Guanosine diphosphate; GEF, Guanine nucleotide exchange factor; GGTase I, Geranylgeranyltransferase I; GGTI, Geranylgeranyltransferase inhibitor; GTP, Guanosine triphosphate;
GTPase, Guanosine triphosphatase; H-Ras, Harvey-Ras; HTS, High-throughput screening; HVR, Hypervariable region; IC50, Half-maximal inhibitory concentration; ICMT, Isoprenylcysteine
carboxymethyltransferase; K-Ras, Kirsten-Ras; MAPK, Mitogen-activated protein kinase; MEK, Mitogen-activated protein/extracellular signal-regulated kinase kinase; Mg2+, Magnesium
ion; N-Ras, Neuroblastoma-Ras; NMR, Nuclear magnetic resonance; NSCLC, Non-small cell lung cancer; PDAC, Pancreatic ductal adenocarcinoma; PDB, Protein Data Bank; PFTS, Phospho-
farnesylthiosalicylic acid; PI3K, Phosphoinositide-3-kinase; Raf, Rapidly accelerated fibrosarcoma; Ras, Rat sarcoma; SAR, Structure–activity relationship; SBDD, Structure-based drug de-
sign; SRJ23, 3,19-(3-chloro-4-fluorobenzylidene) andrographolide; SOS, Son of Sevenless.
⁎ Corresponding author. Tel.: +60 3 89472310; fax: +60 3 89472759.
E-mail addresses: rcxjs@upm.edu.my, jstanslas@yahoo.co.uk (J. Stanslas).
1
Equal contribution.

http://dx.doi.org/10.1016/j.pharmthera.2016.03.010
0163-7258/© 2016 Elsevier Inc. All rights reserved.
36 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

1. Introduction RAS is a proto-oncogene, mutations of which have been implicated in


abnormalities of cell proliferation, survival, migration, and differentia-
The rat sarcoma (Ras) protein superfamily comprises small guano- tion, which ultimately lead to cancer initiation and progression
sine triphosphate (GTP)-binding proteins with inherent guanosine (Wennerberg et al., 2005). When extracellular signals, such as epidermal
triphosphatase (GTPase) activity, hydrolyzing GTP to guanosine diphos- growth factor (EGF), nerve growth factor (NGF), platelet-derived growth
phate (GDP) (Wennerberg et al., 2005). The members of the Ras family factor (PDGF), fibroblast growth factor (FGF), activate receptor tyrosine
include Harvey-Ras (H-Ras); neuroblastoma-Ras (N-Ras); and the kinases (RTKs), Ras activates two canonical cancer-driving mitogen-
splice variants of Kirsten-Ras (K-Ras), K-Ras-4A, and K-Ras-4B. These activated protein kinase (MAPK) and phosphoinositide-3-kinase (PI3K)
isoforms share highly conserved GDP/GTP-binding motif elements pathways (Rodriguez-Viciana et al., 1994; Lodish et al., 2000; Vivanco
(also known as G boxes) that span the N-terminal G domain & Sawyers, 2002; Sever & Brugge, 2015; Fig. 2). Mutations of Ras have
(Wennerberg et al., 2005). The C-terminal hypervariable region (HVR) been found in 25%–30% of human cancers, including 90% of pancreatic
of Ras is subjected to posttranslational modifications to facilitate the ductal adenocarcinomas (PDACs) (Scheffzek et al., 1997; Prior et al.,
lipid-mediated membrane localization of Ras (Konstantinopoulos 2012). Among the Ras isoforms, K-Ras mutations account for 86% of
et al., 2007). The G domain possesses two switches: switch I (residues RAS-driven cancers, followed by N-Ras (11%) and H-Ras (3%) (Forbes
25–40) and switch II (residues 59–75) (Vetter & Wittinghofer, 2001; et al., 2011). Ras mutations most commonly occur at codon glycine-12
Kapoor & Travesset, 2015). These switches are essential for the recogni- (G12), glycine-13 (G13), and glutamine-61 (Q61) (Prior et al., 2012).
tion of the two nucleotide-bound states of Ras by regulatory proteins G12 mutations cause steric hindrance that prevents Ras–GAP interac-
and effectors, thus determining the protein–protein interactions be- tion, whereas Q61 mutations attenuate the GAP-activating GTPase activ-
tween Ras and these Ras-binding proteins (Vetter & Wittinghofer, ity of Ras, resulting in prolonged activation of Ras signaling, ultimately
2001; Herrmann, 2003; Wennerberg et al., 2005). leading to cancer initiation and progression (Gideon et al., 1992;
Ras proteins are molecular switches that can be turned “on” and Scheffzek et al., 1997). The effect of G13 mutations on the biochemical
“off” by effecting conformational changes of switch I and II via the bind- properties of Ras remains to be elucidated (Gideon et al., 1992). It follows
ing of GTP and GDP, respectively (Vetter & Wittinghofer, 2001). As Ras from the structure of the Ras–GAP complex that G13 mutations involv-
exhibits slow intrinsic activities of GDP–GTP exchange and GTP hydro- ing structurally larger amino acids would cause steric hindrance in the
lysis, the interconversion between its active and inactive conformations Ras–GAP interaction (Prior et al., 2012). Nevertheless, the structural biol-
must be catalyzed by two regulatory proteins: guanine nucleotide ex- ogy of G13 mutations must be investigated further.
change factor (GEF) and GTPase-activating protein (GAP). The GEFs in- Decades of research into drugs that can inhibit oncogenic Ras have
clude Son of Sevenless (SOS) and Ras guanine nucleotide-releasing been unsuccessful. For this reason, Ras is described as “undruggable.”
factor 1 (RasGRF1). GEF promotes an active GTP-bound state of RAS The undruggability of Ras is mainly attributed to the lack of a well-
by facilitating GDP–GTP exchange, whereas GAP maintains Ras in the in- defined binding pocket in the Ras protein structure to accommodate a
active GDP-bound state by accelerating the intrinsic GTPase activity of drug (Ledford, 2015). As the GTP loading of Ras is the hallmark of Ras ac-
Ras (Schmidt & Hall, 2002; Bernards & Settleman, 2004; Fig. 1). tivation, using a GTP-competitive inhibitor to block oncogenic Ras

Fig. 1. Regulation of Ras functions. Extracellular signaling molecules (e.g., EGF) activate RTKs, leading to dimerization of the receptor and reciprocal phosphorylation between the two
adjacent kinases (also known as autophosphorylation). An adaptor protein, GRB2, binds activated RTK at its phosphotyrosine-binding domain and subsequently recruits GEF to the
inner plasma membrane. Ras latches onto the plasma membrane via a lipid molecule and interacts with GEF for its activation. The GEF facilitates the GTP loading of Ras by removing
GDP, allowing the intracellularly abundant GTP to diffuse into the GDP/GTP-binding pocket and bind to Ras. The active GTP-bound form of Ras interacts with the effector (Eff) to relay
a cascade of signaling. To stop this process, GAP interacts physically with GTP-bound active Ras and activates its GTPase activity. The formation of the Ras–GAP complex favors the
hydrolysis of GTP into GDP. The inactive GDP-bound form of Ras acquires a protein conformation incompatible for Ras–effector binding, thus turning off Ras signaling. *Oncogenic Ras
harbors mutations at codon G12, G13, and Q61. These mutations attenuate or completely disable the GAP-activating GTPase activity of Ras, thus leading us to conclude that oncogenic
Ras is constitutively active. Nevertheless, oncogenic Ras retains its low intrinsic GTPase activity, which slowly hydrolyzes Ras-bound GTP to GDP. Therefore, oncogenic Ras is not
constitutively active per se; rather, its active state is prolonged. Abbreviations: EGF, Epidermal growth factor; GAP, GTPase-activating protein; GDP, Guanosine diphosphate; GEF,
Guanine nucleotide exchange factor; GRB2, Growth factor receptor-bound protein 2; GTP, Guanosine triphosphate; GTPase, Guanosine triphosphatase; Ras, Ras sarcoma; RTK, Receptor
tyrosine kinase.
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 37

Fig. 2. Ras-mediated signaling pathways. Ras activates several signaling pathways in the mammalian cells, and the two best-studied pathways are MAPK pathway (Raf–MEK–ERK
pathway) and the PI3K pathway. The two direct downstream effectors of active Ras in the MAPK and PI3K pathways are Raf and PI3K, respectively. The interactions of these effectors
with Ras at the RBD initiate these signaling pathways critical for cell proliferation, survival, migration, and differentiation. The Ras–Raf interaction facilitates the activation of Raf via
phosphorylation. Subsequently, a cascade of activating phosphorylation involving downstream kinases (MEK and ERK) relays Ras-mediated intracellular signaling. Active ERK
subsequently initiates gene transcription by activating transcription factors via phosphorylation. In the PI3K signaling pathway, the active membrane-bound Ras recruits PI3K to the
plasma membrane, where PI3K activates AKT via indirect phosphorylation. Active AKT mediates the activation of cellular targets important in cell growth, proliferation, and survival.
Abbreviations: AKT, v-akt murine thymoma viral oncogene; ERK, Extracellular signal-regulated kinase; GDP, Guanosine diphosphate; GEF, Guanine nucleotide exchange factor; GRB2,
Growth factor receptor-bound protein 2; GTP, Guanosine triphosphate; MAPK, Mitogen-activated protein kinase; MEK, Mitogen-activated protein/extracellular signal-regulated kinase
kinase; PI3K, Phosphoinositide-3-kinase; Raf, Rapidly accelerated fibrosarcoma; Ras, Ras sarcoma; RBD, Ras-binding domain; RTK, Receptor tyrosine kinase.

activation by competing with cellular GTP for binding to mutant Ras can highlighted the possibility of targeting proteins other than Ras to
be considered. Drawing from the success of adenosine triphosphate block oncogenic Ras signaling (Cox et al., 2015). Ras membrane associ-
(ATP)-competitive kinase inhibitors, the use of GTP analogs can be con- ation is a prerequisite for Ras activation and signaling, which is initiated
sidered via modifications at the sugar and base moieties of GTP to inhibit by the addition of a fatty acid or a farnesyl group to the C-terminal HVR
Ras activation (Noonan et al., 1991). Nevertheless, the potency of GTP- of Ras (Prior & Hancock, 2012; Fig. 4). The farnesyl group localizes Ras to
competitive inhibitors was found to be lacking for two reasons: GTP the plasma membrane, which is then embedded in the inner surface of
binds Ras with an extremely high affinity and GTP is abundantly present the plasma membrane (Fig. 4). The farnesylation of Ras is initiated by
intracellularly (John et al., 1993). farnesyltransferase (FTase), an enzyme that is relatively more druggable
As the research on drugs targeting oncogenic Ras directly was un- than Ras (Fig. 4). Indeed, farnesyltransferase inhibitors (FTIs) have
successful, the focus was shifted to the modulation of proteins involved shown promising in vitro and preclinical anticancer activity (Berndt
in Ras activation. Elucidating the mechanism of Ras membrane associa- et al., 2011). With this development, FTIs were considered as solutions
tion, which explains the spatiotemporal activation of Ras, has to the blocking of oncogenic Ras signaling. However, the anticancer
38 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

activity of FTIs at bench could not be translated to the expected efficacy Secondary metabolites represent an abundant resource in drug discov-
in clinical trials (Berndt et al., 2011). Subsequently, the lipidation of Ras ery (Newman & Cragg, 2012). From 1981 to 2010, natural products and
was found to occur via the attachment of a geranylgeranyl group at the their derivatives made up 79.8% of the anticancer drugs approved glob-
C-terminal HVR of N-Ras and K-Ras in the presence of an alternative en- ally (Newman & Cragg, 2012). Most of the small-molecule anticancer
zyme known as geranylgeranyltransferase I (GGTase I) despite the inhi- drugs were discovered by isolating phytochemical compounds from
bition of farnesylation by FTIs (Whyte et al., 1997; Fig. 4). plants, such as alkaloids, flavonoids, phenolic acids, and terpenoids,
Despite the unsuccessful attempts over the decades, researchers followed by structurally modifying active compounds without altering
continue to search for inhibitors that block oncogenic Ras signaling. The their pharmacophores to enhance the biological activity of the com-
recent years have seen a major shift in focus back to targeting oncogenic pounds. Here, we discuss the application of natural products and their
Ras itself. With the advancement of computer modeling in the past derivatives in blocking oncogenic Ras signaling based on four known
5 years, drugging oncogenic Ras via in silico compound screening is an strategies (Fig. 3): inhibition of Ras–effector interactions, interference
exciting prospect for this field (Ledford, 2015). Nevertheless, drugging of Ras membrane association, prevention of Ras–GTP formation, and
the regulatory proteins of Ras signaling continues to be the optimal downregulation of Ras proteins. Another promising strategy that has
method of blocking oncogenic Ras signaling, in addition to drugging not been investigated in depth is enhancement of the intrinsic Ras
Ras directly. GTPase activity by small chemical entities, which forms the basis for de-
Natural products are organic compounds produced by living veloping drugs that target oncogenic Ras signaling in the future (Fig. 3).
organisms. Natural compounds with no direct role in the growth In this review, we specifically focused on small chemical entities of nat-
and development of host organisms but a biological effect on other ural products and their derivatives, which emerge as potential drug can-
organisms are known as secondary metabolites (Hanson, 2003). didates in the treatment of RAS-driven malignancies.

Fig. 3. Strategies for targeting Ras signaling. There are four known strategies by which natural products and their derivatives inhibit Ras function: inhibition of Ras–effector interactions (1),
interference of Ras membrane association (2), abrogation of Ras–GTP formation (3), and downregulation of Ras proteins (4). The efficacy of a prospective strategy that enhances the
intrinsic Ras GTPase activity (?) is yet to be proven, and it can be considered in the future development of small-molecule Ras inhibitors. Further investigations are required to
delineate the exact mechanism of anti-Ras signaling activity in some of the promising compounds (!). Abbreviations: GDP, Guanosine diphosphate; GEF, Guanine nucleotide exchange
factor; GTP, Guanosine triphosphate; GTPase, Guanosine triphosphatase; Ras, Ras sarcoma.
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 39

2. Approaches of developing lead compounds focused on identifying small molecules that can bind to pockets adja-
cent to the Ras–SOS interaction surface by an NMR-based saturation
2.1. Structure-based drug design transfer difference (STD) assay. This assay pooled a group of fragments
for screening and identified ligand segments that are in direct contact
For developing anti-Ras cancer drugs, the focus of research has with the protein. When the fragment binds to a Ras protein, the change
shifted to a structure-based drug design (SBDD) to obtain structural in- in chemical shift of a labeled Ras protein is measured by the changes in
formation in terms of the interaction of small compounds with a drug the resonance position of cross-peaks in the H/N two-dimensional
target. This approach is useful in establishing the structure–activity rela- heteronuclear single-quantum coherence (HSQC) spectra (Zartler &
tionship (SAR) as a measure of therapeutic efficacy (Greer et al., 1994). Mo, 2007). Similarly, the free energy of binding for the protein–ligand
The SBDD approach can be used to identify small-molecule compounds complexes can be calculated via molecular mechanics. It is worth noting
that can fit into pockets of unique conformation that are found on the that ligand screening is based on the conservation of the binding mode
three-dimensional (3D) surface of activated Ras proteins (i.e., GTP of fragments on the protein targets with an increase in ligand size, given
bound). The GTP-bound Ras protein has several active sites. One of that the fragment overlaps well within the main hot spot of the binding
these sites is filled by a molecule that complements the space in terms site with its interacting functional groups (Kozakov et al., 2015). In ad-
of its shape, charge, and other binding components. In fact, SBDD can dition to identifying ligands for a known binding site, FBDD also gener-
be performed even when the target structure is lacking in ated high-resolution structures of the ligand–protein complexes and
pharmacophore modeling, which works on the principle of “molecular revealed unique binding cavities in proteins.
similarity” of small molecules (Chen et al., 2012). In brief, the SBDD
approach provides information on the active conformations or 2.3. High-throughput screening
pharmacophores essential for biological activities and allows the con-
structive modification of small ligands for producing novel structures High-throughput screening (HTS) is used to identify chemical com-
with specific activity and potency. Molecular mechanics such as the mo- pounds with diverse structures, which show activity against RAS pheno-
lecular mechanics/generalized Born surface area (MM-GBSA) method types and can inhibit Ras signaling at different stages (Wang et al.,
can be used to predict the free energy of binding and the strain energy 2013). Synthetic lethal screening is an example of HTS, which is a cell-
of a group of small-molecule ligands in the protein–ligand complex based screening method for identifying small molecules that can selec-
(Prime v3.4, 2013). tively kill cells harboring a specific mutation (Guo et al., 2008). Synthetic
To identify the molecular target of anticancer agents, that is, lethality is dependent on two distinct genes or two related molecular
andrographolide (AGP) and its benzylidene derivatives (Fig. 7), we ap- signaling pathways that influence a common essential biological func-
plied the concept of SBDD to demonstrate the ability of these com- tion, by which cells will die in the case of mutations in the two genes
pounds to bind Ras (Hocker et al., 2013). A total of 75 K-RasQ61H or abnormalities in both signaling pathways. Dysfunction of either one
conformers derived from molecular dynamic (MD) simulations were gene or one signaling pathway will not lead to cell death (Guo et al.,
used for ensemble docking to determine the binding properties of 2008). The mutant Ras proteins in tumor cells are highly active, which
these compounds to multiple protein structures. These conformers prompts cells to eventually rely on other molecular pathways for sur-
were found to have rare sampled structures with some open binding vival (de la Cruz et al., 2015). For example, mutant KRAS-driven non-
sites that were invisible in crystal structures (extracted from the Protein small cell lung cancer (NSCLC) cells produce antiapoptotic signals es-
Data Bank (PDB) managed by Research Collaboratory for Structural Bio- sential for NSCLC survival by activating the nuclear factor kappa light-
informatics (RCSB)). These sites could potentially be accessible to small- chain enhancer of activated B cells (NF-κB) pathway (de Castro
molecule inhibitors. Without specifying any particular pocket, Hocker Carpeño & Belda-Iniesta, 2013). Genetically defined cell lines are used
et al. (2013) blindly docked the compounds against such structures in the HTS approach for screening, with the underlying principle that
(Fig. 8) in the presence of guanine nucleotides to exclude nonspecific RAS and other molecular pathways, such as NF-κB pathway in the
hits at the catalytic site. In particular, three distinct binding pockets above-mentioned case of NSCLC, are blocked and only cancer cells
were found to be preferentially targeted by the compounds. Blind with the mutant RAS gene are killed, thus sparing the normal counter-
docking, which requires the receptor to be rigid while the rotatable parts (Guo et al., 2008). Despite its potential, the use of HTS is limited
bonds of the ligands are flexible, is a ligand-based scheme that seeks by its ability to only determine compounds that can destroy an onco-
binding hot spots and identifies binding sites based on the size of both genic protein without identifying the protein target (Guo et al., 2008).
the ligand and receptor clusters.
3. Inhibition of Ras–effector interactions
2.2. Fragment-based drug discovery
Oncogenic Ras mediates constitutive signal transduction through
The fragment-based drug discovery (FBDD) approach is generally protein–protein interaction with effector proteins such as PI3K and rap-
used to screen and identify small-molecule fragments or lead molecules idly accelerated fibrosarcoma (Raf), Raf-1 in particular. To inhibit Ras
that target the binding sites on protein surfaces or at the interface of signaling, small molecules that can fit and inhibit protein–protein inter-
protein–protein interaction sites (Valkov et al., 2012). Some of these actions are promising candidates in cancer therapy. Ras–GTP can inter-
sites normally lack a defined binding pocket or are considered convert between two conformational states (states 1 and 2) at a
“undruggable” (Maurer et al., 2012; Chandrashekar & Adams, 2013). millisecond time scale; this dynamic equilibrium involves two main re-
Small-molecule fragments are low-molecular-weight (≥ 120 Da but gions of the protein: switch I and switch II. State 2 corresponds to the
b300 Da), moderately lipophilic, highly soluble organic molecules that high-affinity state to effectors whereas state 1 corresponds to a state
usually act as ligands to bind target proteins with low affinity (Rees of lessened affinity to effectors (Spoerner et al., 2005).
et al., 2004). These fragments are usually screened at concentrations Sulindac sulfide (Fig. 5) is obtained via the reductive metabolism of
that permit structure characterization with biophysical techniques, the nonsteroidal anti-inflammatory prodrug sulindac (Fig. 5), which is
such as nuclear magnetic resonance (NMR) spectroscopy and X-ray synthesized from natural safrole (Fig. 5), the main chemical constituent
crystallography (Brenke et al., 2009) and biochemical examination of the essential oil of a Sassafras plant (Barreiro & Lima, 1992; Brunell
(Maurer et al., 2012). All of these techniques form the underlying prin- et al., 2011). Sulindac sulfide is a highly specific dual inhibitor that
ciple of the hit-to-lead optimization process. FBDD allows the ultrasen- binds reversibly to Ras (Herrmann et al., 1998). This compound
sitive detection of interactions between proteins and ligands (Maurer interferes with the interaction between Ras and Raf-1 kinase protein, ul-
et al., 2012). For instance, researchers such as Maurer et al. (2012). timately inhibiting Ras signaling (Herrmann et al., 1998). At
40 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

concentrations ranging from 10 to 50 μM, this compound clearly signal-regulated kinase kinase (MEK) (Herrmann et al., 1998). In addi-
inhibited Ras-induced Raf-1 kinase activity as indicated by the de- tion, sulindac sulfide impairs the nucleotide exchange of Ras by
creased phosphorylation of mitogen-activated protein/extracellular inhibiting Ras–GEF interaction (Herrmann et al., 1998). Sulindac sulfone

Fig. 4. Targeting Ras intracellular trafficking and membrane association. Intracellular localization of Ras to the plasma membrane is a prerequisite for GEF-mediated Ras activation. In the
cytoplasm, all Ras isoforms are attached with a C15 farnesyl isoprenoid by FTase at the cysteine residue of a tetrapeptide CAAX motif (AA: aliphatic residues; X: variable residue) or a C20
geranylgeranyl isoprenoid (N-Ras and K-Ras) by GGTase I in response to the inhibition of farnesylation by FTIs. Subsequently, the AAX peptide of the farnesylated Ras is removed by RCE1
via catalytic proteolysis, producing the terminal farnesylated cysteine residue that is reversibly carboxylmethylated by ICMT in the ER. In the vesicle-mediated classic secretory pathway,
PAT adds C16 palmitate fatty acid to H-Ras, N-Ras, and K-Ras-4 A at the immediate cysteine residues upstream of the terminal farnesylated cysteine residue on the Golgi membrane (Cox,
2010). The palmitoyl group provides Ras with an additional anchorage for plasma membrane association. Once on the plasma membrane, APT1 removes the palmitoyl group on Ras,
allowing Ras to dissociate from the plasma membrane. Conversely, K-Ras-4B lacks the immediate cysteine residues upstream of the terminal farnesylated cysteine residue, making
palmityolation impossible. Instead, farnesylated K-Ras-4B is transported to the plasma membrane via a non-vesicular pathway with the help of PDEδ, which solubilizes farnesylated K-
Ras-4B. In contrast to other Ras isoforms, K-Ras-4B possesses a positively charged polylysine stretch (KKKKKK) as an additional anchoring feature for membrane association by
interacting with the negatively charged phosphate group of phospholipids (Cox et al., 2014). Upon GEF-mediated activation, GTP-bound H-Ras and K-Ras-4B were found to bind to a
membrane-bound protein (galectin), also known as the docking protein of the oncoprotein, via the farnesyl tail (Elad-Sfadia et al., 2002, 2004; Meder & Simons, 2005). *N-Ras is
associated with the plasma membrane in a different manner, although all Ras isoforms are significantly homologous. **The interaction between K-Ras-4 A and galectin has not been
documented. Several natural compounds and their derivatives (salirasib, palmostatin B, spermatinamine, and aplysamine 6) have been reported to block Ras signaling by interfering
with its membrane association. Abbreviations: APT1, acyl protein thioesterase 1; ER: endoplasmic reticulum; RCE1, Ras-converting enzyme-1; FTase, Farnesyltransferase; FTIs,
Farnesyltransferase inhibitors; GEF, Guanine nucleotide exchange factor; GGTase I, Geranylgeranyltransferase I; GTP, Guanosine triphosphate; HVR, Hypervariable region; ICMT,
Isoprenylcysteine carboxymethyltransferase; PAT, protein acyltransferase; Ras, Ras sarcoma.
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 41

is another active metabolite obtained by the oxidation of sulindac Salirasib does not inhibit farnesyltransferase nor does it affect Ras mat-
(Brunell et al., 2011). The sulfide and sulfone metabolites have been uration. According to previous reports, salirasib acts independently of
studied extensively in the case of colorectal cancer with a high frequen- farnesyltransferases and has been found to antagonize Ras activity spe-
cy of K-Ras mutations. Despite the structural similarities between cifically by competing with the active GTP-bound form of Ras (Egozi
sulindac sulfide and sulindac sulfone, these compounds inhibit the et al., 1999; Kloog & Cox, 2004). This compound can dislodge GTP-
Ras-dependent cyclooxygenase-2 (COX-2) pathway in colorectal cancer bound Ras from its membrane anchorage domain (galectin) and accel-
cells through two distinct mechanisms: the inhibition of COX-2 enzyme erate the degradation of the GTP-bound Ras in the cytoplasm (Egozi
activity and of COX-2 expression, respectively (Williams et al., 1999; et al., 1999; Baker & Der, 2013). An in vitro growth inhibitory assay
Taylor et al., 2000). Treatment with sulindac sulfone reduced tumorige- using a PDAC cell line, PANC-1, revealed that salirasib can decrease the
nicity in mice with severe combined immunodeficiency disease (SCID) total amount of activated Ras proteins at concentrations ranging from
bearing human KRAS-driven colorectal and HRAS-driven mammary can- 25 to 50 μM, with a maximal reduction of approximately 50%; serum-
cer cells (Thompson et al., 1997; Taylor et al., 2000). Nevertheless, the dependent and EGF-stimulated extracellular signal-regulated kinase
direct binding of sulindac sulfone to Ras and its inhibitory effect on (ERK) activation was also found to be inhibited (Weisz et al., 1999). Fur-
the Ras signaling pathway are yet to be elucidated. thermore, this compound could prevent the transformation of pancreat-
ic cancer cells harboring oncogenic Ras by inhibiting their anchorage-
4. Interference of Ras membrane association dependent and anchorage-independent growth (Weisz et al., 1999).
On testing the compound for any antitumor activity in vivo, no toxic ef-
The posttranslational lipid modifications of Ras proteins are essential fect was observed in PDAC-xenografted nude mice (Weisz et al., 1999).
for the localization and membrane association of Ras (Fig. 4). Enzymes It is worth noting that prolonged salirasib treatment in vitro did not
that control the posttranslational modification of Ras have emerged as cause drug resistance in pancreatic and colorectal cancer cells; more im-
important targets in the design of anti-Ras drugs (Zhang & Casey, portantly, these salirasib-treated cells became highly sensitive to the
1996). Small molecules that inhibit these enzymes can be potentially chemotherapeutic drug gemcitabine (Gana-Weisz et al., 2002). The sys-
developed into anti-Ras drugs. Prenylation of Ras for its localization to temic exposure of salirasib did not change even when administered in
the plasma membrane involves farnesylation and geranylgeranylation. combination with gemcitabine in the treatment of pancreatic cancer
The carboxyl-terminal residue of the CAAX motif is essential for deter- (Laheru et al., 2012). Although salirasib monotherapy was unsuccessful
mining the isoprenoid added to the protein (Kloog & Cox, 2004). The in a phase II clinical trial involving patients with lung adenocarci-
initial strategy used to inhibit the localization of Ras led to the develop- nomas harboring K-Ras mutations, a combination therapy of
ment of farnesyltransferase inhibitors (Wang et al., 2013). Most FTIs gemcitabine and salirasib in phase I clinical trials was successful in
showed in vivo efficacy against Ras; however, they were ineffective in patients with advanced PDAC (Riely et al., 2011; Laheru et al.,
late-stage clinical trials, as they could not inhibit the signaling of activat- 2012). A recent study showed that patients with refractory hemato-
ed K-Ras and N-Ras isoforms, which could be alternatively prenylated logic malignancies were appreciably tolerant to salirasib monother-
via geranylgeranylation (Whyte et al., 1997; Fig. 4). This finding have apy, which warranted the initiation of phase II clinical trials (Badar
led researchers to shift their focus from drugs targeting Ras by interfer- et al., 2015). Although all previous clinical trials involving salirasib
ing with the Ras membrane association to inhibitors of signaling cas- focused on cancers driven by Ras mutations, none of the patients in
cades downstream of Ras. the recent clinical trial harbored Ras mutations (Badar et al., 2015).
Alternatively, farnesyl cysteine mimetics, such as S-trans, trans- Salirasib can be considered in the treatment of glioblastoma
farnesylthiosalicylic acid (FTS, also known as salirasib; Marom et al., multiforme (GBM), which arises from an overactivation of Ras via
1995; Fig. 6A), have also been developed. This compound is derived an overexpression of EGF and PDGF receptors (Guha et al., 1997;
from salicylic acid (Fig. 6A), a phytochemical found in willow tree Goldberg & Kloog, 2006). Salirasib was found to inhibit glioblastoma
(Chen et al., 2009). The farnesylcysteine of salirasib mimics the cell migration and invasion by the downregulation of PI3K, thus re-
carboxyl-terminal farnesylcysteine carboxymethyl ester of Ras crucial ducing Rac-1 activity and increasing RhoA activation; both of these
for its attachment to the membrane. This compound competes with activities play a role in the arrangement of the actin cytoskeleton
GTP-bound H-Ras and K-Ras-4B proteins for binding to the docking pro- (Goldberg & Kloog, 2006). Salirasib treatment immobilized glioblas-
tein of the oncoprotein (galectin) on the inner leaflet of the plasma toma cells by strengthening the cell attachment to the extracellular
membrane; thus, the Ras membrane association and its localization to matrix via actin rearrangement (Goldberg & Kloog, 2006). In addi-
specific cellular localities for signaling are disrupted (Aharonson et al., tion, salirasib induced apoptotic cell death by decreasing the survivin
1998; Kloog & Cox, 2004; Haklai et al., 2008; Baker & Der, 2013). gene transcripts (Blum et al., 2006).

Reduction Oxidation

Sulindac sulfide Sulindac Sulindac sulfone

Safrole

Fig. 5. Chemical structures of compounds that inhibit Ras–effector interactions.


42 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

A.

Salicylic acid

S-trans-geranyl thiosalicylic acid (GTS) S-geranylgeranyl thiosalicylic acid (GGTS)

Farnesyl group

R1 R2 R3 R4 Compound

OH H H H FTS, also known

as salirasib

OCH3 (methyl ester) H H H FTS-ME

OCH2OCH3 H H H FTS-MOME

(methoxymethylester)

OCH(CH3)2 H H H FTS-IE

(isopropylester)

OCH2C6H5 (bezylester) H H H FTS-BE

NH2 (amide) H H H FTS-A

NHCH3 (methylamide) H H H FTS-MA

N(CH3)2 (dimethylamide) H H H FTS-DMA

OH F (5-fluoro) H H 5-F-FTS

OH Cl (5-chloro) H H 5-Cl-FTS

OH H F (4-fluoro) H 4-F-FTS

OH H Cl (4-chloro) H 4-Cl-FTS

OH H H Cl (3-chloro) 3-Cl-FTS

Phospho-farnesylthiosalicylic acid (PFTS)


S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 43

B. C.

TLN-4601

D.

Aplysamine 6 Spermatinamine

E.

Palmostatin B

Lipstatin
Fig. 6. Chemical structures of compounds that interfere with Ras membrane association. The 15-carbon farnesyl group of salirasib (solid box) and the presence of free carboxylic acid are
crucial for its Ras inhibitory activity. However, the addition of amide or ester and butane–diethyl phosphate (dashed box) to the parent compound could enhance the efficacy of the
derivatives in inhibiting pancreatic cancer cell growth.

In the late 20th century, some researchers made modifications to the 15 μM) in H-ras-transformed EJ cells, an endometrial adenocarcinoma
S-prenyl backbone of salirasib to generate new anti-Ras inhibitors cell line (Aharonson et al., 1998). Similarly, Goldberg et al. synthesized
(Fig. 6A) for a more effective inhibitory effect on the growth of tumor a series of compounds by modifying the carboxyl group of salirasib
cells with overactivated Ras (Aharonson et al., 1998). However, studies (Fig. 6A). Both studies clearly showed that the salirasib derivatives in-
into targeting Ras using the S-prenyl derivatives of salirasib were unsuc- hibit Ras activation by dislodging the Ras proteins from the membrane,
cessful, as suggested by the growth inhibitory activity of the S-prenyl as proven by the marked reduction in Ras–GTP levels and subsequent
derivatives (IC50 = 10–N50 μM) compared with salirasib (IC50 = 5– suppression of the downstream Ras–MAPK pathway (Aharonson et al.,
44 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

1998; Goldberg et al., 2009). Some of the derivatives, such as halogenic- extract of an Australian Verongida marine sponge, Pseudoceratina sp.
substituted derivatives at the C5 position of the benzene ring and ester This compound was found to inhibit ICMT at an IC50 of 1.9 μM
or amide derivatives of thiosalicylic acid, showed greater potency than (Buchanan et al., 2007). Although it is chemically unsuitable as a drug
their parent compound by selectively inhibiting the growth of cells lead, this compound exhibits good potency, which can be further opti-
with activated Ras. Mackenzie et al. (2013) reported a derivative of mized via chemical modifications to target ICMT (Buchanan et al.,
salirasib with improved biological activity, phospho- 2007). Subsequently, this compound was proven to be highly potent
farnesylthiosalicylic acid (PFTS, also known as MDC-1016; Fig. 6A), syn- in inhibiting the growth of cancer cell lines such as leukemia (K562)
thesized by adding butane–diethyl phosphate to the carboxylic group of and colon cancer (HT29) cell lines with a mean 50% growth inhibition
salirasib. Indeed, the in vitro anticancer activity of PFTS (IC50 = 44.1 ± (GI50) value of 0.23 μM (Kottakota et al., 2012). Aplysamine 6 is also a
2.8 μM) against MIA PaCa-2 pancreatic cancer cells was found to be su- natural product that inhibits ICMT, with bromotyrosine as the scaffold
perior to the activity of its parental compound (IC50 N 80 μM). This com- (Fig. 6D). This is also an alkaloid derived from the same species as
pound is nontoxic with significant antitumor activity reported at spermatinamine, with an IC50 value of 14 μM for ICMT (Buchanan
tolerable doses in both an MIA PaCa-2 tumor xenograft model in nude et al., 2008).
mice (50 mg/kg per day, 5 days/week, orally for 25 days) and a KRAS- As illustrated in Fig. 4, it is critical that certain Ras isoforms such as
driven pancreatic carcinogenesis murine model (100 mg/kg per day, N-Ras, H-Ras, and K-Ras-4A are palmitoylated and depalmitoylated
5 days/week, orally for 9 days). Interestingly, only 1 μM PFTS was left after methylation so as to control the steady-state transportation of
unreacted and the rest were metabolized immediately into two metab- Ras between the plasma membrane and the Golgi complex as well as
olites, namely salirasib (29.7 μM) and salirasib glucuronide (10.3 μM), in to prevent nonspecific localization on endomembranes other than the
the murine plasma 1 h after oral administration at a dose of 200 mg/kg. Golgi complex (Rocks et al., 2005). After the CAAX motif is processed
An equimolar dose of its parental compound, salirasib, remains partially at ER, the cysteine residues that are palmitoylated at the Golgi complex
intact (19.8 μM) whereas the rest is metabolized into salirasib glucuro- serve as second signals necessary for vesicular transport to the plasma
nide (37.4 μM) (Mackenzie et al., 2013). PFTS treatment in vitro and membrane for association (Plowman & Hancock, 2005). In contrast to
in vivo led to the inhibition of Ras activation and subsequent reduction the other Ras isoforms, K-Ras-4B does not undergo palmitoylation and
of Ras–MAPK and Ras–PI3K signaling (Mackenzie et al., 2013). To deter- its localization to the plasma membrane is facilitated by a prenyl
mine the preventive effect of PFTS on cell proliferation, the compound binding protein, known as PDEδ (Chandra et al., 2012). The polylysine
was tested in combination with another compound, phospho-valproic region of K-Ras-4B participates in an ionic interaction with the negative-
acid (P-V), a novel signal transducer and activator of transcription 3 ly charged plasma membrane for membrane association (Plowman
(STAT3) inhibitor. Interestingly, Mackenzie et al. (2013) showed that & Hancock, 2005). Depalmitoylation catalyzed by acyl protein
this combination exhibited a synergistic and additive inhibitory effect thioesterase 1 (APT1) is crucial for the redistribution of farnesylated
on the growth of cells harboring mutant K-Ras (MIA PaCa-2) and Ras from all other endomembranes to the Golgi complex, the site at
wild-type K-Ras (BxPC-3), respectively. It is interesting to note that which farnesylated Ras is repalmitoylated and relocalized to the plasma
Aharonson et al. (1998) reported that the presence of at least 15 carbon membrane via a classical secretory pathway (Rocks et al., 2005; Cox,
chains (as in the farnesyl group) and free carboxylic acid were required 2010). Palmostatin B (Fig. 6E), derived from the natural product
for Ras inhibition. Conversely, Goldberg et al. and Mackenzie et al. lipstatin, has been shown to inhibit APT1 (Dekker et al., 2010;
showed that modifications to the carboxylic group of salirasib can en- Hernandez et al., 2013). Lipstatin, a natural product isolated from the
hance the compound's anti-Ras and growth inhibitory activities. actinobacterium Streptomyces toxytricini (Zhu et al., 2014), has been
Another natural compound called TLN-4601 (Fig. 6B), a novel found to inhibit porcine pancreatic lipase (Weibel et al., 1987; Dekker
farnesylated dibenzodiazepinone, was isolated as a secondary metabo- et al., 2010). Based on protein structure similarity clustering (PSSC),
lite from the marine actinomycete bacterium Micromonospora sp. the active site of the lipase, which is also the binding site of lipstatin,
(Charan et al., 2004). Due to its farnesylated moiety, this compound is highly similar to APT1 (Dekker et al., 2010). Lipstatin contains a
was initially considered to inhibit Ras prenylation, but Boufaied et al. β-lactone, and palmostatin B is derived by chemical modification of
subsequently showed that this compound induces proteasomal- the constituents surrounding the β-lactone core (Dekker et al., 2010).
dependent Raf-1 degradation instead. This compound was also found Palmostatin B inhibits APT1 via reversible competitive inhibition
to inhibit the formation of active Ras–GTP, which leads to the downreg- (IC50 = 670 nM) in vitro (Dekker et al., 2010). This β-lactone inhibitor
ulation of Ras–MAPK signaling (Campbell et al., 2010). A phase I clinical disturbs the palmitoylation and depalmitoylation cycle of Ras, thus lo-
study involving patients with GBM, showed the safety and excellent tol- calizing the palmitoylated Ras to the endomembranes, instead of the
erance of this compound at drug plasma concentrations in the micro- plasma membrane (Cox, 2010). This phenomenon subsequently leads
molar range (Mason et al., 2010). In addition, the in vitro and in vivo to partial phenotypic reversion in oncogenic H-RasG12V-transformed fi-
anticancer activity of K-Ras-driven PDAC was also reported, which war- broblasts, particularly the loss of spindle-like phenotype and reestab-
ranted clinical trials in patients with advanced pancreatic cancer lishment of cell–cell contact (Dekker et al., 2010).
(Campbell et al., 2010). As GBM is caused by dysregulated Ras–MAPK
signaling, the anti-GBM activity of the compound can be attributed to 5. Prevention of Ras–GTP formation
the inhibition of Ras–MAPK signaling. Despite the promising results of
the phase I clinical trials, this compound was ineffective in phase II trials Like other small G proteins, Ras contains a natural binding pocket
involving patients with progressive GBM (Mason et al., 2012). The exact that can accommodate either GDP or GTP in its inactive or active form,
mechanism of Raf-1 protein degradation and Ras–GTP depletion in- respectively. In Ras activation, the GDP–GTP nucleotide exchange is a
duced by TLN-4601 remains to be elucidated. crucial step regulated by the GEF (Fig. 1). The GEF proteins, particularly
In addition to FTIs, inhibitors of isoprenylcysteine carboxy- SOS, contain a catalytic subunit composed of a Ras exchanger motif
methyltransferase (ICMT) can prevent the localization of Ras to the (REM) and a cell division cycle 25 (Cdc25) homology core domain. Ac-
plasma membrane (Fig. 4) (Baines et al., 2011). ICMT, an integral mem- cording to Boriack-Sjodin et al. (1998), SOS facilitates GTP loading in
brane protein found on the endoplasmic reticulum (ER), catalyzes the two ways. First, SOS inserts a helical hairpin motif of the Cdc25 domain
methylation of the C-terminal prenylcysteine on prenylated proteins; to displace the switch I region. Second, insertion of SOS distorts the con-
thus, it is crucial for the correct localization of prenylated proteins to cel- formation of the switch II region and alters the chemical environment of
lular membranes (Wright et al., 2009). While conducting an HTS of nat- the nucleotide-binding site for the associated magnesium ion (Mg2+)
ural products for inhibitors of ICMT, a novel polyamine alkaloid, and the phosphate groups of the nucleotide. The resulting steric hin-
spermatinamine, was identified (Fig. 6C), which was purified from the drance on switch I and the remodeling of switch II cause the
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 45

AGP SRJ09

SRJ10 SRJ23
Fig. 7. Chemical structures of AGP and its benzylidene derivatives. The three-membered ring scaffold (dashed box shown in SRJ09) of all AGP derivatives is an important pharmacophore
that allows the binding of these compounds to p1 of the Ras protein.

Switch I

p1

p2

Switch II

Fig. 8. Structural basis for inhibition of Ras–SOS interaction by andrographolide-based compounds. Docking pose of SRJ23 in p1 and p2 of K-RasQ61H protein. SRJ23 bound at p1, which is
located behind switch I (amino acid residues 25–40), by anchoring the lactone ring into the pocket and interacting with the nucleotide. SRJ23 also bound at the groove between switches I
and II (amino acid residues 59–75), with the lactone ring directed up towards the nucleotide binding site. Notably, p2 was found to be critical for accommodating small-molecule inhibitors
in mutant K-Ras.
46 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

A.
Phenylhydroxyl-
amine group

Naphthyl group

Sulfonamide
group

R= H

SCH-53239 SCH-54292 SCH-53870

B.

Compound R1 R2 R3 C-2 bond stereo

(solid arrow)

5 OH

6 OH

7 OH

8 OH

9 OH OH

10 OH OH
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 47

nucleotide-binding site to open and GDP to be released subsequently compounds were initially designed to compete with GDP for the
(Cherfils & Zeghouf, 2013). The GEF activity is promoted by picomolar nucleotide-binding site. They share a common scaffold but are dissimi-
nucleotide affinities of Ras combined with millimolar cytosolic concen- lar in the substituent group on the sulfonamidic nitrogen (Fig. 9A). As
trations of GTP (Spiegel et al., 2014). the guanine and sugar moieties of SCH-53239 and SCH-54292, respec-
The GTPase cycle is complete only when active Ras proteins are tively, were directed away from the binding site, these derivatives
deactivated, whereby GTP is cleaved into GDP. It is worth noting that showed insignificant ligand binding. These molecules showed a high
even when GTPase activity is not initiated by GAP, the Ras protein can potency in inhibiting the GEF-dependent nucleotide exchange with
be inactivated slowly due to the intrinsic GTPase activity (Bos et al., considerably low IC50 values of 0.5–0.7 μM (Palmioli, 2009). Without
2007). Thus, GTP hydrolysis is efficient when regulated by the GAPs. displacing the bound GDP, these SCH compounds bound to a previously
However, single-point mutations in tumorigenic Ras, especially at unidentified hydrophobic pocket adjacent to switch II and close to the
Q61, lead to impaired intrinsic GTP hydrolysis and insensitivity to nucleotide-binding site on the mutant Ras–GDP protein (Taveras et al.,
GAPs (Scheffzek et al., 1997). Impaired intrinsic GTPase activity in- 1997). Drawing from these studies, Peri et al. (2005) designed and syn-
creases the dependence of the nucleotide state of Ras on the relative thesized a new class of Ras inhibitors by retaining the putative
concentration and nucleotide affinity, thus preferring GTP to GDP and pharmacophore groups such as phenylhydroxylamine and sulfonimide
increasing the proportion of active GTP-bound Ras (Scherer et al., groups and attaching them to a bicyclic core derived from the natural
1989). This limits the development and use of GTP-competitive inhibi- sugar D-arabinose (Fig. 9B). These arabinose-derived bicyclic com-
tors; thus, research efforts have shifted from strategies impairing pro- pounds bound stronger, as shown in in silico modeling. However,
tein activation to those preventing the initial Ras–GTP formation. With these compounds are limited by the toxicity and poor metabolic stabil-
the aim of developing novel anti-Ras antitumor drugs, recent research fo- ity of the hydroxylamine group, which is crucial for inhibiting the nucle-
cused on regulatory proteins, as well as Ras itself, as the targets for otide exchange. Few other Ras inhibitors were synthesized based on the
inhibiting nucleotide exchange and the subsequent formation of activated backbone of SCH compounds, in which the sugar moiety was conjugat-
protein. Some of these agents were derived from natural products and ed to increase water solubility and the hydroxylamine group was re-
have led to the discovery of agents that can inhibit Ras–GTP formation. moved to eliminate the nonspecific toxicity (Fig. 10) (Palmioli et al.,
AGP, a bicyclic diterpenoid lactone extracted from the well- 2009; Sacco et al., 2011). When the hydroxylamine group was removed,
known Asian herb Andrographis paniculata (“King of Bitters”), and the nucleotide-exchange inhibitory activity of these newly synthesized
its benzylidene derivatives, namely 3,19-(2-bromobenzylidene) compounds significantly reduced. Nevertheless, the sugar moiety has
andrographolide (SRJ09), 3,19-(3-bromobenzylidene) andrographolide been shown to partially compensate the nucleotide-exchange inhibito-
(SRJ10), and 3,19-(3-chloro-4-fluorobenzylidene) andrographolide ry activity of these derivatives in vitro (IC50 values = 35.5–100 μM), by
(SRJ23) (Fig. 7), have shown promising anticancer activity in vitro, as re- enhancing its binding affinity to Ras. Interestingly, these sugar-based
ported by Jada et al. (2008). All benzylidene derivatives showed more compounds were found to bind p2, which is also the binding site of
potency than the parent compound against breast (MCF-7) and colon other Ras inhibitors as reported by Grant et al. (2011), Maurer et al.
cancer (HCT-116) cells, with the GI50 value ranging from 3.8 to (2012), Sun et al. (2012) and Hocker et al. (2013). All sugar-derived
5.3 μM. Ras was found to be a potential target when determining the compounds were found to inhibit Ras-dependent cell growth and re-
exact molecular target of these compounds by the SBDD approach and duce the phosphorylated MAPK levels, although their IC50 values
ensemble docking (Hocker et al., 2013). Three binding pockets (p1, exceeded those of the SCH compounds (Peri et al., 2005; Palmioli
p2, and p3) preferentially targeted by AGP and its benzylidene deriva- et al., 2009; Sacco et al., 2011).
tives were found on the activated protein surface (PDB ID: 3GFT, Drawing from the work at Schering-Plough involving natural
Fig. 8). p1 is a transient pore located behind switch I, with which products with hydroxy-containing moieties, Müller et al. (2009)
these compounds interacted the most and bound with good affinity also developed new Ras inhibitors derived from levoglucosenone
(micro- to nanomolar). MD simulations showed that SRJ23 bound to (Fig. 11A), a nonplanar chiral bicyclic molecule derived from cellu-
p1 and was stabilized at the binding site by the formation of hydrogen lose. The α,β-unsaturated carbonyl group of levoglucosenone was
bonds between the hydroxyl group on the lactone ring and the α- made to react with aromatic nitrones by 1,3-dipolar cycloaddition
phosphate of GTP, as well as electrostatic contact with Mg2+ (Hocker and was further modified by fusion with an isoxazolidine ring to pro-
et al., 2013, Fig. 8). The AGP and SRJ derivatives were also found to target duce a tricyclic scaffold (Fig. 11A). The novel compounds based on
p2, a groove between switches I and II (Fig. 8). When the ligands bound the tricyclic scaffold varied in terms of the substituents R1, R2, and
to these pockets, the key residues on the switch regions were reoriented R3 . The two most potent compounds (compounds 1 and
and Ras–GEF interaction was subsequently impaired. As it cannot bind 4) (Fig. 11B) were tested in vitro at 50 μM for their inhibitory effect
GAP, oncogenic Ras, which is constitutively GTP loaded, is dependent on the GEF-mediated nucleotide exchange in human Ras protein,
on intrinsic hydrolysis to return to its GDP-bound ground state. Then, p21hRas (Müller et al., 2009). When testing their inhibitory effects
GEF interacts with oncogenic Ras to facilitate GTP loading. This process on various cancer cells, compounds 6, 7, and 8 (Fig. 11B) showed
was blocked, as evidenced by the reduced fraction of GTP-bound high potency in human ovarian (A2780) and breast adenocarcinoma
K-RasG12V in cell-based assays and a concomitant reduction in associ- (MCF-7) cancer cell lines (range of IC50 values = 3.1–7.1 μM) (Müller
ated MAPK signaling after prolonged treatment with AGP derivatives et al., 2009). The research team further tested compounds 1 and 4
(Hocker et al., 2013). against wild-type and mutant K-Ras-transformed NIH3T3 mouse fi-
A group of researchers at the Schering-Plough Research Institute de- broblasts. In contrast to compound 1, which had almost no effect
veloped a guanine nucleotide analog termed SCH-53239 and two non- on both types of cells, compound 4 led to partial and complete
nucleotidic derivatives, SCH-54292 (a derivative of SCH-53239 with a growth inhibition in normal and mutant K-Ras-transformed cells, re-
glucose moiety) and SCH-53870 (Taveras et al., 1997; Fig. 9A), to act spectively. Thus, it was proven to be selective in its inhibitory effect
as Ras binders for inhibiting GEF-catalyzed nucleotide exchange. These on the Ras-dependent growth of mutant cells.

Fig. 9. Chemical structure of SCH compounds and related D-arabinose-derived compounds. (A) Core scaffold of SCH compounds contain three main functional groups: the naphthyl group
(dashed box), the phenylhydroxylamine group (solid box), and the sulfonamide group (dash-dotted box). The naphthyl group of these compounds was inserted into a hydrophobic
binding pocket close to the switch II region. The phenylhydroxylamine group was located close to the bivalent Mg2+ ion and the β-phosphate group of the bound GDP in the binding site.
The interaction occurred via hydrogen bonds with Thr58. The sulfonamide group is significant as it forms hydrogen bond interactions with Gly60, which is important for the restructuring
of switch II and for stabilizing the formation of Ras-GTP. (B) On the basis of SCH compound structures, Peri et al. (2005, 2006) developed the first (compounds 1–4) and second (com-
pounds 5–10) generation of inhibitors containing the bicyclic scaffold (solid circle), by retaining the functional phenylhydroxylamine and benzyl moieties.
48 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

Glucose Lactose
N-acetyl
glucosamine

Compound 2 Compound 3 Compound 4

Fig. 10. Chemical structures of glycosylated Ras inhibitors developed by Sacco et al. (2011) by conjugating sugar moieties found most abundantly in plants and animals, which are glucose,
N-acetyl glucosamine, and lactose (solid boxes).

Lai et al. (2009) identified at least 11 hydrophobic phytoconstituents Talaromyces sp. 3656-A1 (Tomikawa et al., 2000), Trichurus terrophilus
including pheophorbide a (an epimer), pheophorbide a' (an epimer), (Fujimoto et al., 2000; Akiyama et al., 2003; Fujimoto et al., 2005), and
pyropheophorbide a (an epimer), methyl pyropheophorbide a, Doratomyces sp. (Xiao et al., 2014). Compared to its effect on a non-
hexadecanoic acid, oleic acid, linoleic acid, linolenic acid, campesterol, RAS-dependent cell line, this fungal secondary metabolite with an IC50
stigmasterol, and β-sitosterol from a Malaysian medicinal plant value of 0.37 μM was shown to induce apoptosis in Ba/F3-V12 cells
known as Typhonium flagelliforme (Keladi tikus). These compounds (transfected murine cells (Ba/F3) expressing H-RasG12V) (Tomikawa
did not exhibit a significant anticancer effect when administered indi- et al., 2000). The target selectivity of the compound was tested in vitro
vidually, but in combination they showed a synergistic antiproliferative against two different human pancreatic cancer cell lines, BxPC-3
effect on NCI-H23 lung cancer cell line. The first four aforementioned (wild-type KRAS) and Panc-1 (mutated KRAS), and in vivo against
compounds are chlorophyll catabolites and were proven to be Panc-1 (Xiao et al., 2014). It showed a selective inhibitory effect on
photosensitive; that is, they were found to inhibit cell growth with Panc-1 compared with BxPC-3 with an IC50 value of 5.5 μM versus
high potency upon light activation. With mutant K-Ras considered the 10 μM, respectively. Apart from inhibiting cell proliferation, the com-
important prognostic biomarker for NSCLCs (Marks & Pao, 2009), pound also suppressed contact-independent growth, cell migratory ac-
five of the potent compounds (pheophorbide a, pheophorbide a', tivity, and the invasiveness of Panc-1 cells in a dose-dependent manner
pyropheophorbide a, hexadecanoic acid, and 1-hexadecene; Fig. 12) re- (Xiao et al., 2014). In these experiments, salirasib was used as a positive
ported by Lai et al. (2009) were evaluated for their mechanism of action control as it has a well-defined mechanism of action. Like salirasib,
using an in silico approach (Fatima & Yee, 2014). The researchers exam- rasfonin reduced EGF-induced Ras activation in a dose-dependent man-
ined the SOS-dependent nucleotide-exchange inhibitory activity by ner. In addition, it also inhibited kinases downstream of Ras, such as c-
docking the compounds, either directly or allosterically, onto different Raf, MEK, and ERK without affecting the expression of these proteins,
reported binding sites on mutant K-Ras proteins (K-RasG12D protein unlike salirasib, which reduces the cellular levels of the Ras protein
– PDB ID: 4DST with co-crystallized ligand 4,6-dichloro-2-methyl- (Xiao et al., 2014). Moreover, based on the GTP/GDP exchange assay,
3aminoethyl-indole (DCAI) bound to the SOS pocket (Maurer et al., the suppression of Ras activity was attributed to the reduction in ex-
2012); and K-RasG12C proteins – PDB ID: 4LUC with a sulfonamide as pression of SOS1, a GEF. In the in vivo experiment, rasfonin showed
a co-crystallized ligand and 4LYF with a vinyl sulfonamide as a co- greater potency than salirasib in suppressing Panc-1 tumor growth.
crystallized ligand (Ostrem et al., 2013) bound to allosteric sites).
Fatima and Yee demonstrated that pheophorbide epimers bound 6. Downregulation of Ras proteins
strongly to two binding sites, one of which was present at the region
of the Ras–SOS interacting surface whereas the other was located The expression of RAS oncogene is crucial during tumor initiation
at the switch II region. By contrast, long-chained hydrocarbon and maintenance, which is regulated by several Ras–effector pathways
hexadecanoic acid inserted itself well into the binding site similar to including MAPK/ERK and PI3K/AKT pathways (Lim & Counter, 2005).
the co-crystallized ligand, by folding into a U-shape. This compound, Natural compounds in food such as polyphenolics can serve as potential
as well as the unsaturated hydrocarbon 1-hexadecene, can easily slide anticancer agents by modulating cancer cell proliferation and survival.
into and bind with low affinity (range of −4.5 to −5.0 kcal/mol) a pock- Quercetin, also known as 3,3′,4′,5,7-pentahydroxyflavone (Fig. 14), is
et that was previously reported by Ostrem et al. (2013). Consequently, a naturally occurring polyphenolic flavonoid molecule normally found
the binding disrupted the switch regions. Overall, these compounds as a plant pigment in red wine, fruits, vegetables, and grains. Previous
can block oncogenic Ras proteins by inhibiting SOS-initiated nucleotide studies have shown that quercetin is a promising chemopreventive
exchange and subsequent Ras–GTP formation. It should be noted, how- agent in many types of cancer, such as colorectal (Ranelletti et al.,
ever, that such docking studies conducted without MD simulation and 2000; Psahoulia et al., 2007), prostate (Slusarz et al., 2010), and breast
bioassays are speculative (Chen, 2015). Moreover, Ras cannot be con- (Deng et al., 2013) cancer. Particularly in the Ras signaling pathway,
cluded readily as the target of these compounds. This is because confir- treatment of primary colon tumors and human colon cancer cell lines
mations with bioassays are needed to show the inhibitory effect of these with quercetin led to the inhibition of p21-Ras gene expression and de-
compounds on the K-Ras pathway and to further determine their effect crease in the p21–Ras protein levels (Ranelletti et al., 2000). Notably,
on the GEF-mediated nucleotide exchange. Ras expression was significantly inhibited in two colon cancer cell
Rasfonin (Fig. 13) is an unsaturated δ-lactone (or α-pyrone) deriva- lines (Colo-320HSR and Colo-205) after a 48-h treatment with querce-
tive, isolated from different parts of a few fungal species, including tin (IC50 = 0.5 μM) (Ranelletti et al., 2000). The compound was shown
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 49

to be equally effective in inhibiting the expression of all three Ras iso- mechanism was unclear. The chemopreventive effect of quercetin can
forms: H-Ras, K-Ras, and N-Ras (Ranelletti et al., 2000). Similar results be attributed to its binding to either the Raf-1 or MEK protein in the
were obtained in other studies reporting the antiproliferative and anti- Ras–MAPK pathway, as shown by Lee et al. (2008) in the skin epidermal
tumor effects of quercetin on colorectal cancer cells (Psahoulia et al., cell line with inhibition of tumor promoter-induced neoplastic transfor-
2007; Xavier et al., 2009). The responses to quercetin treatment include mation. In another study, Song et al. (2013) showed that quercetin sup-
reduced production of Ras oncoproteins, induced autophagy and apo- pressed cancer cell invasion and migration by binding to PI3K, another
ptosis, and inhibition of cell viability. Psahoulia et al. (2007) also dem- downstream kinase protein directly activated by Ras-GTP, at the ATP
onstrated the effect of quercetin in accelerating the degradation of Ras binding site (Walker et al., 2000). Indeed, quercetin was also shown
oncoproteins, mediated by proteasomal function, although the to inhibit the proliferation of colon cancer cells through its effect on

A.
4
5 3a 3

6 9 2
8a
8 1
7

Levoglucosenone Tricyclic scaffold

B.

1/2 3

4 5
Fig. 11. Chemical structures of levoglucosenone and its derivatives. (A) Chemical structures of levoglucosenone and the tricyclic scaffold derived from levoglucosenone by 1,3-dipolar
addition. (B) Chemical structures of tricyclic compounds 1–9. Compounds 1 and 2 differ in the stereo-configuration at carbon 3 (C3, solid arrow). The presence of ketone group (solid
box) in compounds 1 and 4 could result in decreased cell penetration and, thus, metabolic instability. Compounds 6 and 7 differ in the stereo-configuration at carbon 4 (C4, dotted
arrow). They show potency in inhibiting cell proliferation, probably due to the presence of the benzyl ether (dashed box) at the C4, which increases the lipophilicity that favors cell
penetration.
50 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

6/7 8

9
Fig. 11 (continued).

K-Ras by regulating both the Ras-MAPK and PI3K pathways (Xavier whereby a significant reduction of Ras protein levels was observed in
et al., 2009). The latter pathway is critical for maintaining transformed both colon cancer cell lines and primary colorectal tumors carrying
growth in RAS mutant cell lines both in vitro and in RAS-driven tumor Ras oncogenes. Delgado et al. (2000) attributed this selectivity to the ar-
xenografts in mice (Lim & Counter, 2005). Quercetin displayed selectiv- rest of oncogenic RAS-induced growth and an increase in the messenger
ity for cells harboring oncogenic forms of Ras, thus preferentially induc- RNA (mRNA) and protein levels of p21WAF1, a cyclin-dependent kinase
ing autophagy, cell cycle arrest, and degradation of mutant activated (CDK) inhibitor regulated by the tumor suppressor protein p53. Accord-
Ras proteins over their inactive wild-type counterparts (Psahoulia ing to Delgado et al. (2000), the p21WAF1 gene promoter was signifi-
et al., 2007). Ranelletti et al. (2000) observed a similar phenomenon, cantly activated in transfected NIH3T3 cells with mutant RAS genes in

Pheophorbide a Pheophorbide a Pyropheophorbide a

Hexadecanoic acid 1-Hexadecene


Fig. 12. Chemical structures of selected active compounds found in Typhonium flagelliforme (Keladi tikus).
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 51

treatments that involved prenylated proteins other than K-Ras. Because


of the narrow therapeutic window, FTI-GGTI and DPIs may not be a
promising strategy for blocking oncogenic Ras signaling.
Recently, several novel natural compounds and their derivatives that
target oncogenic Ras and its signaling have emerged, showing antican-
cer activity in the preclinical stage, thus leading to further development
of drugs targeting RAS-driven cancers (Table 1). In this timely review,
we elaborated on natural compounds and their derivatives with an in-
hibitory effect on Ras signaling and anticancer activity at different stages
Fig. 13. Chemical structure of rasfonin. of drug discovery. We categorized these compounds based on the four
known strategies for blocking Ras signaling: inhibition of Ras–effector
interactions, interference of Ras membrane association, prevention of
a p53-independent manner, with less potent effects on K562 cells carry- Ras–GTP formation, and downregulation of Ras proteins. We also
ing wild-type RAS. discussed the anticancer potential of a novel prospective anti-Ras strat-
egy: enhancement of the intrinsic Ras GTPase activity.
7. Discussion and conclusion Sulindac sulfide is a Ras binder with a dual inhibitory effect on Ras
signaling. The compound simultaneously impairs the nucleotide ex-
Natural products represent a rich source of lead compounds for drug change of Ras and inhibits Ras–Raf-1 protein–protein interaction, thus
discovery. These natural compounds can either be used directly as ther- blocking Ras signaling. Due to its dual inhibitory effect on Ras-
apeutic agents or be chemically modified to enhance their biological mediated signaling, this compound is a pharmacologically attractive
activity. compound for targeting RAS-driven cancers. Nevertheless, studies
Ras mutations are linked to some of the most fatal cancers, such as have shown that the biological activity of sulindac sulfide is rather
those of the pancreas, lungs, and colon. However, researchers have weak and its anticancer activity has not been investigated in depth.
failed to discover a clinically efficacious drug that directly targets the on- Therefore, a more potent compound must be derived from sulindac sul-
cogenic Ras protein and its signaling despite several decades of research. fide to strengthen the anti-Ras activity of its existing pharmacophore.
Thus, oncogenic Ras is considered “undruggable” because it lacks bind- Another biologically active metabolite of sulindac, sulindac sulfone, is
ing pockets for the interaction between Ras and biologically active small a promising candidate that must be examined for its role in Ras signal-
molecules, thus making the inhibition of oncogenic Ras activity impos- ing. It has been studied in RAS-driven colorectal and mammary cancers;
sible. Alternative approaches to blocking oncogenic Ras signaling have however, none of these studies showed that the antitumor effect of
been explored by targeting the regulatory proteins of Ras activation, sulindac sulfone arose from the direct inhibition of Ras signaling. Al-
but research efforts have been unsuccessful to date. Blocking Ras activa- though a direct physical interaction between Ras and sulindac sulfone
tion with FTIs is an effective method of abrogating oncogenic Ras signal- has not been documented, the structural similarity of these two com-
ing by targeting other key regulatory proteins in the pathway. Although pounds suggests the possibility of the direct binding of sulindac sulfone
these small molecules initially showed potential as inhibitors of Ras to Ras. However, this hypothesis must be validated by in vitro binding
membrane association, a key step of Ras activation, clinical trials involv- assays and crystallography studies.
ing these compounds for anticancer treatment have been largely Despite the failure of FTIs in clinical trials, blocking oncogenic Ras
unsuccessful. signaling by targeting Ras membrane localization remains an effective
Nevertheless, researchers continue to explore new strategies to target approach, because any key regulatory protein in the Ras membrane lo-
oncogenic Ras and its signaling. As FTIs were found to be clinically inef- calization process serves as a potential target. Salirasib, a farnesyl mimic
fective largely due to an alternative mechanism of Ras lipidation, namely that competes with farnesylated Ras for membrane association, can be
geranylgeranylation, simultaneous inhibition of both farnesylation potentially developed into a drug targeting PDAC, 90% of which harbors
and geranylgeranylation with a combinatory therapy of FTI and a K-Ras mutations (Almoguera et al., 1988; Laghi et al., 2002). Salirasib
geranylgeranyltransferase inhibitor (GGTI) or a dual prenylation inhibi- was found to dislodge active GTP-bound H-Ras and K-Ras-4B mutants
tor (DPI) was considered. Lobell et al. (2001) followed through on this from the plasma membrane by competitively binding with the docking
concept in preclinical models: mice and dogs. Although FTI–GGTI combi- protein of the oncoprotein, galectin (Fig. 4). Subsequently, the active cy-
natory therapy and DPI were known to effectively cause deprenylation of toplasmic H-Ras and K-Ras-4B proteins were found to be degraded. The
K-Ras in vivo, the effect was not selective to K-Ras in either normal or earliest phase II clinical trial was conducted in patients with lung adeno-
tumor mouse tissue. They proved that GGTIs and DPIs caused systemic carcinoma. Unfortunately, the trials were unsuccessful as salirasib was
toxicity at effective doses, mainly through an unknown mechanism in ad- found to be clinically ineffective. In a later study, the clinical efficacy of
dition to myelosuppression, which was ultimately fatal. As a result, the salirasib was reevaluated in combination with gemcitabine in an animal
therapeutic potential of FTI–GGTI and DPIs was significantly limited by model (Laheru et al., 2012). The combination therapy exhibited good
the inherent toxicity, which might be a response to FTI–GGTI and DPI anticancer activity against patient-derived PDAC xenografts and im-
proved the survival rate in nude mice, thus indicating the efficacy of a
combination therapy in targeting RAS-driven cancers (Haklai et al.,
2008; Laheru et al., 2012). This promising preclinical result highlights
the need to evaluate the clinical efficacy of this combination therapy
in patients with advanced PDAC. Combination therapy involving
gemcitabine and other cytotoxic drugs has been used in cancer treat-
ment previously. Erlotinib and gemcitabine was found to significantly
improve the survival rate of advanced pancreatic cancer (Moore et al.,
2007).
PFTS was found to be more potent than its parental compound,
salirasib, in vitro and in vivo (Mackenzie et al., 2013). Salirasib was
Fig. 14. Chemical structure of quercetin. The chromone structure (dashed box) is critical
found to be partially metabolized into salirasib glucuronide, whereas
for its binding to the ATP-binding site of PI3K, which is a protein directly downstream of PFTS was completely metabolized into salirasib and salirasib glucuro-
Ras. nide in the blood plasma after oral administration. Compared with
52
Table 1
Summary of natural compounds and their derivatives as anti-Ras agents.

Compound Class of compound Origin Ras-targeting Specific Phases of drug Inhibitory activity Reference(s)
strategy mechanism development

Sulindac sulfide Phenylpropanoid Derived from safrole, a Inhibition of Ras–effector Binds Ras In vitro Focus formation – Barreiro & Lima, 1992;
compound found in sassafras interactions and inhibition of H-rasG12V-transformed Herrmann et al., 1998;
plants Ras–GEF nucleotide exchange REF (IC50 = 10–50 μM) Brunell et al., 2011
Salirasib Phenolic acids Derived from salicylic acid, a Interference of Ras membrane Binds Phase I clinical trials in MIA PaCa-2 (IC50 ≥ 80 μM); Marom et al., 1995;
compound isolated from association galectin patients with advanced H-ras-transformed EJ cells Aharonson et al., 1998; Egozi
willow tree, prokaryotic and PDAC and refractory (IC50 = 5–15 μM); et al., 1999; Weisz et al.,
eukaryotic organisms hematologic malignancies. Panc-1 1999; Gana-Weisz et al.,

S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57


Phase II clinical trial in (IC50 = 35 μM); 2002; Kloog & Cox, 2004;
patients with lung U87 Haklai et al., 2008; Chen
adenocarcinomas harboring (IC50 = 50 μM) et al., 2009; Goldberg et al.,
K-Ras mutations 2009; Riely et al., 2011;
Laheru et al., 2012; Baker &
Der, 2013; Mackenzie et al.,
2013; Badar et al., 2015
S-prenyl derivatives Phenolic acids Derived from salirasib Interference of Ras membrane Binds In vitro H-ras transformed EJ cells Aharonson et al., 1998
association galectin (IC50 = 10–N50 μM); Goldberg et al., 2009
Panc-1 (IC50 = 10–95 μM);
U87 (IC50 = 10–52 μM)
Carboxyl-modified Phenolic acids Derived from salirasib Interference of Ras membrane Binds In vivo — mice MIA PaCa-2 (IC50 = Mackenzie et al., 2013
derivatives, association galectin 44.1 ± 2.8 μM)
including PFTS
TLN-4601 Alkaloid Marine actinomycete Interference of Ras membrane Depletes Raf Phase I and II clinical trials MIA PaCa-2 (IC50 = 10 μM) Charan et al., 2004; Boufaied
bacterium, Micromonospora association in patients with GBM et al., 2010; Campbell et al.,
sp. 2010; Mason et al., 2010,
2012
Spermatinamine and Alkaloids Australian Verongida marine Interference of Ras membrane Bind ICMT In vitro • Spermatinamine – ICMT Buchanan et al., 2007, 2008;
aplysamine 6 sponge, Pseudoceratina sp. association (IC50 = 1.9 μM); K562 and Kottakota et al., 2012
HT29 (GI50 = 0.23 μM)
• Aplysamine 6 –ICMT
(IC50 = 14 μM)
Palmostatin B Lactone Derived from lipstatin, a Interference of Ras membrane Binds APT1 In vitro APT 1 (IC50 = 670 nM) Weibel et al., 1987; Dekker
compound isolated from the association et al., 2010; Cox, 2010;
sctinobacterium Dekker & Hedberg, 2011;
Streptomyces toxytricini Hernandez et al., 2013; Zhu
et al., 2014
SRJ09, SRJ10, and SRJ23 Terpenoids Derived from AGP, a Inhibition of Ras–GEF Bind Ras In vitro • SRJ09 – MCF-7 Jada et al., 2008; Hocker et al.,
compound isolated from an nucleotide exchange (GI50 = 4.2 μM); MDA-231 2013
Asian herb, Andrographis (GI50 = 5.0 μM); PC-3
paniculata (GI50 = 12.4 μM); HCT-116
(GI50 = 3.1 μM)
• SRJ10 – MCF-7
(GI50 = 5.3 μM); HCT-116
(GI50 = 4.7 μM)
• SRJ23 – MCF-7
(GI50 = 5.2 μM); MDA-231
(GI50 = 5.8 μM); PC-3
(GI50 = 6.6 μM); HCT-116
(GI50 = 3.8 μM)
SCH-53870, SCH-53239, • SCH-53,239 – • SCH-53,239 – guanine nu- Inhibition of Ras–GEF Bind Ras In vitro • SCH-53,870 and SCH-53,239 – Taveras et al., 1997; Palmioli,
and SCH-54292 nucleobase cleotide nucleotide exchange Nucleotide-exchange 2009
• SCH-54,292 – sugar • SCH-54,292 – glucose (IC50 = 0.5 μM)
• SCH-54,292 - Nucleotide
exchange (IC50 = 0.7 μM)
Compounds 1–10 derived Sugars D-arabinose Inhibition of Ras–GEF Bind Ras In vitro • Compound 1 – Nucleotide Peri et al., 2005, 2006
from D-arabinose nucleotide exchange exchange (IC50 = 76.0 μM)
• Compound 2 – Nucleotide
exchange (IC50 = 57.3 μM)
• Compound 3 – Nucleotide
exchange (IC50 = 35.5 μM)
• Compound 4 – Nucleotide
exchange (IC50 = 62.3 μM)
• Compounds 5–10 have no
clearly defined IC50 values
Glycosylated Compounds Sugars • Compound 2 — glucose Inhibition of Ras–GEF Bind Ras In vitro • Compound 2 – Nucleotide Sacco et al., 2011
2–4 • Compound 3 — N-acetyl nucleotide exchange exchange (IC50 = 80 μM)
glucosamine • Compound 3 – Nucleotide

S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57


• Compound 4 — lactose exchange (IC50 = 80 μM)
• Compound 4 – Nucleotide
exchange (IC50 = 100 μM)
Compounds 1–9 derived Carbohydrates Levoglucosenone (cellulose) Inhibition of Ras–GEF Bind Ras In vitro A2780 (IC50 = 3.1 to N40 μM); Müller et al., 2009
from levoglucosenone nucleotide exchange MCF-7 (IC50 = 5.5 to N40 μM);
A549 (IC50 = 15.5 to N40 μM;
HCT-116 (IC50 = 17 to N40 μM )
Pheophorbide a, • Pheophorbide a, Typhonium flagelliforme Inhibition of Ras–GEF Bind Ras In silico - Lai et al., 2009; Fatima &
pheophorbide a’, pheophorbide a’, (Keladi tikus) nucleotide exchange Yee, 2014
pyropheophorbide a, pyropheophorbide
hexadecanoic acid, a–alkaloids
1-hexadecene • Hexadecanoic
acid–carboxylic
acid
• Hexadecene–alkene
Rasfonin Lactone From fungi: Talaromyces Reduction of SOS1 (GEF) Is not clearly In vivo — mice Ba/F3-V12 Tomikawa et al., 2000;
sp. 3656-A1, Trichurus protein expression defined (IC50 = 0.37 μM); Panc-1 (IC50 Fujimoto et al., 2000;
terrophilus, and Doratomyces = 5.5 μM); BxPC-3 (IC50 = 10 Akiyama et al., 2003;
sp. μM) Fujimoto et al., 2005;
Xiao et al., 2014
Quercetin Flavonoid Red wine, many fruits, Downregulation of Ras Binds Raf or In vitro Ras expression – Ranelletti et al., 2000;
vegetables, and grains proteins MEK and Colo-320HSR and Psahoulia et al., 2007; Xavier
PI3K proteins Colo-205 et al., 2009; Slusarz et al.,
(IC50 = 0.5 μM) 2010; Deng et al., 2013; Song
et al., 2013

Abbreviations: AGP, Andrographolide; APT1, acyl protein thioesterase 1; GBM, Glioblastoma multiforme; GEF, Guanine nucleotide exchange factor; ICMT, Isoprenylcysteine carboxymethyltransferase; K-Ras, Kirsten-Ras; MEK, Mitogen-activated
protein/extracellular signal-regulated kinase kinase; PDAC, Pancreatic ductal adenocarcinoma; PFTS, Phospho-farnesylthiosalicylic acid; PI3K, Phosphoinositide-3-kinase; Raf, Rapidly accelerated fibrosarcoma; Ras, Ras sarcoma; REF, Rat embryo
fibroblasts; SOS1, Son of Sevenless 1.
Cancer Cell Lines: A549, lung; A2780, ovary; BxPC-3, pancreas; Colo-205, colon; Colo-320HSR, colon; HCT-116, colon; HT29, colon; K562, bone marrow; MCF-7, breast; MDA-231, breast; MIA PaCa-2, pancreas; Panc-1, pancreas; PC-3, prostate; U87,
brain.

53
54 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

salirasib metabolism, an equimolar concentration of PFTS led to higher For wild-type Ras to be activated, GDP must be removed from the
levels of salirasib and lower levels of salirasib glucuronide. Based on nucleotide-binding pocket by GEF, thus providing space for the intracel-
these findings, salirasib can be considered the primary contributor to lularly abundant GTP to bind to Ras. Whether oncogenic Ras, mutations
the high potency of PFTS. of which cause so-called constitutive activation, would ever return to its
The farnesyl group of salirasib is the core pharmacophore responsi- GDP-bound inactive state and require GEF for its reactivation remains to
ble for the reported anti-Ras activity of this compound. The farnesyl be elucidated. In a recent study, contrary to popular opinion, oncogenic
group acts by competing with GTP-bound H-Ras and K-Ras-4B proteins Ras was not found to be constitutively active and independent of GEF-
for binding to the binding domain on galectin at the plasma membrane. mediated activation (Huang et al., 2014). Although Ras mutations inter-
Notably, the length of the isoprenoid group on a compound determines fere with Ras–GAP interaction and impair the GTP hydrolysis of active
the inhibitory effect of this compound on Ras activity and Ras-mediated Ras, its intrinsic GTPase activity remains unaffected. Therefore, the
cell proliferation and survival (Aharonson et al., 1998). A compound GTP hydrolysis of active oncogenic Ras can still occur in the absence of
with a 10-carbon geranyl group was found to lack the inhibitory effect GAP, but more slowly, leading to prolonged Ras activation. Once onco-
found in compounds with a 15-carbon farnesyl group and a 20-carbon genic Ras returns to its GDP-bound inactive form due to its intrinsic
geranylgeranyl group (Aharonson et al., 1998). Although the mecha- GTPase activity, it relies on GEF for its activation.
nism of Ras inhibition by salirasib derivatives was not elucidated previ- To this end, we were able to confirm the feasibility of targeting Ras–
ously, we propose that they may demonstrate the same anti-Ras GEF interaction and GTP loading of Ras as a strategy to block oncogenic
mechanism as salirasib as they share a common 15-carbon farnesyl Ras signaling. As oncogenic Ras is activated in a GEF-dependent manner,
group. Structural modifications of salirasib, such as halogenation of the similar to wild-type Ras, interfering with Ras–GEF interaction and
benzene ring, esterification, or amidation of the carboxyl group of the inhibiting GTP loading of Ras remains an effective approach to target on-
thiosalicylic acid, generate new compounds with varying anti-Ras activ- cogenic Ras signaling. AGP and its benzylidene derivatives (SRJ09 and
ity in vitro and in vivo. Among the salirasib derivatives, amide deriva- SRJ23) were found to block the nucleotide exchange of wild-type and
tives showed improved efficacy compared with their parental oncogenic Ras by abrogating Ras–GEF interaction. Based on the SBDD
compound, although these derivatives retained the inhibitory selectivi- approach and ensemble docking, Hocker et al. (2013) reported Ras as
ty of salirasib toward the GTP-bound Ras protein. Therefore, amide de- a potential target of AGP and its benzylidene derivatives. AGP is a natu-
rivatives of salirasib may be used in place of salirasib as promising Ras ral compound with significant anticancer activity. The benzylidene de-
inhibitors for cancer therapy. rivatives of AGP exhibit excellent anticancer activity against several
TLN-4601 also shares the same farnesyl group with salirasib. This cancer cell lines. Along with in silico docking results, SRJ09 and SRJ23
compound was reported to reduce the GTP-bound K-Ras levels in addi- were shown to inhibit the EGF-induced activation of endogenous
tion to depleting the total Raf-1 protein in pancreatic cancer cells and wild-type Ras, indicating that SRJ09 and SRJ23 inhibited GEF-mediated
transformed cells, both of which express oncogenic K-Ras (Campbell Ras activation by directly binding to Ras. Prolonged treatment of cells
et al., 2010). We postulate that TLN-4601 inhibits Ras activity by inter- stably expressing exogenous oncogenic Ras with SRJ09 and SRJ23 re-
fering with its association to the plasma membrane. Similar to salirasib, duced the levels of GTP-bound oncogenic Ras. This showed that onco-
this compound may compete with GTP-bound H-Ras and K-Ras-4B pro- genic Ras slowly returned to its inactive state due to its intrinsic
teins for binding to galectin, thus increasing the unbound activated Ras GTPase activity, and AGP derivatives prevented the reactivation of onco-
that can be readily degraded in the cytoplasm. genic Ras by GEF. It is interesting to note that AGP derivatives selectively
The posttranslational modification of Ras, so as to transport Ras intra- showed low micromolar growth inhibitory activity against cancer cell
cellularly to the plasma membrane and embed it to the inner leaflet of lines harboring Ras mutations, in spite of unbiased docking of these
the plasma membrane, is regulated by several enzymes. These enzymes compounds into wild-type and oncogenic Ras in silico. Altogether, AGP
are alternative targets for blocking oncogenic Ras signaling via the pre- and its derivatives may serve as effective anti-Ras drugs with minimal
vention of Ras maturation in a subsequent GEF-mediated activation at adverse effects on the normal cells.
the plasma membrane. In this review, we described several natural Recently, a promising natural compound, rasfonin, was found to tar-
compounds that target these enzymes. Among these compounds, get oncogenic Ras signaling by downregulating the protein expression
spermatinamine emerged as the leading compound with significant inhi- of GEF. Rasfonin was expected to equally inhibit both signaling path-
bition on Ras activation. Although spermatinamine is structurally large, it ways mediated by wild-type and oncogenic Ras, considering that
can serve as an effective inhibitor of Ras activation because of its relative- these pathways are activated in a GEF-dependent manner. Neverthe-
ly strong inhibitory effect on the catalytic activity of ICMT and promising less, Xiao et al. (2014) reported that rasfonin was selective in inhibiting
sub-micromolar anticancer effect on leukemia and colon cancer cell lines. the growth of a PDAC cell line harboring oncogenic K-Ras over a cell line
Future studies that specifically focus on modifying the chemical structure harboring wild-type K-Ras. This finding is indeed interesting, as is the
of spermatinamine are needed to improve its drug likeness. Neverthe- molecular mechanism proposed to explain the greater dependency of
less, changing the chemical structure of spermatinamine without affect- RAS oncogene-driven cell proliferation on GEF. The inhibitory effect of
ing its inherent anticancer activity is challenging. rasfonin on RAS oncogene-driven cell growth, migration, and invasion
Palmostatin B led to sub-micromolar inhibition of APT1 activity was also shown to be confined to inhibiting the Ras–MAPK pathway.
in vitro, and its inhibitory effect on Ras signaling was proven in a cell Xiao et al. (2014) investigated the effect of rasfonin on several EGFR-
line that stably expressed oncogenic Ras (Dekker et al., 2010). The com- mediated pathways, among which the MAPK pathway showed signifi-
pound was found to prevent palmitoylated Ras from relocalizing from cant response to rasfonin treatment. These findings suggested that
the endomembranes to the plasma membrane, thus inhibiting GEF- rasfonin may serve as a potential inhibitor of Ras signaling with signifi-
dependent Ras activation. Thus, it can serve as an effective inhibitor of cant target specificity, thus overcoming the limitation of poor selectivity
Ras signaling. However, it is primarily limited by its poor substrate se- commonly encountered in drug discovery.
lectivity. The anti-depalmitoylation activity of palmostatin B targets a Quercetin is another promising natural candidate for the treatment
wide range of palmitoylated proteins in addition to Ras (Zeidman of RAS-driven cancers. Studies have shown that quercetin inhibits RAS
et al., 2009). These palmitoylated proteins are distributed in various tis- oncogene expression and promotes the degradation of Ras oncoprotein
sues, which are crucial for maintaining the functioning of an organism in primary colon tumors and human colon cancer cell lines. The antican-
(Zeidman et al., 2009). The poor selectivity poses a limitation to tests cer effect of quercetin on colorectal cancer is clear as Ras signaling is im-
in animal models. To overcome this limitation, palmostatin B can be plicated in cell proliferation and survival. Quercetin was also found to
modified to increase its selectivity toward oncogenic Ras, thus minimiz- inhibit cancer cell growth, survival, invasion, and migration via physical
ing its systemic toxicity. interaction with the Ras effector, thus blocking oncogenic Ras signaling.
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 55

Ranelletti et al. (2000) and Psahoulia et al. (2007) showed that the se- Brenke, R., Kozakov, D., Chuang, G. Y., Beglov, D., Hall, D., Landon, M. R., et al. (2009).
Fragment-based identification of druggable 'hot spots' of proteins using Fourier do-
lectivity of quercetin in inhibiting the growth of cancer cells harboring main correlation techniques. Bioinformatics 25(5), 621–627.
oncogenic Ras can serve as the basis for the future development of Brunell, D., Sagher, D., Kesaraju, S., Brot, N., & Weissbach, H. (2011). Studies on the metab-
anti-Ras drugs with reduced off-target effects. olism and biological activity of the epimers of sulindac. Drug Metab Dispos 39(6),
1014–1021.
Oncogenic Ras shows an impaired response to GAP-mediated GTP Buchanan, M. S., Carroll, A. R., Fechner, G. A., Boyle, A., Simpson, M. M., & Addepalli, R.
hydrolysis, leading to prolonged activation of oncogenic Ras. Neverthe- (2008). Aplysamine 6, an alkaloidal inhibitor of isoprenylcysteine carboxyl methyl-
less, it is yet to be examined whether the GTPase activity of oncogenic transferase from the sponge Pseudoceratina sp. J Nat Prod 71(6), 1066–1067.
Buchanan, M. S., Carroll, A. R., Fechner, G. A., Boyle, A., Simpson, M. M., Addepalli, R., et al.
Ras can be enhanced using a small molecule to compensate for its insen- (2007). Spermatinamine, the first natural product inhibitor of isoprenylcysteine
sitivity to GAP-mediated inactivation. Here, we propose a strategy ap- carboxylmethyltransferase, a new cancer target. Bioorg Med Chem Lett 17(24),
plying computer-assisted screening on natural compounds and their 6860–6863.
Campbell, P. M., Boufaied, N., Fiordalisi, J. J., Cox, A. D., Falardeau, P., Der, C. J., et al. (2010).
derivatives that readily dock into the active site of GTP hydrolysis on
TLN-4601 suppresses growth and induces apoptosis of pancreatic carcinoma cells
Ras, wherein one or more pharmacophores of the compounds are in- through inhibition of Ras–ERK MAPK signaling. J Mol Signal 5, 18.
volved in stabilizing the transition state of GTP hydrolysis. Alternatively, Chandra, A., Grecco, H. E., Pisupati, V., Perera, D., Cassidy, L., Skoulidis, F., et al. (2012). The
discovering a compound that can nullify the effect of Ras mutations GDI-like solubilizing factor PDEδ sustains the spatial organization and signaling of Ras
family proteins. Nat Cell Biol 14, 148–158.
and facilitate Ras–GAP interaction is also feasible for counteracting Chandrashekar, R., & Adams, P. D. (2013). Prospective development of small molecule tar-
prolonged Ras activation. gets to oncogenic Ras proteins. J Biophys 3, 207–211.
Finding an effective therapeutic agent for RAS-driven cancers is not Charan, R. D., Schlingmann, G., Janso, J., Bernan, V., Feng, X., & Carter, G. T. (2004).
Diazepinomicin, a new antimicrobial alkaloid from a marine Micromonospora sp.
impossible altogether, despite the challenges in drugging Ras. With J Nat Prod 67(8), 1431–1433.
the development of popular anticancer agents from natural compounds, Chen, Y. (2015). Beware of docking. Trends Pharmacol Sci 36(2), 78–95.
we believe that natural products represent a rich source of novel anti- Chen, L., Morrow, J. K., Tran, H. T., Phatak, S. S., Du-Cuny, L., & Zhang, S. (2012). From lap-
top to benchtop to bedside: Structure-based drug design on protein targets. Curr
cancer agents, particularly agents that target oncogenic Ras and its sig- Pharm Des 18(9), 1217–1239.
naling. Decades of research efforts and failures can inform future Chen, Z., Zheng, Z., Huang, J., Lai, Z., & Fan, B. (2009). Biosynthesis of salicylic acid in plants.
effective methods of drugging Ras as well as blocking its signaling. It is Plant Signal Behav 4(6), 493–496.
Cherfils, J., & Zeghouf, M. (2013). Regulation of small GTPases by GEFs, GAPs, and GDIs.
all the more imperative to discover agents that target oncogenic Ras
Physiol Rev 93, 269–309.
and its signaling. Cox, A. D. (2010). Protein localization: Can too much lipid glue stop Ras? Nat Chem Biol
6(7), 483–485.
Cox, A. D., Der, C. J., & Philips, M. R. (2015). Targeting RAS Membrane Association: Back to
Conflict of interest the future for anti-RAS drug discovery? Clin Cancer Res 21(8), 1819–1827.
Cox, A. D., Fesik, S. W., Kimmelman, A. C., Luo, J., & Der, C. J. (2014). Drugging the
undruggable RAS: Mission possible? Nat Rev Drug Discov 13, 828–851.
The authors declare that there are no conflicts of interest.
de Castro Carpeño, J., & Belda-Iniesta, C. (2013). KRAS mutant NSCLC, a new opportunity for
the synthetic lethality therapeutic approach. Transl Lung Cancer Res 2(2), 142–151.
de la Cruz, F. F., Gapp, B. V., & Nijman, S. M. B. (2015). Synthetic lethal vulnerabilities of
Acknowledgements
cancer. Annu Rev Pharmacol Toxicol 55, 513–531.
Dekker, F. J., & Hedberg, C. (2011). Small molecule inhibition of protein depalmitoylation
The authors would like to thank the Malaysian Ministry of Higher as a new approach towards downregulation of oncogenic Ras signaling. Bioorg Med
Education (MOHE) for funding projects through the Research Universi- Chem 19(4), 1376–1380.
Dekker, F. J., Rocks, O., Vartak, N., Menninger, S., Hedberg, C., Balamurugan, R., et al.
ty Grant Scheme (RUGS, 04-02-12-2017RU) and Fundamental Research (2010). Small-molecule inhibition of APT1 affects Ras localization and signaling. Nat
Grant Scheme (FRGS, 04-02-13-1324FR). Chem Biol 6, 449–456.
Delgado, M. D., Vaqué, J. P., Arozarena, I., Ló-Ilasaca, M. A., Martínez, C., Crespo, P., et al.
(2000). H-, K- and N-Ras inhibit myeloid leukemia cell proliferation by a p21WAF1-
References dependent mechanism. Oncogene 19, 783–790.
Deng, X., Song, H., Zhou, Y., Yuan, G., & Zheng, F. (2013). Effects of quercetin on the proliferation of
Aharonson, Z., Gana-Weisz, M., Varsano, T., Haklai, R., Marciano, D., & Kloog, Y. (1998). breast cancer cells and expression of surviving in vitro. Exp Ther Med 6(5), 1155–1158.
Stringent structural requirements for anti-Ras activity of S-prenyl analogues. Egozi, Y., Weisz, B., Gana-Weisz, M., Ben-Baruch, G., & Kloog, Y. (1999). Growth inhibition
Biochim Biophys Acta 1406(1), 40–50. of Ras-dependent tumors in nude mice by a potent ras-dislodging antagonist. Int J
Akiyama, K., Kawamoto, S., Fujimoto, H., & Ishibashi, M. (2003). Absolute stereochemistry Cancer 80(6), 911–918.
of TT-1 (rasfonin), an α-pyrone-containing natural product from a fungus, Trichurus Elad-Sfadia, G., Haklai, R., Balan, K., & Kloog, Y. (2004). Galectin-3 augments K-Ras activa-
terropilus. Tetrahedron Lett 44, 8427–8431. tion and triggers a Ras signal that attenuates ERK but not phosphoinositide 3-kinase
Almoguera, C., Shibata, D., Forrester, K., Martin, J., Arnheim, N., & Perucho, M. (1988). activity. J Biol Chem 279(33), 34922–34930.
Most human carcinoma of the exocrine pancreas contain mutant c-K-Ras genes. Elad-Sfadia, G., Haklai, R., Ballan, E., Gabius, H. J., & Kloog, Y. (2002). Galectin-1 augments
Cell 53, 549–554. Ras activation and diverts Ras signals of Raf-1 at the expense of phosphoinositide 3-
Badar, T., Cortes, J. E., Ravandi, F., O'Brien, S., Verstovsek, S., Garcia-Manero, G., et al. kinase. J Biol Chem 277(40), 37169–37175.
(2015). Phase I study of S-trans,trans-farnesylthiosalicylic acid (salirasib), a novel Fatima, A., & Yee, H. F. (2014). In silico screening of mutated K-Ras inhibitors from
oral RAS inhibitor in patients with refractory hematologic malignancies. Clin Malaysian Typhonium flagelliforme for non-small cell lung cancer. Adv Bioinformatics
Lymphoma Myeloma Leuk 15(7), 433–438. 2014. http://dx.doi.org/10.1155/2014/431696.
Baines, A. T., Xu, D., & Der, C. J. (2011). Inhibition of Ras for cancer treatment: the search Forbes, S. A., Bindal, N., Bamford, S., Cole, C., Kok, C. Y., Beare, D., et al. (2011). COSMIC:
continues. Future Med Chem 3, 1787–1808. Mining complete cancer genomes in the catalogue of somatic mutations in cancer.
Baker, N. M., & Der, C. J. (2013). Cancer: Drug for an ‘undruggable’ protein. Nature Nucleic Acids Res 39, D945–D950.
497(7451), 577–578. Fujimoto, H., Okamoto, Y., Sone, E., Maeda, S., Akiyama, K., & Ishibashi, M. (2005). Eleven
Barreiro, E. J., & Lima, M. E. (1992). The synthesis and anti-inflammatory properties of a new 2-pyrones from a fungi imperfecti, Trichurus terrophilus, found in a screening
new sulindac analogue synthesized from natural safrole. J Pharm Sci 81(12), study guided by immunomodulatory activity. Chem Pharm Bull 53(8), 923–929.
1219–1222. Fujimoto, H., Sone, E., Okuyama, E., & Ishibashi, M. (2000). Annual meeting of the pharma-
Bernards, A., & Settleman, J. (2004). GAP control: Regulating the regulators of small ceutical society of Japan. Abstr Pap 2, 68.
GTPases. Trends Cell Biol 14, 377–385. Gana-Weisz, M., Halaschek-Wiener, J., Jansen, B., Elad, G., Haklai, R., & Kloog, Y. (2002).
Berndt, N., Hamilton, A. D., & Sebti, S. M. (2011). Targeting protein prenylation for cancer The Ras inhibitor S-trans, trans-farnesylthiosalicylic acid chemosensitizes human
therapy. Nat Rev Cancer 11, 775–791. tumor cells without causing resistance. Clin Cancer Res 8(2), 555–565.
Blum, R., Jacob-Hirsch, J., Rechavi, G., & Kloog, Y. (2006). Suppression of survivin expres- Gideon, P., John, J., Frech, M., Lautwein, A., Clark, R., Scheffler, J. E., & Wittinghofer, A.
sion in glioblastoma cells by the Ras inhibitor farnesylthiosalicylic acid promotes (1992). Mutational and kinetic analyses of the GTPase-activating protein (GAP)-
caspase-dependent apoptosis. Mol Cancer Ther 5(9), 2337–2347. p21 interaction: The C-terminal domain of GAP is not sufficient for full activity. Mol
Boriack-Sjodin, P. A., Margarit, S. M., Bar-Sagi, D., & Kuriyan, J. (1998). The structural basis Cell Biol 12(5), 2050–2056.
of the activation of Ras by SOS. Nature 394(6691), 337–343. Goldberg, L., & Kloog, Y. (2006). A ras inhibitor tilts the balance between Rac and Rho and
Bos, J. L., Rehmann, H., & Wittinghofer, A. (2007). GEFs and GAPs: Critical elements in the blocks phosphatidylinositol 3-kinase-dependent glioblastoma cell migration. Cancer
control of small G Proteins. Cell 129(5), 865–877. Res 66(24), 11709–11717.
Boufaied, N., Wioland, M. A., Falardeau, P., & Gourdeau, H. (2010). TLN-4601, a novel an- Goldberg, L., Haklai, R., Bauer, V., Heiss, A., & Kloog, Y. (2009). New derivatives of
ticancer agent, inhibits Ras signalling post Ras prenylation and before MEK activation. farnesylthiosalicylic acid (salirasib) for cancer treatment: Farnesylthiosalicylamide
Anticancer Drugs 21(5), 543–552. inhibits tumor growth in nude mice models. J Med Chem 52, 197–205.
56 S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57

Grant, B. J., Lukman, S., Hocker, H. J., Sayyah, J., Brown, J. H., McCammon, J. A., et al. (2011). pancreatic cancer: a phase III trial of the National Cancer Institute of Canada Clinical
Novel allosteric sites on ras for lead generation. PLoS One 6(10), e25711. Trials Group. J Clin Oncol 25(15), 1960–1966.
Greer, J., Erickson, J. W., Baldwin, J. J., & Varney, M. D. (1994). Application of the three- Müller, C., Frau, M. A. G., Ballinari, D., Colombo, S., Bitto, A., Martegani, E., et al. (2009). De-
dimensional structures of protein target molecules in structure-based drug design. sign, synthesis, and biological evaluation of levoglucosenone-derived Ras activation
J Med Chem 37(8), 1035–1054. inhibitors. ChemMedChem 4, 524–528.
Guha, A., Feldkamp, M. M., Lau, N., Boss, G., & Pawson, A. (1997). Proliferation of human Newman, D. J., & Cragg, G. M. (2012). Natural products as sources of new drugs over the
maglinant astrocytomas is dependent on Ras activation. Oncogene 15(23), 30 years from 1981 to 2010. J Nat Prod 75(3), 311–335.
2755–2765. Noonan, T., Brown, N., Dudycz, L., & Wright, G. (1991). Interaction of GTP derivatives with
Guo, W., Wu, S., Liu, J., & Fang, B. (2008). Identification of a small molecule with synthetic cellular and oncogenic ras-p21 proteins. J Med Chem 34(4), 1302–1307.
lethality for K-RAS and protein kinase C iota. Cancer Res 68, 7403–7408. Ostrem, J. M., Peters, U., Sos, M. L., Wells, J. A., & Shokat, K. M. (2013). K-Ras(G12C) inhibitors
Haklai, R., Elad-Sfadia, G., Egozi, Y., & Kloog, Y. (2008). Orally administered FTS (salirasib) allosterically control GTP affinity and effector interactions. Nature 503, 548–551.
inhibits human pancreatic tumor growth in nude mice. Cancer Chemother Pharmacol Palmioli, A. (2009) Synthesis and biological characterization of new pharmacologically active
61(1), 89–96. molecules derived from natural compounds (published doctoral thesis). University of
Hanson, J. R. (2003). The classes of natural products and their isolation. In E. W. Abel (Ed.), Milano-Bicocca, Milano, Italy.
Natural products: The secondary metabolites (pp. 1–34). Cambridge, United Kingdom: Palmioli, A., Sacco, E., Abraham, S., Thomas, C. J., Di Domizio, A., De Gioia, L., et al. (2009).
Royal Society of Chemistry Publishing. First experimental identification of Ras-inhibitor binding interface using a water-
Hernandez, J. L., Majmudar, J. D., & Martin, B. R. (2013). Profiling and inhibiting reversible soluble Ras ligand. Bioorg Med Chem Lett 19, 4217–4222.
palmitoylation. Curr Opin Chem Biol 17(1), 20–26. Peri, F., Airoldi, C., Colombo, S., Mari, S., Jiménez-Barbero, J., Martegani, E., et al. (2006).
Herrmann, C. (2003). Ras-effector interactions: After one decade. Curr Opin Struct Biol 13, Sugar-derived Ras inhibitors: Group epitope mapping by NMR spectroscopy and bio-
122–129. logical evaluation. Eur J Org Chem, 3707–3720.
Herrmann, C., Block, C., Geisen, C., Haas, K., Weber, C., & Winde, G. (1998). Sulindac sulfide Peri, F., Airoldi, C., Colombo, S., Martegani, E., van Neuren, S., Stein, M., et al. (2005). De-
inhibits Ras signaling. Oncogene 17, 1769–1776. sign, synthesis and biological evaluation of sugar-derived Ras inhibitors.
Hocker, H. J., Cho, K., Chen, C. K., Rambahal, N., Sagineedu, S. R., Shaari, K., et al. (2013). Chembiochem 6, 1839–1848.
Andrographolide derivatives inhibit guanine nucleotide exchange and abrogate on- Plowman, S. J., & Hancock, J. F. (2005). Ras signaling from plasma membrane and
cogenic Ras function. Proc Natl Acad Sci 110, 10201–10206. endomembrane microdomains. Biochim Biophys Acta 1746, 274–283.
Huang, H., Daniluk, J., Liu, Y., Chu, J., Li, Z., Ji, B., & Logsdon, C. D. (2014). Oncogenic K-Ras Prime, version 3.4, Schrödinger, LLC, New York, NY, 2013.
requires activation for enhanced activity. Oncogene 33, 532–535. Prior, I. A., & Hancock, J. F. (2012). Ras trafficking, localization and compartmentalized sig-
Jada, S. R., Matthews, C., Saad, M. S., Hamzah, A. S., Lajis, N. H., Stevens, M. F. G., et al. (2008). nalling. Semin Cell Dev Biol 23(2), 145–153.
Benzylidene derivatives of andrographolide inhibit growth of breast and colon cancer Prior, I. A., Lewis, P. D., & Mattos, C. (2012). A comprehensive survey of Ras mutations in
cells in vitro by inducing G1 arrest and apoptosis. Br J Pharmacol 155, 641–654. cancer. Cancer Res 72(10), 2457–2467.
John, J., Rensland, H., Schlichting, I., Vetter, I., Borasio, G. D., Goody, R. S., et al. (1993). Ki- Psahoulia, F. H., Moumtzi, S., Roberts, M. L., Sasazuki, T., Shirasawa, S., & Pintzas, A. (2007).
netic and structural analysis of the Mg(2+)-binding site of the guanine nucleotide- Quercetin mediates preferential degradation of oncogenic Ras and causes autophagy
binding protein p21H-ras. J Biol Chem 268, 923–929. in H-RAS-transformed human colon cells. Carcinogenesis 28(5), 1021–1031.
Kapoor, A., & Travesset, A. (2015). Differential dynamics of RAS isoforms in GDP- and Ranelletti, F. O., Maggiano, N., Serra, F. G., Ricci, R., Larocca, L. M., Lanza, P., et al. (2000).
GTP-bound states. Proteins 83(6), 1091–1106. Quercetin inhibits p21-RAS expression in human colon cancer cell lines and in prima-
Kloog, Y., & Cox, A. D. (2004). Prenyl-binding domains: potential targets for Ras inhibitors ry colorectal tumors. Int J Cancer 85(3), 438–445.
and anti-cancer drugs. Semin Cancer Biol 14(4), 253–261. Rees, D. C., Congreve, M., Murray, C. W., & Carr, R. (2004). Fragment based lead discovery.
Konstantinopoulos, P. A., Karamouzis, M. V., & Papavassiliou, A. G. (2007). Post- Nat Rev Drug Discov 3, 660–672.
translational modifications and regulation of the RAS superfamily of GTPases as anti- Riely, G. J., Johnson, M. L., Medina, C., Rizvi, N. A., Miller, V. A., & Kris, M. G. (2011). A phase
cancer targets. Nat Rev Drug Discov 6, 541–555. II trial of salirasib in patients with lung adenocarcinomas with KRAS mutations.
Kottakota, S. K., Benton, M., Evangelopoulos, D., Guzman, J. D., Bhakta, S., & McHugh, T. D. J Thorac Oncol 6(8), 1435–1437.
(2012). Versatile routes to marine sponge metabolites through benzylidene Riely, G. J., Marks, J., & Pao, W. (2009). KRAS mutations in non-small cell lung cancer. Proc
rhodanines. Org Lett 14(24), 6310–6313. Am Thorac Soc 6, 201–205.
Kozakov, D., Hall, D. R., Jehle, S., Luo, L., Ochiana, S. O., Jone, E. V., et al. (2015). Ligand de- Rocks, O., Peyker, A., Kahms, M., Verveer, P. J., Koerner, C., & Lumbierres, M. (2005). An ac-
construction: Why some fragment binding positions are conserved and others are ylation cycle regulates localization and activity of palmitoylated ras isoforms. Science
not. Proc Natl Acad Sci 112(28), E2585–E2594. 307(5716), 1746–1752.
Laghi, L., Orbetegli, O., Bianchi, P., Zerbi, A., Di Carlo, V., Boland, C. R., et al. (2002). Com- Rodriguez-Viciana, P., Warne, P. H., Dhand, R., Vanhaesebroeck, B., Gout, I., Fry, M. J., et al.
mon occurrence of multiple K-RAS mutations in pancreatic cancers with associated (1994). Phosphatidylinositol-3-OH kinase as a direct target of Ras. Nature 370(6490),
precursor lesions and in biliary cancers. Oncogene 21(27), 4301–4306. 527–532.
Laheru, D., Shah, P., Rajeshkumar, N. V., McAllister, F., Taylor, G., & Goldsweig, H. (2012). Sacco, E., Abraham, S. J., Palmioli, A., Damore, G., Bragna, A., Mazzoleni, E., et al. (2011).
Integrated preclinical and clinical development of S-trans, trans-farnesylthiosalicylic Binding properties and biological characterization of new sugar-derived ligands.
acid (FTS, salirasib) in pancreatic cancer. Invest New Drugs 30, 2391–2399. Med Chem Commun. http://dx.doi.org/10.1039/c0md00264j.
Lai, C. S., Mas, R. H., Nair, N. K., Manson, S. M., & Navaratnam, V. (2009). Chemical constit- Scheffzek, K., Ahmadian, M. R., Kabsch, W., Wiesmuller, L., Lautwen, A., Schmitz, F., et al.
uents and in vitro anticancer activity of Typhonium flagelliforme (Araceae). (1997). The Ras-RasGAP complex: structural basis for GTPase activation and its loss
J Ethnopharmacol 127(2), 486–494. in oncogenic Ras mutants. Science 277(5324), 333–338.
Ledford, H. (2015). Cancer: The Ras renaissance. Nature 520(7547), 278–280. Scherer, A., John, J., Linke, R., Goody, R. S., Wittinghofer, A., Pao, E. F., et al. (1989). Crystal-
Lee, K. W., Kang, N. J., Heo, Y., Rogozin, E. A., Angelo, P., Hwang, M. K., et al. (2008). Raf and lization and preliminary X-ray analysis of the human c-H-Ras-oncogene product p21
MEK Protein kinases are direct molecular targets for the chemopreventive effect of complexed with GTP analogues. J Mol Biol 206, 257–259.
quercetin, a major flavonol in red wine. Cancer Res 68(3), 946–955. Schmidt, A., & Hall, A. (2002). Guanine nucleotide exchange factors for Rho GTPases: turn-
Lim, K., & Counter, C. M. (2005). Reduction in the requirement of oncogenic Ras signaling ing on the switch. Genes Dev 16, 1587–1609.
to activation of PI3K/AKT pathway during tumor maintenance. Cancer Cell 8, Sever, R., & Brugge, J. S. (2015). Signal transduction in cancer. Cold Spring Harb Perspect
381–392. Med 5(4). http://dx.doi.org/10.1101/cshperspect.a006098.
Lobell, R. B., Omer, C. A., Abrams, M. T., Bhimnathwala, H. G., Brucker, M. J., Buser, C. A., Slusarz, A., Shenouda, N. S., Sakla, M. S., Drenkhahn, S. K., Narula, A. S., MacDonald, R. S.,
et al. (2001). Evaluation of farnesyl:protein transferase and geranylgeranyl:protein et al. (2010). Common botanical compounds inhibit the hedgehog signalling pathway
transferase inhibitor combinations in preclinical models. Cancer Res 61, 8758–8768. in prostate cancer. Cancer Res 70(8), 3382–3390.
Lodish, H., Berk, A., Zipursky, S. L., Matsudaira, P., Baltimore, D., & Darnell, J. (2000). Song, N. R., Chung, M., Kang, N. J., Seo, S. G., Jang, T. S., Lee, H. J., et al. (2013). Quercetin
Section 20.4, receptor tyrosine kinases and Ras. molecular cell biology (4th ed.). New suppresses invasion and migration of H-Ras-transformed MCF10A human epithelial
York: W. H. Freeman. cells by inhibiting phosphatidylinositol 3-kinase. Food Chem 142, 66–71.
Mackenzie, G. G., Bartels, L. E., Xie, G., Papayannis, I., Alston, N., Alston, N., et al. (2013). A Spiegel, J., Cromm, P. M., Zimmermann, G., Grossmann, T. N., & Waldmann, H. (2014). Direct
novel Ras inhibitor (MDC-1016) reduces human pancreatic tumor growth in mice. targeting of rab-GTPase-effector interactions. Angew Chem Int Ed Engl 53, 2498–2503.
Neoplasia 15, 1184–1195. Spoerner, M., Graf, T., Konig, B., & Kalbitzer, H. R. (2005). A novel mechanism for the mod-
Marom, M., Haklai, R., Ben-Baruch, G., Marciano, D., Egozi, Y., & Kloog, Y. (1995). Selective ulation of the ras-effector interaction by small molecules. Biochem Biophys Res
inhibition of Ras-dependent cell growth by farnesylthiosalisylic acid. J Biol Chem Commun 334, 709–713.
270(38), 22263–22270. Sun, Q., Burke, J. P., Phan, J., Burns, M. C., Olejniczak, E. T., Waterson, A. G., et al. (2012).
Mason, W. P., Belanger, K., Nicholas, G., Vallières, I., Mathieu, D., Desjardins, A., et al. Discovery of small molecules that bind to K-Ras and inhibit SOS-mediated activation.
(2010). A phase II trial of TLN-4601 in patients with glioblastoma multiforme Angew Chem Int Ed Engl 51, 6140–6143.
(GBM) at first progression. J Clin Oncol 28, 2094. Taveras, A. G., Remiszewski, S. W., Doll, R. J., Cesarz, D., Huang, E. C., Kirschmeier, P., et al.
Mason, W. P., Belanger, K., Nicholas, G., Vallières, I., Mathieu, D., & Kavan, P. (2012). A (1997). Ras oncoprotein inhibitors: the discovery of potent, ras nucleotide exchange
phase II study of the Ras-MAPK signaling pathway inhibitor TLN-4601 in patients inhibitors and the structural determination of a drug-protein complex. Bioorg Med
with glioblastoma at first progression. J Neurooncol 107(2), 343–349. Chem 5(1), 125–133.
Maurer, T., Garrenton, L. S., Oh, A., Pitts, K., Anderson, D. J., Skelton, N. J., et al. (2012). Taylor, M. T., Lawson, K. R., Ignatenko, N. A., Marek, S. E., Stringer, D. E., Skovan, B. A., &
Small-molecule ligands bind to a distinct pocket in RAS and inhibit SOS-mediated nu- Gerner, E. W. (2000). Sulindac sulfone inhibits K-ras-dependent cyclooxygenase-2
cleotide exchange activity. Proc Natl Acad Sci 109, 5299–5304. expression in human colon cancer cells. Cancer Res 60(23), 6607–6610.
Meder, D., & Simons, K. (2005). Ras on the roundabout. Science 307(5716), 1731–1733. Thompson, H. J., Jiang, C., Lu, J., Mehta, R. G., Piazza, G. A., Paranka, N. S., et al. (1997).
Moore, M. J., Goldstein, D., Hamm, J., Figer, A., Hecht, J. R., Gallinger, S., et al. (2007). Erlo- Sulfone metabolite of sulindac inhibits mammary carcinogenesis. Cancer Res 57(2),
tinib plus gemcitabine compared with gemcitabine alone in patients with advanced 267–271.
S.Y. Quah et al. / Pharmacology & Therapeutics 162 (2016) 35–57 57

Tomikawa, T., Shin-Ya, K., Furihata, K., Kinoshita, T., Miyajima, A., Seto, H., et al. (2000). Williams, C. S., Goldman, A. P., Sheng, H., Morrow, J. D., & DuBois, R. N. (1999). Sulindac
Rasfonin, a new apoptosis inducer in ras-dependent cells from Talaromyces sp. sulfide, but not sulindac sulfone, inhibits colorectal cancer growth. Neoplasia 1(2),
J Antibiot 53, 848–850. 170–176.
Valkov, E., Sharpe, T., Marsh, M., Greive, S., & Hyvönen, M. (2012). Targeting protein– Wright, L. P., Court, H., Mor, A., Ahearn, I. M., Casey, P. J., & Philips, M. R. (2009). Topology
protein interactions and fragment-based drug discovery. Top Curr Chem 317, of mammalian isoprenylcysteine carboxyl methyltransferase determined in live cells
145–179. with a fluorescent probe. Mol Cell Biol 29(7), 1826–1833.
Vetter, I. R., & Wittinghofer, A. (2001). The guanine nucleotide-binding switch in three di- Xavier, C. P. R., Lima, C. F., Preto, A., Seruca, R., Fernandes-Ferreira, M., & Pereira-Wilson, C.
mensions. Science 294, 1299–1304. (2009). Luteolin, quercetin and ursolic acid are potent inhibitors of proliferation and
Vivanco, I., & Sawyers, C. L. (2002). The phosphatidylinositol 3-kinase AKT pathway in inducers of apoptosis in both KRAS and BRAF mutated human colorectal cancer cells.
human cancer. Nat Rev Cancer 2(7), 489–501. Cancer Lett 281, 162–170.
Walker, E. H., Pacold, M. E., Perisic, O., Stephens, L., Hawkins, P. T., Wymann, M. P., et al. Xiao, Z., Li, L., Li, Y., Zhou, W., Cheng, J., Liu, F., et al. (2014). Rasfonin, a novel 2-pyrone de-
(2000). Structural determinants of phosphoinositide 3-kinase inhibition by rivative, induces RAS-mutated Panc-1 pancreatic tumor cell death in nude mice. Cell
wortmannin, LY294002, quercetin, myricetin, and staurosporine. Mol Cell 6, 909–919. Death Dis 5. http://dx.doi.org/10.1038/cddis.2014.213.
Wang, Y., Kaiser, C. E., Frett, B., & Li, H. Y. (2013). Targeting mutant KRAS for anticancer Zartler, E. R., & Mo, H. (2007). Practical aspects of NMR-based fragment discovery. Curr
therapeutics: A review of novel small molecule modulators. J Med Chem 56(13), Top Med Chem 7, 1592–1599.
5219–5230. Zeidman, R., Jackson, C. S., & Magee, Al (2009). Protein acyl thiotransferases (review). Mol
Weibel, E. K., Hadvary, P., Hochuli, E., Kupfer, E., & Lengsfeld, H. (1987). Lipstatin, an inhib- Membr Biol 26(1), 32–41.
itor pancreatic lipase, produced by Streptomyces toxytricini. I. Producing organism, Zhang, F. L., & Casey, P. J. (1996). Protein prenylation: Molecular mechanisms and func-
fermentation, isolation and biological activity. J Antibiot 40(8), 1081–1085. tional consequences. Annu Rev Biochem 65, 241–269.
Weisz, B., Giehl, K., Gana-Weisz, M., Egozi, Y., Ben-Baruch, G., & Marciano, D. (1999). A Zhu, T., Wang, L., Wang, W., Hu, Z., Yu, M., & Wang, K. (2014). Enhanced production of
new functional Ras antagonist inhibits human pancreatic tumour growth in nude lipstatin from Streptomyces toxytricini by optimizing fermentation conditions and
mice. Oncogene 18(16), 2579–2588. medium. J Gen Appl Microbiol 60(3), 106–111.
Wennerberg, K., Rossman, K. L., & Der, C. J. (2005). The Ras superfamily at a glance. J Cell
Sci 118(Pt 5), 843–846.
Whyte, D. B., Kirschmeier, P., Hockenberry, T. N., Nunez-Oliva, I., James, L., Catino, J. J., et al.
(1997). K-Ras and N-Ras are geranylgeranylated in cells treated with farnesyl protein
transferase inhibitors. J Biol Chem 272, 14459–14464.

You might also like