Feeds:
Posts
Comments

Posts Tagged ‘Pancreatic cancer’

New Mutant KRAS Inhibitors Are Showing Promise in Cancer Clinical Trials: Hope For the Once ‘Undruggable’ Target, Volume 2 (Volume Two: Latest in Genomics Methodologies for Therapeutics: Gene Editing, NGS and BioInformatics, Simulations and the Genome Ontology), Part 1: Next Generation Sequencing (NGS)

New Mutant KRAS Inhibitors Are Showing Promise in Cancer Clinical Trials: Hope For the Once ‘Undruggable’ Target

Curator: Stephen J. Williams, Ph.D.

UPDATED 10/2/2021

A Newly Identified Mutant RAS G12C  GTPase Activating Protein (GAP) may lead to discovery of new class of RAS inhibitors 

From the journal Science in Filling in the GAPs in understanding RAS: A newly identified regulator increases the efficacy of a new class of targeted anti-RAS drugs.

Source:

SCIENCE•8 Oct 2021•Vol 374Issue 6564•pp. 152-153•DOI: 10.1126/science.abl3639

    According to canonical thinking, mutationally activated RAS (constitutive activation) is insensitive to GTPase-activating proteins (GAPs).  GAPs accelerated the conversion of GTP bound to RAS to GDP, a critical step in the inactivation of RAS signaling cycle.  However a small molecule has just been FDA approved, sotorasib, in cancer cells by binding only the GDP bound KRAS G12C mutant. A few non small cell lung cancers (NSCLC) are resistant to another such inhibitor, adagrasib.  In the same Science issue, another paper by Li et. al. explains the potential that a GAP, RGS3, may play in this conundrum and demonstrates that certain other inhibitors of the RAS cycle, mainly the GEF (guanine exchange factor) SOS1, in combination with MEK inhibitors may circumvent this resistance.

Inhibitors of mutant KRAS
The KRASG12C (Gly12→Cys) mutant is refractory to canonical GAPs, p120 RASGAP and neurofibromin, but not RGS3, which promotes GTP hydrolysis. The resulting GDP-bound KRASG12C is an anticancer drug target. Combination with inhibitors of SOS1 or with inhibitors of downstream signaling may further improve efficacy.
GRAPHIC: C. BICKEL/SCIENCE
 
First a discussion of the RAS signalling cycle is shown below in a good review of RAS activation and signal termination.
 
PMCID: PMC3124093
NIHMSID: NIHMS294865
PMID: 21102635
Ras superfamily GEFs and GAPs: validated and tractable targets for cancer therapy?

Abstract

There is now considerable and increasing evidence for a causal role of aberrant activity of the Ras superfamily of small GTPases in human cancers. These GTPases act as GDP-GTP-regulated binary switches that control many fundamental cellular processes. A common mechanism of GTPase deregulation in cancer is the deregulated expression and/or activity of their regulatory proteins, guanine nucleotide exchange factors (GEFs) that promote formation of the active GTP-bound state and GTPase activating proteins (GAPs) that return the GTPase to its GDP-bound inactive state. We assess the association of GEFs and GAPs with cancer and their druggability for cancer therapeutics.

Box 1

Ras superfamily of small GTPases

The human Ras superfamily comprised of over 150 members which is divided into five major branches on the basis of sequence and functional similarities. In addition to the three Ras isoforms, other members of the Ras family with important roles in cancer include Rheb and Ral proteins. The ~20 kDa core G domain (corresponding to Ras residues 4–166) is conserved among all Ras superfamily proteins and is involved in GTP binding and hydrolysis. This domain is comprised of five conserved guanine nucleotide consensus sequence elements (Ras residue numbering) involved in binding phosphate/Mg2+ (PM) or the guanine base (G). The switch I (Ras residues 30–38) and II (59–76) regions change in conformation during GDP-GTP cycling and contribute to preferential effector binding to the GTP-bound state and the core effector domain (E; Ras residues 32–40). Ras and Rho family proteins have additional C-terminal hypervariable (HV) sequences that commonly terminate with a CAAX motif that signals for farnesyl or geranylgeranyl isoprenoid addition to the cysteine residue, proteolytic removal of the AAX residues and carboxylmethylation of the prenylated cysteine. Some are modified additionally by a palmitate fatty acid to cysteine residues in the HV sequence that contributes to membrane association. Rab proteins also contain a C-terminal HV region that terminates with cysteine-containing motifs that are modified by addition of geranylgeranyl lipids, with some undergoing carboxylmethylation. Arf family proteins are characterized by an N-terminal extension involved in membrane interaction, with some cotranslationally modified by addition of a myristate fatty acid. Ran is not lipid modified but contains a C-terminal extension that is essential for function. Rho proteins are characterized by an up to 13 amino acid “Rho insert” sequence positioned between Ras residues 122 and 123 involved in effector regulation.

 

An external file that holds a picture, illustration, etc.
Object name is nihms294865u1a.jpg

 

The GDP-GTP cycle

Ras superfamily proteins possess intrinsic guanine nucleotide exchange and GTP hydrolysis activities. However, these activities are too low to allow efficient and rapid cycling between their active GTP-bound and inactive GDP-bound states. GEFs and GAPs accelerate and regulate these intrinsic activities. Members of the different branches of the superfamily are regulated by GEFs and GAPs with structurally distinct catalytic domains. Here we have utilized the Rho family as an example to illustrate the complexity of this process, where multiple GEFs and GAPs may regulate one specific GTPase. For the 20 human Rho GTPases there are 83 GEFs and 67 GAPs and a subset of Rho GTPases are not likely regulated by GEFs and GAPs (e.g., Rnd3/RhoE). Rho GTPases are activated by distinct RhoGEF families. Dbl family RhoGEFs (68) possesses a tandem Dbl homology (DH) catalytic and pleckstrin homology (PH) regulatory domain topology. DOCK family RhoGEFs (11) are characterized by two regions of high sequence conservation that are designated Dock-homology region regulatory DHR-1 and catalytic DHR-2 domains. Two other RhoGEFs have been described (SWAP70 and SLAT) contain a PH but no DH domain (2) and smgGDS (1) is an unusual GEF in that it functions as a GEF for some Rho as well as non-Rho family GTPases. At least 24 Dbl RhoGEFs have been reported to activate RhoA. Rho (and Rab) GTPases are also controlled by a third class of regulatory proteins, Rho dissociation inhibitors (RhoGDI) (of which there are 3) whose main function involves regulation of Rho GTPase membrane association by masking the isoprenoid group.

 

An external file that holds a picture, illustration, etc.
Object name is nihms294865u2.jpg

So GEFs like SOS 1,2,3 are responsible for activation of the G Protien RAS upon Receptor Tyrosine Kinase (RTK) activation by a multitude of growth factors.  The GAPs are responsible for termination of the activated RAS cycle and now RAS, in its normal unmuted version, can now be activated again.  The G12C mutant keeps RAS in its activated state as GAPs cannot fascilitate the hydroylsis of GTP.  

Li et al. (in this issue) theorized that there must be cellular factors that stimulate formation of GDP-bound KRAS G12C, enabling its vulnerability to KRAS G12C inhibitors, and discovered a regulator of G protein signaling 3 (RGS3), which had GAP activity previously for heterotrimeric G proteins GαI and Gαq with unexpected GAP activity for the KRAS G12C mutant.   They also found the fully activated GTP bound KRAS G12C is dependent on RTK-mediated activation of SOS1.  This suggested that SOS1 inhibitors with MEK inhibitors could be effective against mutant KRAS.

The paper by Li is summarized below:

The G protein signaling regulator RGS3 enhances the GTPase activity of KRAS

SCIENCE•8 Oct 2021•Vol 374Issue 6564•pp. 197-201•DOI: 10.1126/science.abf1730

Abstract

Recently reported to be effective in patients with lung cancer, KRASG12C inhibitors bind to the inactive, or guanosine diphosphate (GDP)–bound, state of the oncoprotein and require guanosine triphosphate (GTP) hydrolysis for inhibition. However, KRAS mutations prevent the catalytic arginine of GTPase-activating proteins (GAPs) from enhancing an otherwise slow hydrolysis rate. If KRAS mutants are indeed insensitive to GAPs, it is unclear how KRASG12C hydrolyzes sufficient GTP to allow inactive state–selective inhibition. Here, we show that RGS3, a GAP previously known for regulating G protein–coupled receptors, can also enhance the GTPase activity of mutant and wild-type KRAS proteins. Our study reveals an unexpected mechanism that inactivates KRAS and explains the vulnerability to emerging clinically effective therapeutics.

UPDATED 09/26/2021

The KRAS G12C Inhibitor MRTX849 Provides Insight toward Therapeutic Susceptibility of KRAS-Mutant Cancers in Mouse Models and Patients

Source: Hallin J, Engstrom LD, Hargis L, Calinisan A, Aranda R, Briere DM, Sudhakar N, Bowcut V, Baer BR, Ballard JA, Burkard MR, Fell JB, Fischer JP, Vigers GP, Xue Y, Gatto S, Fernandez-Banet J, Pavlicek A, Velastagui K, Chao RC, Barton J, Pierobon M, Baldelli E, Patricoin EF 3rd, Cassidy DP, Marx MA, Rybkin II, Johnson ML, Ou SI, Lito P, Papadopoulos KP, Jänne PA, Olson P, Christensen JG. The KRASG12C Inhibitor MRTX849 Provides Insight toward Therapeutic Susceptibility of KRAS-Mutant Cancers in Mouse Models and Patients. Cancer Discov. 2020 Jan;10(1):54-71. doi: 10.1158/2159-8290.CD-19-1167. Epub 2019 Oct 28. PMID: 31658955; PMCID: PMC6954325.

Abstract

Despite decades of research, efforts to directly target KRAS have been challenging. MRTX849 was identified as a potent, selective, and covalent KRASG12C inhibitor that exhibits favorable drug-like properties, selectively modifies mutant cysteine 12 in GDP-bound KRASG12C, and inhibits KRAS-dependent signaling. MRTX849 demonstrated pronounced tumor regression in 17 of 26 (65%) KRASG12C-positive cell line- and patient-derived xenograft models from multiple tumor types, and objective responses have been observed in patients with KRASG12C-positive lung and colon adenocarcinomas. Comprehensive pharmacodynamic and pharmacogenomic profiling in sensitive and partially resistant nonclinical models identified mechanisms implicated in limiting antitumor activity including KRAS nucleotide cycling and pathways that induce feedback reactivation and/or bypass KRAS dependence. These factors included activation of receptor tyrosine kinases (RTK), bypass of KRAS dependence, and genetic dysregulation of cell cycle. Combinations of MRTX849 with agents that target RTKs, mTOR, or cell cycle demonstrated enhanced response and marked tumor regression in several tumor models, including MRTX849-refractory models. SIGNIFICANCE: The discovery of MRTX849 provides a long-awaited opportunity to selectively target KRASG12C in patients. The in-depth characterization of MRTX849 activity, elucidation of response and resistance mechanisms, and identification of effective combinations provide new insight toward KRAS dependence and the rational development of this class of agents.

Introduction

KRAS is one of the most frequently mutated oncogenes in cancer; however, efforts to directly target KRAS have been largely unsuccessful due to its high affinity for GTP/GDP and the lack of a clear binding pocket (1–4). More recently, compounds were identified that covalently bind to KRASG12C at the cysteine 12 residue, lock the protein in its inactive GDP-bound conformation, inhibit KRAS-dependent signaling, and elicit antitumor responses in tumor models (5–7). Advances on early findings demonstrated that the binding pocket under the switch II region was exploitable for drug discovery, culminating in the identification of more potent KRASG12C inhibitors with improved physiochemical properties that are now entering clinical trials. The identification of KRASG12C inhibitors provides a renewed opportunity to develop a more comprehensive understanding of the role of KRAS as a driver oncogene and to explore the clinical utility of direct KRAS inhibition.

KRASG12C mutations are present in lung and colon adenocarcinomas as well as smaller fractions of other cancers. The genetic context of KRASG12C alteration can vary significantly among tumors and is predicted to affect response to KRAS inhibition. KRAS mutations are often enriched in tumors due to amplification of mutant or loss of wild-type allele (8, 9). In addition, KRAS mutations often co-occur with other key genetic alterations including TP53 and CDKN2A in multiple cancers, KEAP1 and/or STK11 in lung adenocarcinoma, or APC and PIK3CA in colon cancer (3, 8–12). Whether differences in KRAS-mutant allele fraction or co-occurrence with other mutations influence response to KRAS blockade is not yet well understood. In addition, due to the critical importance of the RAS pathway in normal cellular function, there is extensive pathway isoform redundancy and a comprehensive regulatory network in normal cells to ensure tight control of temporal pathway signaling. RAS pathway negative feedback signaling is mediated by ERK1/2 and receptor tyrosine kinases (RTK) as well as by ERK pathway target genes including dual-specificity phosphatases (DUSP) and Sprouty (SPRY) proteins (13–17). One important clinically relevant example is provided by the reactivation of ERK signaling observed following treatment of BRAFV600E-mutant cancers with selective BRAF inhibitors (18–20). The observed intertumoral heterogeneity and extensive feedback signaling network in KRAS-mutant cancers may necessitate strategies to more comprehensively block oncogenic signal transduction and deepen the antitumor response in concert with KRAS blockade (15, 21, 22).

Potential strategies to augment the response to KRASG12C inhibitor treatment are evident at multiple nodes of the signaling pathway regulatory machinery. RAS proteins are small GTPases that normally cycle between an active, GTP-bound state and an inactive, GDP-bound state. RAS proteins are loaded with GTP through guanine nucleotide exchange factors (e.g., SOS1) which are activated by upstream RTKs, triggering subsequent interaction with effector proteins that activate RAS-dependent signaling. RAS proteins hydrolyze GTP to GDP through their intrinsic GTPase activity, which is dramatically enhanced by GTPase-activating proteins (GAP). Mutations at codons 12 and 13 in RAS proteins impair GAP-stimulated GTP hydrolysis, leaving RAS predominantly in the GTP-bound, active state.

Potent covalent KRASG12C inhibitors described to date bind only GDP-bound KRAS (5–7). Although codon 12 and 13 mutations decrease the fraction of GDP-bound KRAS, recent biochemical analyses revealed that KRASG12C exhibits the highest intrinsic GTP hydrolysis rate and highest nucleotide exchange rate among KRAS mutants (23). Furthermore, the nucleotide-bound state of KRASG12C can be shifted toward the GDP-bound state by pharmacologically modulating upstream signaling with RTK inhibitors that increase the activity of KRASG12C inhibitors (7, 22, 24). Likewise, SHP2 is a phosphatase that positively transduces RTK signaling to KRAS. Accordingly, SHP2 inhibitors are active in cancers driven by KRAS mutations that are dependent on nucleotide cycling, including KRASG12C (25–27).

MRTX849 is among the first KRASG12C inhibitors to advance to clinical trials. The comprehensive and durable inhibition of KRASG12C by MRTX849 provides a unique opportunity to understand the extent to which KRAS functions as an oncogenic driver. In addition, the observation that the response to blockade of KRAS is markedly different in vitro and in vivo indicates that evaluation of the consequences of KRAS blockade in in vivo model systems is critical to understand the role of KRAS-driven tumor progression. The demonstration of partial responses in patients with lung and colon adenocarcinomas treated with MRTX849 in clinical trials indicates that results observed in tumor models extend to KRASG12C-positive human cancers. Our comprehensive molecular characterization of multiple tumor models at baseline and during response to KRAS inhibition has provided further insight toward the contextual role of KRAS mutation in the setting of genetic and tumoral heterogeneity. Finally, further interrogation of these genetic alterations and signaling pathways utilizing functional genomics strategies including CRISPR and combination approaches uncovered regulatory nodes that sensitize tumors to KRAS inhibition when cotargeted.

Introduction

KRAS is one of the most frequently mutated oncogenes in cancer; however, efforts to directly target KRAS have been largely unsuccessful due to its high affinity for GTP/GDP and the lack of a clear binding pocket (1–4). More recently, compounds were identified that covalently bind to KRASG12C at the cysteine 12 residue, lock the protein in its inactive GDP-bound conformation, inhibit KRAS-dependent signaling, and elicit antitumor responses in tumor models (5–7). Advances on early findings demonstrated that the binding pocket under the switch II region was exploitable for drug discovery, culminating in the identification of more potent KRASG12C inhibitors with improved physiochemical properties that are now entering clinical trials. The identification of KRASG12C inhibitors provides a renewed opportunity to develop a more comprehensive understanding of the role of KRAS as a driver oncogene and to explore the clinical utility of direct KRAS inhibition.

KRASG12C mutations are present in lung and colon adenocarcinomas as well as smaller fractions of other cancers. The genetic context of KRASG12C alteration can vary significantly among tumors and is predicted to affect response to KRAS inhibition. KRAS mutations are often enriched in tumors due to amplification of mutant or loss of wild-type allele (8, 9). In addition, KRAS mutations often co-occur with other key genetic alterations including TP53 and CDKN2A in multiple cancers, KEAP1 and/or STK11 in lung adenocarcinoma, or APC and PIK3CA in colon cancer (3, 8–12). Whether differences in KRAS-mutant allele fraction or co-occurrence with other mutations influence response to KRAS blockade is not yet well understood. In addition, due to the critical importance of the RAS pathway in normal cellular function, there is extensive pathway isoform redundancy and a comprehensive regulatory network in normal cells to ensure tight control of temporal pathway signaling. RAS pathway negative feedback signaling is mediated by ERK1/2 and receptor tyrosine kinases (RTK) as well as by ERK pathway target genes including dual-specificity phosphatases (DUSP) and Sprouty (SPRY) proteins (13–17). One important clinically relevant example is provided by the reactivation of ERK signaling observed following treatment of BRAFV600E-mutant cancers with selective BRAF inhibitors (18–20). The observed intertumoral heterogeneity and extensive feedback signaling network in KRAS-mutant cancers may necessitate strategies to more comprehensively block oncogenic signal transduction and deepen the antitumor response in concert with KRAS blockade (15, 21, 22).

Potential strategies to augment the response to KRASG12C inhibitor treatment are evident at multiple nodes of the signaling pathway regulatory machinery. RAS proteins are small GTPases that normally cycle between an active, GTP-bound state and an inactive, GDP-bound state. RAS proteins are loaded with GTP through guanine nucleotide exchange factors (e.g., SOS1) which are activated by upstream RTKs, triggering subsequent interaction with effector proteins that activate RAS-dependent signaling. RAS proteins hydrolyze GTP to GDP through their intrinsic GTPase activity, which is dramatically enhanced by GTPase-activating proteins (GAP). Mutations at codons 12 and 13 in RAS proteins impair GAP-stimulated GTP hydrolysis, leaving RAS predominantly in the GTP-bound, active state.

Potent covalent KRASG12C inhibitors described to date bind only GDP-bound KRAS (5–7). Although codon 12 and 13 mutations decrease the fraction of GDP-bound KRAS, recent biochemical analyses revealed that KRASG12C exhibits the highest intrinsic GTP hydrolysis rate and highest nucleotide exchange rate among KRAS mutants (23). Furthermore, the nucleotide-bound state of KRASG12C can be shifted toward the GDP-bound state by pharmacologically modulating upstream signaling with RTK inhibitors that increase the activity of KRASG12C inhibitors (7, 22, 24). Likewise, SHP2 is a phosphatase that positively transduces RTK signaling to KRAS. Accordingly, SHP2 inhibitors are active in cancers driven by KRAS mutations that are dependent on nucleotide cycling, including KRASG12C (25–27).

MRTX849 is among the first KRASG12C inhibitors to advance to clinical trials. The comprehensive and durable inhibition of KRASG12C by MRTX849 provides a unique opportunity to understand the extent to which KRAS functions as an oncogenic driver. In addition, the observation that the response to blockade of KRAS is markedly different in vitro and in vivo indicates that evaluation of the consequences of KRAS blockade in in vivo model systems is critical to understand the role of KRAS-driven tumor progression. The demonstration of partial responses in patients with lung and colon adenocarcinomas treated with MRTX849 in clinical trials indicates that results observed in tumor models extend to KRASG12C-positive human cancers. Our comprehensive molecular characterization of multiple tumor models at baseline and during response to KRAS inhibition has provided further insight toward the contextual role of KRAS mutation in the setting of genetic and tumoral heterogeneity. Finally, further interrogation of these genetic alterations and signaling pathways utilizing functional genomics strategies including CRISPR and combination approaches uncovered regulatory nodes that sensitize tumors to KRAS inhibition when cotargeted.

Results

MRTX849 Is a Potent and Selective Inhibitor of KRASG12C, KRAS-Dependent Signal Transduction, and Cell Viability In Vitro

A structure-based drug design approach, including optimization for favorable drug-like properties, led to the discovery of MRTX849 as a potent, covalent KRASG12C inhibitor (Fig. 1A; Supplementary Table S1). An LC/MS-based KRASG12C protein modification assay revealed that MRTX849 demonstrated much greater modification of KRASG12C when preloaded with GDP compared with GTP (Supplementary Table S2), supporting that MRTX849 binds to and stabilizes the inactive GDP-bound form of KRASG12C. Indeed, introducing a comutation that impairs the GTPase activity of KRASG12C (24) attenuated the inhibitory activity of MRTX1257, a close analogue of MRTX849 (Supplementary Fig. S1A). Secondary mutations that modulate the nucleotide exchange function of KRASG12C also affected inhibition by MRTX1257, supporting that the MRTX compound series is dependent on KRASG12C nucleotide cycling.

Figure 1.

 

MRTX849 is a potent, covalent KRASG12C inhibitor in vitroA, Structure of MRTX849. B, Immunoblot protein Western blot analyses of KRAS pathway targets in MIA PaCa-2 cells treated from 1 hours to 72 hours with MRTX849 at 100 nmol/L. C, Immunoblot protein Western blot analyses of KRAS pathway targets in MIA PaCa-2 cells treated for 24 hours with MRTX849 over a 13-point dose response. D, Left y-axis shows active RAS ELISA assay to determine the reduction in RAS-GTP abundance following MRTX849 treatment in MIA PaCa-2 cells for 24 hours. The vehicle value was normalized to 1 by dividing all average values by the vehicle value. Right y-axis shows quantitation of KRAS band shift by MRTX849 treatment in MIA PaCa-2 cells for 24 hours as assessed by Western blot and densitometry. E, In-cell Western blot assay to evaluate modulation of pERK in MIA PaCa-2 cells grown in standard tissue-culture conditions treated with MRTX849 over a time course. F, CellTiter-Glo assay to evaluate cell viability performed on seven KRASG12C-mutant cell lines and three non–KRASG12C-mutant cell lines grown in 2-D tissue-culture conditions in a 3-day assay (left plot) or 3-D conditions using 96-well, ULA plates in a 12-day assay (right plot).

 

We next determined the cellular activity of MRTX849 utilizing the KRASG12C-mutant H358 lung and MIA PaCa-2 pancreatic cancer cell lines. In both models, MRTX849 demonstrated an upward electrophoretic mobility shift of KRASG12C protein band migration by immunoblot, indicative of covalent modification of KRASG12C. A maximal mobility shift was observed by 1 hour, was maintained through 72 hours (Fig 1B; Supplementary Fig. S1B), and was evident at concentrations as low as 2 nmol/L with near-maximal modification observed at 15.6 nmol/L (Fig. 1C; Supplementary Fig. S1C). Comparable inhibition of active RAS was observed as determined by a RAF RAS-binding domain capture ELISA assay (Fig. 1D; 1D). MRTX849 also inhibited KRAS-dependent signaling targets including ERK1/2 phosphorylation (pERK; Thr202/Tyr204 ERK1), S6 phosphorylation (pS6; RSK-dependent Ser235/236), and expression of the ERK-regulated DUSP6, each with IC50 values in the single-digit nanomolar range in both cell lines (Fig. 1B and C; Supplementary Fig. S1B and S1C). The evaluation of pERK over a time course of 48 hours indicated maximal inhibition was observed at 24 hours (Fig. 1E; Supplementary Fig. S1E). Treatment with the des-acrylamide version of MRTX849, which is unable to covalently bind to KRASG12C, did not demonstrate significant inhibition of ERK phosphorylation (Supplementary Fig. S1F). The H358 cell line was selected for determination of MRTX849 cysteine selectivity utilizing an LC/MS-based proteomics approach able to detect approximately 6,000 cysteine-containing peptides. After treatment for 3 hours, decreased KRASG12C Cys12-free peptide was detected with treated-to-control ratios of 0.029 and 0.008 determined at 1 and 10 μmol/L, respectively, indicating near-complete engagement of the intended target (Supplementary Table S3). In contrast, the only other peptides identified were from lysine-tRNA ligase (KARS) at Cys209 near the detection limit, indicating a high degree of selectivity toward KRASCYS12.

To evaluate the breadth of MRTX849 activity, its effect on cell viability was determined across a panel of 17 KRASG12C-mutant and 3 non–KRASG12C-mutant cancer cell lines using 2-D (3-day, adherent cells) and 3-D (12-day, spheroids) cell-growth conditions. MRTX849 potently inhibited cell growth in the vast majority of KRASG12C-mutant cell lines with IC50 values ranging between 10 and 973 nmol/L in the 2-D format and between 0.2 and 1,042 nmol/L in the 3-D format (Supplementary Table S4; Fig. 1F). In agreement with prior KRASG12C inhibitor studies (5), MRTX849 demonstrated improved potency in the 3-D assay format, as all but one KRASG12C-mutant cell line exhibited an IC50 value below 100 nmol/L. Although MRTX849 was broadly effective in inhibiting viability of KRASG12C-mutant cell lines, IC50 values varied across the cell panel by 100-fold, suggesting a differential degree of sensitivity to treatment. All three non–KRASG12C-mutant cell lines tested demonstrated IC50 values greater than 1 μmol/L in 2-D conditions and greater than 3 μmol/L in 3-D conditions, suggesting the effect of MRTX849 on cell viability was dependent on the presence of KRASG12C.

To determine whether the difference in sensitivity across the cell panel correlated with the ability of MRTX849 to bind to KRAS or inhibit KRAS-dependent signal transduction, seven KRASG12C-mutant cancer cell lines were selected from the panel for further evaluation. In each cell line, MRTX849 demonstrated a very similar concentration-dependent electrophoretic mobility shift (IC50) for KRASG12C protein migration, suggesting that the ability to bind to and modify KRASG12C does not readily account for differences in response in viability studies (Fig. 1B and C; Supplementary Figs. S1B and S1C and S2A and S2B). The effect of MRTX849 on selected phosphoproteins implicated in mediating KRAS-dependent signaling was also evaluated across the cell panel by immunoblot and/or reverse-phase protein array (RPPA) following treatment for 6 or 24 hours. Notably, the concentration–response relationship and maximal effect of MRTX849 on inhibition of ERK and S6S235/236 phosphorylation varied across the cell panel (Supplementary Fig. S2A and S2C; Supplementary Table S7). MRTX849 demonstrated only partial inhibition of pERK in KYSE-410 and SW1573 cells and a minimal effect on pS6S235/236 in SW1573, H2030, and KYSE-410 cells (Supplementary Fig. S2A and S2C). Each of these cell lines were among those that exhibited a submaximal response to MRTX849 in both 2-D and 3-D viability assays (Fig. 1F). Although KRAS is implicated in mediating signal transduction through the PI3K and mTOR pathways, there was minimal evidence of a significant and/or durable effect of MRTX849 on AKT (S473, T308) or 4E-BP1 (T37/T46, S65, T70) phosphorylation at any time point in any cell lines evaluated (Supplementary Fig. S2D). However, MRTX849 demonstrated concentration-dependent partial inhibition of the mTOR-dependent signaling targets p70 S6 kinase (T412) and/or pS6 (S240/44) in the H358, MIA PaCa-2,H2122, and H1373 cell lines, each of which exhibited a maximal response to treatment. Together, these data suggest that maximizing inhibition of KRAS-dependent ERK and S6 signaling may be required to elicit a robust response in tumor-cell viability assays.

MRTX849 Treatment In Vivo Leads to Dose-Dependent KRASG12C Modification, KRAS Pathway Inhibition, and Antitumor Efficacy

Studies were conducted to evaluate MRTX849 antitumor activity along with its pharmacokinetic and pharmacodynamic properties in vivo, both to understand the clinical utility of this agent and to provide insight toward response to treatment. MRTX849 demonstrated moderate plasma clearance and prolonged half-life following oral administration (Supplementary Table S1; Supplementary Fig. S3). To evaluate the pharmacodynamic response to MRTX849 and to correlate drug exposure with target inhibition, MRTX849 was administered via oral gavage over a range of dose levels to H358 xenograft–bearing mice, and plasma and tumors were collected at defined time points. The fraction of covalently modified KRASG12C protein was proportional to the plasma concentration of MRTX849 (Fig. 2A). When evaluated over time after a single oral dose at 30 mg/kg, the modified fraction of KRASG12C was 74% at 6 hours after dose and gradually decreased to 47% by 72 hours (Fig. 2B). This extended pharmacodynamic effect, despite declining levels of MRTX849 in plasma, was consistent with the irreversible inhibition of KRASG12C by MRTX849 and the relatively long half-life for the KRASG12C protein (∼24–48 hours; Supplementary Table S5). The modification of KRASG12C was maximized after repeated daily dosing for 3 days at 30 mg/kg (Fig. 2B), and higher dose levels did not demonstrate additional KRASG12C modification in multiple tumor models (data not shown). The maximum level of modification of approximately 80%, despite increasing dose and plasma levels of MRTX849, suggests that accurate measurement of complete inhibition of KRASG12C utilizing LC/MS may not be attainable, potentially due to a pool of active, noncycling, or unfolded KRASG12C protein in tumors. Together, these studies demonstrated a dose-dependent increase in covalent modification of KRASG12C by MRTX849 and that the majority of targetable KRAS was covalently modified by MRTX849 over a repeated administration schedule at dose levels at or exceeding 30 mg/kg.

Figure 2.

 

MRTX849 modifies KRASG12C and inhibits KRAS signaling and tumor growth in vivoA, MRTX849 was administered orally as a single dose to mice bearing established H358 xenografts (average tumor volume ∼350 mm3) at 10, 30, and 100 mg/kg. KRAS modification and MRTX849 plasma concentration data from n = 3 mice are shown as mean ± SD. KRASG12C modification was statistically significant versus vehicle control using the two-tailed Student t test. **, P < 0.01. B, MRTX849 was administered orally as a single dose or daily (QD) for 3 days to mice bearing established H358 xenografts (average tumor volume ∼350 mm3) at 30 mg/kg. Plasma was collected at 0.5, 2, 6, 24, 48, and 72 hours after administration of the last dose, and tumors were collected at 6, 24, 48, and 72 hours after dose. KRASG12C modification and MRTX849 plasma concentration data are shown from n = 3 mice as mean ± SD. Induction of modified KRASG12C protein at all time points was determined to be statistically significant versus vehicle control using two-way ANOVA. In addition, induction of modified KRASG12C protein at 72 hours in day 1 samples and 48 and 72 hours in day 3 samples was statistically significant versus the 6-hour time point. Brackets indicate P < 0.05 as compared with left-most sample. C, MRTX849 was administered as in A. Tumors were collected 6 hours after dose, and total and phosphorylated ERK1/2 and total and phosphorylated S6 were analyzed by immunoblot and quantified by densitometric analysis. Relative fluorescence intensity of pERK1/2 and pS6 was normalized by dividing pERK1/2 and pS6 by total ERK1/2 and total S6, respectively. Vehicle-treated tumors were normalized to 1 by dividing all average values by the vehicle value. Average pERK1/2 and pS6 values were divided by the average value in vehicle-treated tumors. Data shown represent the average of 2 to 3 tumors per treatment group plus SD. Reduction of pS6 relative fluorescence intensity was determined to be statistically significant versus vehicle control using the two-tailed Student t test. Brackets indicate P < 0.05 compared with left-most sample. D, MRTX849 was administered as in B. Tumors were collected at 6, 24, 48, or 72 hours after administration of the last dose, and total and phosphorylated ERK1/2 and total and phosphorylated S6 were analyzed as in C. Data shown represent the average of 3 to 4 tumors per treatment group plus SD. Reduction of pS6 relative fluorescence intensity on day 3 was determined to be statistically significant versus vehicle control using two-way ANOVA. Brackets indicate P < 0.05 compared with left-most sample. E, MRTX849 was administered via daily oral gavage at the doses indicated to mice bearing established MIA PaCa-2 xenografts. Dosing was initiated when tumors were approximately 350 to 400 mm3.MRTX849 was administered to mice daily until day 16. Data are shown as mean tumor volume ± SEM. Tumor volumes at day 16 were determined to be statistically significant versus vehicle control via two-tailed Student t test. **, P < 0.01; *, P < 0.05.

 

To evaluate the effect of MRTX849 on KRAS-dependent signal transduction in vivo, a single dose of MRTX849 at 10, 30, or 100 mg/kg was administered to H358 tumor–bearing mice. Dose-dependent inhibition of ERK1/2 and pS6S235/36 phosphorylation was observed at 6 hours after dose based on immunoblot and densitometric analysis (Fig. 2C). MRTX849 also demonstrated marked inhibition of ERK1/2 and S6S235/36 phosphorylation after one or three daily doses at 6 or 24 hours, and levels gradually recovered by 72 hours after the final dose (Fig. 2D). pERK1/2 and pS6S235/36 were further evaluated in formalin-fixed, paraffin-embedded sections from vehicle-treated and MRTX849-treated xenografts in four tumor models utilizing IHC methods coupled with image analysis algorithms. These studies demonstrated increased pERK1/2 and pS6 in nontumor/stromal cells following MRTX849 administration, indicating that immunoblotting studies with bulk tumor lysate likely underrepresent the degree of pathway inhibition in tumor cells, whereas IHC-based evaluation may more accurately reflect both the degree and spatial impact of pathway inhibition. Maximal inhibition was observed for both ERK and S6S235/36 phosphorylation after a single dose at the 6-hour time point, with a rebound in signaling evident 24 hours after single dose in each model (Supplementary Fig. S4). Marked inhibition of ERK phosphorylation was observed at 6 hours after administration, with 89%, 94%, and 94% inhibition observed compared with vehicle controls in MIA PaCa-2, H1373, and H2122 tumors, respectively (H358 pERK not quantifiable). This indicates that dose levels at or exceeding 30 mg/kg dose maximized inhibition of ERK phosphorylation in multiple models (Supplementary Fig. S4A and S4B). Inhibition of S6 phosphorylation at 6 hours was more variable, with percent inhibition values of 76%, 50%, 86%, and 56% observed in MIA PaCa-2, H1373, H358, and H2122 tumors, respectively (Supplementary Fig. S4B). Together, these data indicate that consistent acute (6 hours) inhibition of KRAS-dependent ERK phosphorylation was maximized in all evaluated models, whereas inhibition of S6S235/36 was more variable, presumably due to varying degrees of KRAS-independent activation of this pathway in different tumor models.

MIA PaCa-2 and H358 were selected as MRTX849-responsive tumor models, thereby enabling a high-resolution understanding of dose–response relationships. Significant, dose-dependent, antitumor activity was observed at the 3, 10, 30, and 100 mg/kg dose levels in the MIA PaCa-2 model (Fig. 2E). Evidence of rapid tumor regression was observed at the earliest post-treatment tumor measurement, and animals in the 30 and 100 mg/kg cohorts exhibited evidence of a complete response at study day 15. Dosing was stopped at study day 16, and all 4 mice in the 100 mg/kg cohort and 2 of 7 mice in the 30 mg/kg cohort remained tumor-free through study day 70 (Supplementary Fig. S5A). In a second MIA PaCa-2 study, dose-dependent antitumor efficacy was observed at the 5, 10, and 20 mg/kg dose levels, and 2 of 5 mice at the 20 mg/kg dose level exhibited complete tumor regression (Supplementary Fig. S5B). Significant dose-dependent antitumor efficacy was also observed in the H358 model, including 61% and 79% tumor regression at the 30 and 100 mg/kg dose levels, respectively, at day 22 (Supplementary Fig. S5C). MRTX849 was well tolerated, and no effect on body weight was observed at all dose levels evaluated (Supplementary Fig. S5D). These studies indicated that MRTX849 demonstrated dose-dependent antitumor efficacy over a well-tolerated dose range and that the maximally efficacious dose of MRTX849 is between 30 and 100 mg/kg/day.

MRTX849 Demonstrates Broad-Spectrum Tumor Regression in KRASG12C Cell Line and Patient-Derived Xenograft Models

To evaluate the breadth of antitumor activity across genetically and histologically heterogeneous KRASG12C-mutant cancer models, MRTX849 was evaluated at a fixed dose of 100 mg/kg/day (a dose projected to demonstrate near-maximal target inhibition in most models) in a panel of human KRASG12C-mutant cell line–derived xenograft (CDX) and patient-derived xenograft (PDX) models. MRTX849 demonstrated tumor regression exceeding 30% volume reduction from baseline in 17 of 26 models (65%) at approximately 3 weeks of treatment (Fig. 3A; Supplementary Table S6). By comparison, MRTX849 did not exhibit significant antitumor activity at 100 mg/kg in three non–KRASG12C-mutant models (Fig. 3A; Supplementary Table S6). Together, these results indicate that KRASG12C-mutant tumors are broadly dependent upon mutant KRAS for tumor-cell growth and survival and that MRTX849 elicits antitumor activity through a KRASG12C-dependent mechanism.

Figure 3.

 

Antitumor activity of MRTX849 in KRASG12C-mutant and non–KRASG12C-mutant human tumor xenograft models. A, MRTX849 was administered via oral gavage at 100 mg/kg every day to mice bearing the CDX or PDX model indicated. Dosing was initiated when tumors were, on average, approximately 250 to 400 mm3. MRTX849 was formulated as a free base and resuspended as a solution in 10% Captisol and 50 mmol/L citrate buffer, pH 5.0. The percent change from baseline control was calculated at days 19 to 22 for most models. Statistical significance was determined for each model and is shown in Supplementary Table S6. Status of mutations and alterations in key genes is shown below each model. MAF (%), percent KRASG12C-mutant allele fraction by RNA-seq; CNV, copy-number variation; * denotes very high CDK4 expression by RNA-seq and possible amplification. HER family status was determined by averaging EGFRERBB2, and ERBB3 RNA-seq expression for CDX (CCLE) or PDX (Crown huBase) models. Positive HER family calls denote greater than the median expression of the models tested. CDX and PDX model HER family calls were determined independently. B, Tumor-growth inhibition plots from representative xenograft models that were categorized as sensitive, partially sensitive, or treatment refractory.

 

Although MRTX849 exhibited marked antitumor responses in the majority of models tested, a response pattern ranging from delayed tumor growth to complete regression was observed across the xenograft panel. The response to treatment was categorized as sensitive, partially sensitive, or treatment refractory (Fig. 3B). Rank order and Pearson statistical analyses were performed to evaluate the correlation between in vitro potency (IC50 in 2-D or 3-D viability assays) and antitumor response in vivo (% regression or progression on day 22), and a significant correlation between response in cell lines compared with tumor models was not observed (Supplementary Fig. S6A and S6B). Thus, we focused on a comprehensive analysis of correlates with MRTX849 tumor response in vivo, including tumor histology, co-occurring genetic alterations, as well as baseline or drug-induced changes in expression of KRAS-related genes [RNA sequencing (RNA-seq)] and/or protein signaling networks (RPPA in 18 models, ref. 28; Supplementary Fig. S7). No individual genetic alteration, including but not limited to KRAS-mutant allele frequency, TP53, STK11, or CDKN2A, predicted the antitumor activity of MRTX849. Interestingly, baseline gene and/or protein expression of selected members of the HER family of RTKs and of regulators of early cell-cycle transition did exhibit a trend with the degree of antitumor response, suggesting these pathways may influence the response to KRAS inhibitors (Supplementary Fig. S7A). Together, these data indicate that there are no individual binary biomarkers that clearly predict therapeutic response and that the molecular complexity and heterogeneity present in distinct KRAS-mutated tumors may contribute to the response to target blockade.

MRTX849 Antitumor Activity Translates to RECIST Responses in Patients with Cancer

A 45-year-old female former smoker diagnosed with stage IV lung adenocarcinoma and refractory to multiple lines of therapy including carboplatin/pemetrexed/pembrolizumab, docetaxel, and investigational treatment with binimetinib and palbociclib was enrolled onto the MRTX849-001 phase Ib clinical trial with two bilateral lung lesions and mediastinal lymph node as target lesions. Targeted next-generation sequencing (NGS) demonstrated a KRASG12C mutation (c.34G>T). In addition, loss-of-function KEAP1 (K97M) and STK11 (E223*) mutations were detected and are predicted to be deleterious to their respective proteins. The patient was administered MRTX849 (600 mg twice a day) and had marked clinical improvement within 2 weeks, including complete resolution of baseline cough and oxygen dependency. A RECIST-defined partial response of 33% reduction of target lesions was observed at cycle 3 day 1 (45 days), and the patient continues on study (Fig. 4A).

Figure 4.

 

Activity of MRTX849 in patients with lung and colon cancers. A, Pretreatment and 6-week scans of a heavily pretreated patient with a KRASG12C mutation–positive lung adenocarcinoma indicating 33% reduction of target lesions. Patient continues on study. The top plots show a coronal view, and bottom plots an axial view of CT chest images prior to MRTX849 treatment (left) and after two cycles of MRTX849 treatment (right). B, Baseline, 6-week (Cycle 2), and 12-week (Cycle 4) scans of a patient with a KRASG12C mutation–positive colon adenocarcinoma. Partial response (PR) was confirmed at Cycle 4, and patient continues on study. Four lesions (TL1–4) are shown with axial views of CT images prior to MRTX849 treatment (top), after two cycles of MRTX849 treatment (center), and after four cycles of MRTX849 treatment (bottom).

 

A 47-year-old female never-smoker with metastatic adenocarcinoma of the left colon who exhibited progressive disease after receiving multiple lines of systemic therapy, including FOLFOX plus bevacizumab, single-agent capecitabine, FOLFIRI plus bevacizumab, and an investigational antibody–drug conjugate, was enrolled into the MRTX849-001 phase Ib clinical trial. This patient had extensive metastases involving the liver, peritoneum, ovaries, and lymph nodes. Targeted NGS identified a KRASG12C mutation. The patient was administered MRTX849 (600 mg twice a day) and demonstrated marked clinical improvement within 3 weeks and a visible decrease in size of her umbilical Sister Mary Joseph nodule. Her carcinoembryonic antigen levels decreased from 77 ng/mLat baseline to 11 ng/mL at cycle 2 day 1 and 3 ng/mL by cycle 3 day 1 (normal range, 0–5 ng/mL). A RECIST-defined partial response with a 37% reduction of target lesions and complete response of a nontarget lesion was observed at cycle 3 day 1(day 42). Confirmatory CT scans were conducted at cycle 5, day 1 (day 84) and indicated a confirmed RECIST partial response with further reduction of target lesions at −47% from baseline (Fig. 4B). The patient remains on treatment through Cycle 6.

Temporal Effects of MRTX849 on KRAS-Dependent Signaling and Feedback Pathways and Relationship to Antitumor Activity Following Repeat Dosing in Xenograft Models

A comprehensive analysis was conducted to evaluate MRTX849-induced temporal molecular changes to further interrogate mechanisms of drug response across sensitive and partially sensitive models. To evaluate temporal changes in global gene expression, xenograft-bearing mice were administered vehicle or 100 mg/kg MRTX849, and RNA-seq was performed on tumors at 6 and 24 hours after treatment. Gene expression was evaluated at day 1 and day 5 for the sensitive models MIA PaCa-2 and H1373 to ensure sufficient tissue availability from regressing tumors, or at day 7 in the partially sensitive models H358, H2122, and H2030 to coordinate with tumor stasis plateau. The top differentially expressed gene set enrichment analysis (GSEA) hallmark gene sets, regardless of tumor response, in all five models were several KRAS-annotated gene sets confirming MRTX849 selectively inhibits multiple genes directly related to KRAS signaling. MYC, mTOR, cell cycle, and apoptosis/BCL2 pathway gene sets were also strongly differentially expressed, confirming MRTX849 broadly affected multiple well-established, KRAS-regulated pathways, several of which have proved difficult to directly inhibit with previous targeted therapies (Fig. 5A and B; Supplementary Fig. S8A—S8D). The marked impact of MRTX849 on a large number of genes that regulate cell cycle and apoptosis provides further insight into molecular mechanisms which mediate its antitumor activity.

Figure 5.

 

MRTX849 treatment in vivo regulates KRAS-dependent oncogenic signaling and feedback-inhibitory pathways. A, Volcano plots displaying differentially expressed genes in xenograft tumors 24 hours after oral administration of vehicle or 100 mg/kg MRTX849 in a representative MRTX849-sensitive (H1373) and MRTX849-partially sensitive (H358) model. Significance denoted in the legend (Padj < 0.01). B, GSEA heat maps depicting hallmark signature pathways differentially regulated in at least one model 24 hours following oral administration of a single 100 mg/kg MRTX849 dose compared with vehicle. Normalized enrichment score shown in all models 6 or 24 hours after a single dose (QD × 1) or 5 (QD × 5) or 7 (QD × 7) days dosing. C, Genes that feedback-inhibit MAP kinase signaling are downregulated following MRTX849 treatment in all five cell line xenografts assessed by RNA-seq. TPM, Transcripts Per Kilobase Million.

 

Targeted RNA-seq analysis was performed on genes implicated in the temporal regulation of external signaling inputs and feedback pathways which collectively temper signaling flux through the RAS–RAF–MEK–ERK MAP kinase (MAPK) pathway including DUSPSPRY, and PHLDA family genes (13, 18). These MAPK pathway–negative regulators were each ranked among the most strongly decreased genes following MRTX849 treatment, providing evidence that ERK-dependent transcriptional output is blocked and that pathways involved in reactivation of RTK- and ERK-dependent signaling were activated (Fig. 5C; Supplementary Fig. S4A).

On the basis of the observation of dynamic changes in transcriptional programs linked to KRAS pathway reactivation, IHC plus quantitative imaging of tumor cell–specific pERK and pS6 was evaluated over a range of time points. In the sensitive MIA PaCa-2 and H1373 tumor models, treatment with MRTX849 (100 mg/kg) demonstrated ≥90% inhibition of ERK phosphorylation at 6 and 24 hours on both days 1 and 5 (Supplementary Fig. S4). In contrast, in the partially sensitive H358 and H2122 models, robust inhibition of ERK phosphorylation was observed at 6 hours after a single dose; however, marked recovery of ERK phosphorylation was observed at 24 hours after single dose and at both 6 and 24 hours following 7 days of repeat-dose administration. Because DUSPSPRY, and ETV family transcripts remain downregulated through 5 to 7 days in all models, it is evident that other independent factors contribute to temporal reactivation of ERK (Fig. 5C). Similar to what was observed with single-dose administration, the effect of MRTX849 on pS6 was variable over time and did not track with the antitumor activity of MRTX849. Together, these results suggest that the extent and duration of inhibition of pERK may track with the magnitude of antitumor efficacy of KRASG12C inhibitors and that further evaluation of the role of S6 is required to understand if it plays a role in drug sensitivity.

The effect of MRTX849 on cell proliferation and apoptosis was characterized by IHC analysis of Ki-67 or cleaved caspase-3 after a single dose or repeat administration. The fraction of Ki-67–positive cells was significantly reduced in tumors after repeat administration in all four models tested, further supporting a broadly operative antiproliferative mechanism, independent of the magnitude of MRTX849 antitumor response (Supplementary Fig. S4). Induction of apoptosis as determined by cleaved caspase-3 immunostaining was also evident on day 1 of treatment (6 and/or 24 hours after treatment) in the sensitive H358, MIA PaCa-2, and H1373 models (79%–100% maximal regression) but not in the partially sensitive H2122 model (Supplementary Fig. S4). An expanded RPPA-based pathway analysis of several models also indicated a correlation between antitumor activity of MRTX849 and decreased survivin (statistically significant at days 5/7 in 7 models evaluated; Supplementary Fig. S7B) and a trend toward increased cleaved caspase-3 induction (day 1, P = 0.08, 16 models), supporting the induction of apoptosis as a key mediator of a cytoreductive antitumor response (Supplementary Fig. S7C). Interestingly, the magnitude of reduction of MYC and cyclin B1 protein levels at days 5/7 also closely correlated with MRTX849 antitumor activity, consistent with their roles as critical regulators of KRAS-mediated cell growth and survival pathways (Supplementary Fig. S7B). Collectively, these data support that durable inhibition of ERK activity and maximal inhibition of ERK-regulated outputs including MYC and E2F-mediated transcription are associated with induction of apoptosis and maximal response to MRTX849 treatment.

CRISPR/Cas9 Screen Identifies Vulnerabilities and Modifiers of Response to MRTX849 in KRASG12C-Mutant Cancer Cell Lines In Vitro and In Vivo

The correlative analysis of genomic or proteomic markers with response to MRTX849 in the defined panel of models provided only limited insight toward mechanism of therapeutic response or resistance. Therefore, we directly interrogated the role of selected genes in mediating therapeutic response utilizing a focused CRISPR/Cas9 knockout screen targeting approximately 400 genes including many genes involved in KRAS signaling. This was conducted in H358 and H2122 cells in vitro and in H2122 xenografts in vivo in the presence and absence of MRTX849 treatment (Supplementary Fig. S9A–S9F). In MRTX849-anchored screens in vitro, single guide RNAs (sgRNA) that target RAS signaling pathway genes including MYC, SHP2 (H2122), mTOR pathway (MTOR and RPS6), and cell-cycle genes (CDK1, CDK2, CDK4/6, and RB1) were identified to affect cell fitness. sgRNAs that target KEAP1 and CBL were enriched in the H2122 model, demonstrating cell-specific genetic routes toward improved fitness through loss of classic tumor-suppressor genes, including in the context of MRTX849 treatment. KRAS sgRNA dropout was less pronounced in the MRTX849-treated cells compared with DMSO control–treated cells, as would be expected with redundant depletion of the drug target (Supplementary Fig. S9C and S9D). To evaluate whether a distinct KRAS dependence or modulation of MRTX849 therapeutic response was observed in vitro versus in vivo, xenograft-bearing mice bearing H2122 cells (∼250 mm3) transduced with the sgRNA library were orally administered vehicle or MRTX849 for 2 weeks (Supplementary Fig. S9A, S9E, and S9F). In MRTX849-treated xenografts, sgRNAs targeting cell cycle, SHP2, MYC, and mTOR pathway genes remained among the top depleted sgRNAs, demonstrating that inhibition of these targets in vivo, in the context of KRAS inhibition, leads to further tumor-growth inhibition over and above the effects of KRAS inhibition alone (Supplementary Fig. S9E and S9F). sgRNAs targeting the tumor suppressor KEAP1 were enriched in MRTX849-treated xenografts, suggesting loss of KEAP1 may represent a mechanism of intrinsic or acquired resistance. Interestingly, NRAS was one of the top enriched genes in the vehicle-treated xenografts, suggesting NRAS functions as a tumor suppressor in this context; however, enrichment was not as pronounced in the MRTX849-treated xenografts, suggesting NRAS may compensate for KRAS in the context of KRAS inhibition (Supplementary Fig. S9F). Collectively, these data demonstrate the importance of selected proteins that regulate RTK- and RAS-dependent signaling and cell-cycle transition in mediating the oncogenic effects of mutant KRAS, and also provide a catalog of potentially druggable vulnerabilities that complement KRAS blockade.

Cancer Therapeutic Combination Screen to Identify Rational and Clinically Tractable Strategies to Address Feedback and Resistance Pathways

To further interrogate pathways that mediate the antitumor response to MRTX849 and to identify combinations capable of enhancing response to MRTX849, a combination screen was conducted in vitro using a focused library of small-molecule inhibitors across a panel of cell lines (Supplementary Fig. S10A and S10B; Supplementary Table S8). Approximately 70 compounds targeting relevant pathways (RTKs, MAPK/ERK, PI3K, mTOR, cell cycle) were tested in a 3- or 7-day viability assay, and synergistic combinations were identified and ranked. Multiple hits from this screen were then identified for additional evaluation in combination studies with MRTX849, including the HER family inhibitor afatinib, the CDK4/6 inhibitor palbociclib, the SHP2 inhibitor RMC-4550, and mTOR pathway inhibitors.

Combination Strategies That Target Upstream Signaling Pathways Implicated in Extrinsic Regulation of KRAS Nucleotide Cycling and Feedback/Bypass Pathways

MRTX849 in combination with HER family inhibitors synergistically inhibited tumor-cell viability in the majority of cell lines evaluated and were the top hit in the combination screen in vitro (Supplementary Fig. S10). Cell lines with the highest (top 50th percentile) average composite baseline RNA expression values of selected HER family members exhibited the highest synergy scores to these combinations (Supplementary Fig. S11A). Afatinib was selected as a prototype HER family inhibitor based on its broad in vitro combination activity. Combination studies were conducted with MRTX849 and afatinib in five tumor models that were partially sensitive or treatment refractory to single-agent MRTX849. The MRTX849 and afatinib combination demonstrated significantly greater antitumor efficacy compared with either single agent in all five models evaluated, including multiple models exhibiting complete or near-complete responses to the combination (Fig. 6A; Supplementary Fig. S11B).

Figure 6.

 

HER family and SHP2 inhibitor combinations further inhibit KRAS signaling and exhibit increased antitumor responses. A, MRTX849 at 100 mg/kg, afatinib at 12.5 mg/kg, or the combination was administered daily via oral gavage to mice bearing the H2122 or KYSE-410 cell line xenografts (n = 5). Combination treatment led to a statistically significant decrease in tumor growth compared with either single-agent treatment. *, Padj < 0.01. B, Quantification of KRAS mobility shift and pERK in H2122 cells treated for 24 hours with MRTX849 (0.1–73 nmol/L), afatinib (200 nmol/L), or the combination assessed by Western blot densitometry. C, MRTX849 at 100 mg/kg, afatinib at 12.5 mg/kg, or the combination was administered once or daily for 7 days via oral gavage to mice bearing H2122 cell line xenografts (n = 3/group). Tumors were harvested at 6 and 24 hours following the final dose. Tumor sections were stained for pERK and pS6 via IHC methods. Quantitation of images shown by H-score in tumor tissue. Reduction of pERK or pS6 staining intensity was determined to be statistically significant relative to vehicle or either single agent using one-way ANOVA. Brackets indicate P < 0.05 compared with left-most sample. D, Quantitation of KRAS band shift and pERK after 24-hour treatment with MRTX849 (0.1–73 nmol/L), RMC-4550 (1 μmol/L), or the combination in H358 cells assessed by Western blot densitometry. E, MRTX849 at 100 mg/kg, RMC-4550 at 30 mg/kg, or the combination was administered daily via oral gavage to mice bearing the KYSE-410 or H358 cell line xenografts (n = 5/group). Combination treatment led to a statistically significant reduction in tumor growth compared with either single agent on the last day of dosing. *, Padj < 0.05. F, MRTX849 at 100 mg/kg, RMC-4550 at 30 mg/kg, or the combination was administered via oral gavage to mice bearing KYSE-410 cell line xenografts (n = 3/group), and tumors were harvested at 6 and 24 hours post-dose. Tumor sections were stained with pERK or pS6 via IHC methods. Quantitation of images shown by H-score in tumor tissue. Reduction of pERK staining intensity was determined to be statistically significant relative to RMC-4550 alone using one-way ANOVA. Brackets indicate P < 0.05 compared with left-most sample.

 

To evaluate whether afatinib affected covalent modification of KRASG12C by MRTX849, partially sensitive H2122 cells were treated with increasing concentrations of MRTX849 alone or in the presence of afatinib (200 nmol/L, IC90), and the mobility shift in KRAS protein was densitometrically determined from immunoblots. A clear shift in the concentration response to MRTX849 was apparent in the presence of afatinib, indicating that the combination increased the fraction of modified KRASG12C consistent with the putative role of HER family receptors in extrinsic regulation of KRASG12C GTP loading (Fig. 6B). The concentration–response relationship for inhibition of ERK phosphorylation was also clearly shifted in the presence of afatinib. To further evaluate the effect of the combination on KRAS-dependent signaling, four cell lines (H2030, H2122, H358, and KYSE-410) were treated with a range of MRTX849 concentrations in the presence or absence of afatinib for 6 or 24 hours, and key signaling molecules were evaluated by RPPA. Afatinib demonstrated clear inhibition of EGFR (pY1068) and HER2 (pY1248) activity and partial inhibition of ERK, AKT (S473), and p70S6K phosphorylation at both time points (Supplementary Fig. S11C). The effect of afatinib on S6 (S235/236, S240/244) and p90 RSK (S380) phosphorylation was variable and exhibited only minimal inhibition in most of the cell lines evaluated. The combination of afatinib and MRTX849 demonstrated markedly enhanced concentration-dependent inhibition and/or a greater magnitude of effect on ERK, RSK, p70 S6K, and S6 (S235/236) phosphorylation compared with MRTX849 alone at both 6 and 24 hours. Of note, neither afatinib nor MRTX849 alone inhibited S6 phosphorylation at the S240/244 site regulated by mTOR/S6K, whereas the combination demonstrated marked inhibition at 24 hours.

In vivo, the combination also exhibited a trend toward increased pERK and pS6 (S235/236) inhibition in the partially sensitive H2122 model in combination groups as determined by quantitation of immunostaining after 1- or 7-day administration (Fig. 6C). Similar results were observed in the MRTX849-refractory KYSE-410 model, and the combination also increased the number of apoptotic cells in this model (Supplementary Fig. S12A–S12C). Collectively, these data indicate that upstream baseline HER family activation may limit the ability of MRTX849 to achieve robust inhibition of the ERK and mTOR–S6 signaling pathways. Accordingly, the combination of afatinib and MRTX849 can limit feedback reactivation of ERK and demonstrate complementary inhibition of AKT–mTOR–S6 signaling, resulting in significantly improved antitumor activity.

SHP2 inhibition has been shown to inhibit the growth of cells that harbor KRASG12C mutations, and this effect is likely mediated, in part, by decreasing KRAS GTP loading (25–27). To evaluate whether SHP2 inhibition enhanced covalent modification of KRASG12C by MRTX849, H358 and H2122 cells were incubated with increasing MRTX849 concentrations with or without the SHP2 inhibitor RMC-4550. In both cell lines, cotreatment with RMC-4550 (1 μmol/L, IC90) demonstrated a MRTX849 concentration-dependent increase in KRASG12C protein modification and a concomitant decrease in ERK phosphorylation compared with MRTX849 alone (Fig. 6D; Supplementary Fig. S13A). RPPA analysis of KRAS-dependent signaling was conducted at 6 or 24 hours after treatment in three cell lines (H358, H2030, H2122) over a range of MRTX849 concentrations in the presence or absence of RMC-4550. RMC-4550 demonstrated robust inhibition of ERK phosphorylation and partial inhibition of p90 RSK (S380) and p70 S6K (T412) at both time points (Supplementary Fig. S13B). The combination of RMC-4550 and MRTX849 demonstrated incrementally increased concentration-dependent inhibition of ERK and RSK phosphorylation in all cell lines at both 6 and 24 hours and markedly improved inhibition of S6 (S235/236) phosphorylation compared with MRTX849 alone in H2122 and H358 cells at 24 hours. In addition, the combination demonstrated near-complete inactivation of KRAS in MRTX849-refractory KYSE-410 xenografts as determined using an active RAS ELISA assay, and this was significant compared with single agents (Supplementary Fig. S13C). On the basis of these findings, combination studies were conducted with MRTX849 and RMC-4550 in six KRASG12C-mutated tumor models in vivo, and the combination demonstrated significantly greater antitumor efficacy compared with either single agent in 4 of 6 models evaluated (Fig. 6E; Supplementary Fig. S13D). Consistent with the in vitro data, the combination also demonstrated a significant decrease in ERK phosphorylation compared with either single agent in the KYSE-410 model as determined by quantitation of tumor-cell immunostaining on day 1 at 6 and 24 hours and day 7 at 6 hours after dose (Fig. 6F). Together, these data indicate that EGFR family and SHP2 blockade can augment the antitumor activity of KRASG12C inhibitors through enhancing covalent target modification and establishing a more comprehensive blockade of KRAS-dependent signaling.

Combinations That Inhibit Bypass Pathways Downstream of KRAS and Exhibit Increased Antitumor Activity in Xenograft Models

KRAS is implicated in regulation of the oncogenic S6 protein translation pathway through both ERK-dependent activation of RSK, which phosphorylates S6 at Ser235/236, and cross-talk with the PI3K and mTOR pathway that additionally phosphorylates S6 at Ser240/244 (29). However, the S6 pathway can also be activated independently of mutated KRAS in tumor cells through hyperactivated RTK signaling, PI3K activation, or STK11 mutations, each of which converge on mTOR-mediated activation of S6. In the in vitro combination screen, mTOR inhibitors demonstrated synergy in a subset of evaluated cell lines (Supplementary Fig. S14A). To further evaluate the effect of the combination on KRAS and mTOR pathway–dependent signaling, four cell lines were treated with MRTX849 in the presence or absence of the selective ATP-competitive mTOR inhibitor vistusertib (1 μmol/L), for 6 or 24 hours, and several signaling molecules were evaluated by RPPA. Vistusertib demonstrated clear and robust inhibition of several components of the PI3K–mTOR signaling pathway including AKT (S473), p70 S6K (T412), S6 (pS235/236, S240/244), and 4E-BP1 (S65, T70) phosphorylation in each cell line at both time points consistent with its mechanism of action (Supplementary Fig. S14B). MRTX849 alone did not affect 4E-BP1 or S6 (S240/244) activity, and it exhibited a variable and cell line–dependent effect on p70 S6K and S6 (pS235/236) phosphorylation in these cell lines. Vistusertib also demonstrated marked induction of ERK phosphorylation, often several-fold over vehicle control, at both time points in all four cell lines, consistent with prior reports (30). The combination of vistusertib and MRTX849 demonstrated a comparable level of inhibition of ERK phosphorylation compared with single-agent MRTX849, indicating that the activation of ERK signaling by vistusertib was impeded by the combination of the two agents. In addition, MRTX849 combined with vistusertib further inhibited p70 S6K and AKT S473 phosphorylation compared with either single agent. Near-complete inhibition of S6 (S235/236, S240/244) phosphorylation at limit of detection was observed for the combination in each cell line at evaluated time points.

Consequently, a cohort of tumor models was identified, and the combination of MRTX849 with the selective mTOR inhibitor vistusertib demonstrated marked tumor regression and significantly improved antitumor activity compared with either single agent in all six models evaluated (Fig. 7A; Supplementary Fig. S14C). MRTX849 in combination with a second, differentiated mTOR inhibitor, everolimus, which inhibits TORC1 but not TORC2, in the H2030 xenograft model also demonstrated a striking combination effect (Supplementary Fig. S14D). In the KRASG12C, STK11-mutant H2030 model, MRTX849 demonstrated marked inhibition of ERK phosphorylation through 24 hours, but exhibited only partial inhibition of pS6235/36 at 6 hours after dose, on days 1 and 7 (Fig. 7B and C). Vistusertib demonstrated marked inhibition of pS6235/36 at 6 hours after treatment with evidence of recovery by 24 hours. The combination of vistusertib and MRTX849 did not have a further effect on ERK phosphorylation but demonstrated a significant reduction in pS6235/36 on day 1 at 24 hours compared with vistusertib alone and a trend toward reduced pS6235/36 on both day 1 and day 7 at 6 hours compared with either single agent (Fig. 7B; Supplementary Fig. S14E). Together, these data indicate that MRTX849 and mTOR inhibitor combination demonstrates complementary inhibition of the ERK and mTOR–S6 signaling pathways, resulting in broad antitumor activity in KRASG12C-mutant tumor models.

Figure 7.

 

CDK4/6 and mTOR combinations suppress independently hyperactivated downstream pathways and exhibit increased antitumor responses. A, MRTX849 at 100 mg/kg, vistusertib at 15 mg/kg, or the combination was administered daily via oral gavage to mice bearing the H2122 or H2030 cell line xenografts (n = 5/group). Combination treatment led to a statistically significant decrease in tumor growth compared with either single-agent treatment. *, Padj < 0.05. B, MRTX849 at 100 mg/kg, vistusertib at 15 mg/kg, or the combination was administered once or daily for 7 days via oral gavage to mice bearing H2030 cell line xenografts (n = 3/group). Tumors were harvested at 6 and 24 hours following the final dose. Tumor sections were stained with pERK and pS6 via IHC methods. Quantitation of images shown by H-score in tumor tissue. Reduction of pERK or pS6 staining intensity was determined to be statistically significant relative to vehicle or either single agent using one-way ANOVA. Brackets indicate P < 0.05 compared with left-most sample. C, Protein Western blot analysis of KRAS pathway targets in H2030 xenografts treated with MRTX849 (100 mg/kg), vistusertib (15 mg/kg), or the combination, 6 or 24 hours after a single dose. D, Protein Western blot analysis of KRAS pathway and cell-cycle targets in H2122 cells treated for 24 hours with MRTX849, palbociclib, or the combination. E, Normalized RNA-seq gene-expression data on E2F targets in H2122 xenografts treated with MRTX849, palbociclib, or the combination, 6 and 24 hours after a single daily dose or seven daily doses. F, MRTX849 at 100 mg/kg, palbociclib at 130 mg/kg, or the combination was administered daily via oral gavage to mice bearing the H2122 or SW1573 cell line xenografts (n = 5). Combination treatment led to a statistically significant decrease in tumor growth compared with either single-agent treatment. *, Padj < 0.05.

 

Signaling through KRAS is known to mediate cell proliferation, at least in part, through the regulation of the cyclin D family and triggering RB/E2F-dependent entry of cells into cell cycle. Loss-of-function mutations and homozygous deletions in the cell-cycle tumor suppressor CDKN2A (p16) are coincident in a subset of KRAS-mutant non–small cell lung cancer (NSCLC) and hyperactivate CDK4/6-dependent RB phosphorylation and cell-cycle transition. In the CDKN2A-null H2122 and SW1573 cell lines in vitro, MRTX849 demonstrated concentration-dependent partial inhibition of RB phosphorylation (pRB pS807/811) and concurrent increase in p27 in H2122 cells, but not SW1573 cells, at 24 hours (Fig. 7D; Supplementary Fig. S15A). MRTX849 in combination with the CDK4/6 inhibitor palbociclib (1 μmol/L) demonstrated near-complete inhibition of pRB in both H2122 and SW1573 cells and further induced p27 in H2122 cells. Interestingly, pS6 (S235/236) was also much more effectively suppressed by the combination in both H2122 and SW1573 cells, which is consistent with a recent report (31). RNA expression of target genes and RPPA analysis of target protein signaling events were also used as a readout of cell-cycle inhibition in the H2122 tumor model in vivo, and the combination of MRTX849 and palbociclib significantly inhibited E2F1 and selected E2F family target genes, induced p27 protein expression to a greater degree compared with either single agent, and further reduced the number of Ki-67–positive cells after 7 days of administration (Fig. 7E; Supplementary Fig. S15B and S15C). In addition, the combination demonstrated a significant decrease in pRB (S780) compared with either single agent after 7 days of administration in SW1573 tumors in vivo (Supplementary Fig. S15D). This combination also induced tumor regression in five tumor xenograft models that was significant compared with either single-agent control (Fig. 7F; Supplementary Fig. S15E). Although not significant, a trend was noted in which models with CDKN2A homozygous deletion exhibited an increased antitumor response to the combination of MRTX849 and CDK4/6 inhibition compared with models lacking evidence of genetic dysregulation of key cell-cycle genes (Supplementary Fig. S15F and S15G).

Discussion

The identification of MRTX849 as a highly selective KRASG12C inhibitor capable of near-complete inhibition of KRAS in vivo provides a renewed opportunity to better understand the role of this mutation as an oncogenic driver in various cancers and to guide rational clinical trial design. The lack of a significant correlation between sensitivity to MRTX849 antitumor activity in in vitro versus in vivo model systems made it necessary to further study KRAS oncogene dependence in tumor models in vivo, a more clinically relevant setting. The demonstration that MRTX849 exhibited significant antitumor efficacy in all evaluated KRASG12C-mutated cancer models and demonstrated marked regression in the majority (65%) confirms that this mutation is a broadly operative oncogenic driver and that MRTX849 represents a compelling therapeutic opportunity. This evidence of activity extended to patients, as demonstrated by RECIST partial responses in 2 patients enrolled in a phase I clinical trial of MRTX849. Collectively however, these data also illustrate that the degree of dependence of cancer cells on the presence of a KRASG12C mutation for growth and survival can vary across tumors and that co-occurring genetic alterations observed in KRAS-mutated cancers may influence response to direct targeted therapy. The further observation that KRAS mutations occur across different cancers and that no single co-occurring genetic alteration predicted response to treatment illustrates the genetic heterogeneity of KRAS-driven cancers. Findings in the present studies are consistent with other functional genomics or therapeutic strategies to block KRAS function across panels of cell lines or models which demonstrated a highly significant response of KRAS-mutant cells to target knockdown, a heterogeneous magnitude of response, and no clear co-occurring aberrations that predict resistance to target blockade (5, 32, 33). Interestingly, despite the implication that certain mutations that co-occur with KRAS including TP53, STK11, and KEAP1 may limit therapeutic response in KRASG12C-positive lung cancers, none of these mutations correlated with response or resistance in the cell-line panel. In addition, the partial response we reported in the patient with lung adenocarcinoma was observed in a patient harboring deleterious comutations in both STK11 and KEAP1. Together, these data further illustrate the heterogeneity and complexity of KRAS-mutated cancers and suggest that no binary co-occurring genetic event may be predictive of therapeutic response.

Temporal and dose–response analysis indicated maximal modification of KRASG12C and durable inhibition of KRAS-dependent signaling was important in maximizing therapeutic response. The recovery of ERK signaling and the inability to inhibit mTOR–S6 signaling despite continued treatment were each associated with transient or submaximal response to MRTX849. ERK1/2 is implicated in direct phosphorylation and negative feedback regulation of EGFR (T669), FGFR1 (S777), and SOS1, and each of these targets may facilitate KRASG12C-independent resetting of ERK signaling flux (34–36). The rapid and remarkable suppression of ERK pathway–regulated transcripts such as DUSP and SPRY/SPRED family members by MRTX849 in all models evaluated is consistent with that observed for RAF inhibitors and is implicated in reactivation of ERK and RTK signaling (18, 19). The dual-specificity phosphatases DUSP4 and 6 were strongly suppressed by MRTX849 and are implicated in dephosphorylating and inactivating ERK1/2 (14, 18, 37), whereas SPRY family members are implicated in the negative regulation of RTKs and adaptor proteins (e.g., GRB2), and may participate in modifying RAS family nucleotide exchange and effector binding (e.g., RAF1; ref. 38). Although suppression of DUSP and SPRY/SPRED was broadly observed in all models, the magnitude of signaling reactivation and response to MRTX849 varied across models. This suggests some tumor models harbor additional factors that bypass KRAS dependence or affect RAS pathway signaling flux, such as expression or activation of selected RTKs (e.g., ERBB2 amplification in the KYSE-410 model) or STK11 loss-of-function mutations, and may be primed for feedback reactivation of RAS-dependent signaling and/or limit the degree of signaling inhibition by MRTX849. This phenomenon was observed for BRAFV600E-mutant colon cancer (but not melanoma) which exhibits high baseline EGFR expression, is primed for rapid feedback activation of this RTK, and is resistant to single-agent inhibition but highly responsive to cotargeting BRAF (and/or MEK) and EGFR (20). In addition, blockade of BRAF or MEK1/2 resulted in feedback-mediated activation of the PI3K–mTOR signaling pathway in concert with the coactivation of upstream RTKs (e.g., EGFR), resulting in bypass of ERK pathway dependence and therapeutic resistance (17, 20, 39). The observations that baseline expression of HER family RTKs trended with MRTX849 antitumor activity and that CRISPR-based drug-anchored screens implicated EGFR, SHP2, and mTOR–S6 pathways as cotargetable vulnerabilities both support the hypothesis that these targets act as conditional response modifiers.

Activation of RTK signaling in the context of KRASG12C-mutant cancer was predicted to limit MRTX849 therapeutic response both by enhancing extrinsic regulation of GTPase activity and initiating KRAS-independent ERK and mTOR–S6 pathway activation. Therefore, HER family and SHP2 inhibition were employed as strategies to either block the critical RTK family in KRAS-mutant cells or block collective RTK signaling downstream, respectively. As MRTX849 binds only GDP-KRASG12C, both HER family and SHP2 inhibition each enhanced KRASG12C modification by MRTX849 and significantly improved antitumor activity. This observation is consistent with the putative role of activated RTKs in the engagement of SHP2 to mediate SOS1-dependent RAS GTP loading and to diminish RAS GAP activity, each of which converge on enhanced RAS activation state (40). The afatinib combination demonstrated a clear and marked inhibition of both the ERK–RSK and AKT–mTOR–S6 signaling pathways, whereas the SHP2 inhibitor combination demonstrated a clear impact on ERK–RSK signaling and a relatively less prominent impact on mTOR–S6 signaling. Although afatinib may more effectively address mTOR–S6 bypass signaling, SHP2 inhibition should be an effective combinatorial strategy to combat other RTKs outside of the HER family, such as FGFRs or MET, that could affect KRAS dependence. To further address bypass signaling mediated by RTK activation or STK11 mutations, each of which activate the mTOR–S6 signaling pathway independently of KRAS, mTOR inhibition in combination with MRTX849 was also evaluated. MRTX849 in combination with vistusertib, in fact, demonstrated significantly improved antitumor activity in vivo compared with either single agent in all six tumor models evaluated, regardless of STK11 mutational status. Consistent with the mechanism of action of vistusertib, comprehensive inhibition of AKT–mTOR–S6 signaling was observed for vistusertib alone and near-complete inhibition of pS6S235–36 and pS6240–44 was observed in combination. In addition, the marked feedback reactivation of ERK by vistusertib was relieved by the combination. The induction of ERK activity has been observed in tumor cells following mTORC1 inhibition by rapalogs or ATP-competitive inhibitors and has been implicated in limiting antitumor activity of this class of agents (30, 41, 42), supporting the suppression of ERK signaling by MRTX849 as a key mechanism of response to the combination. Notably, all three combination strategies converge on more comprehensive inhibition of KRAS-dependent signaling, converging on ERK and S6 activity. In addition, although the inhibition of the AKT–mTOR–S6 pathway did not correlate with model response to MRTX849 (potentially due to tumor heterogeneity), the observations that both MTOR and RPS6 drop out in drug-anchored CRISPR screens and that effective combination strategies more comprehensively block this pathway illustrate its likely importance in maximizing therapeutic response in KRAS-mutated cancers.

Cell-cycle dysregulation due to genetic alterations in cell-cycle regulators identified additional factors that could modify the therapeutic response to MRTX849. In addition, CDKN2A, RB1, CDK4, and CDK6 were all identified as gene targets that affected cell fitness in CRISPR screens. Genetic alterations including homozygous deletion of CDKN2A or amplification of CDK4 or CCND1 comprise up to 20% of KRAS-mutated NSCLC (43). Combination studies with MRTX849 and palbociclib in vivo demonstrated more comprehensive inhibition of RB and E2F family target genes and increased antitumor activity compared with either single agent in NSCLC models. In addition, these studies indicated that the combination resulted in more effective inhibition of S6 (S235/236) phosphorylation, establishing a previously unappreciated connection between cell-cycle blockade and protein translation pathways. Notably, this combination was especially effective in CDKN2A-deleted models, suggesting that this combination strategy may be primarily beneficial in a molecularly defined subset of patients characterized by decoupling of cell-cycle regulation from KRAS.

Collectively, models exhibiting a cytoreductive response to single-agent MRTX849 demonstrated a more comprehensive and durable inhibition of KRAS-dependent signaling and induction of an apoptotic response. These data suggest that maintaining durable inhibition of KRAS-dependent signaling below a defined threshold is required to elicit tumor regression. The elucidation of mechanisms that limit the therapeutic response to single-agent KRAS inhibition has provided insight toward strategies to enhance therapeutic activity in KRAS-mutant tumors. Of the 35% of models (9/26) that did not exhibit durable regression with single-agent MRTX849 treatment, five models (KYSE410, SW1573, H2122, H2030, and LU6405) were selected for rational combination studies, and at least one combination demonstrated significant improvement in antitumor efficacy and elicited a >50% tumor regression in all five models evaluated. These results suggest that essentially all KRASG12C-mutated cancers can derive clinical benefit from direct KRAS inhibitor–directed therapy either alone or in combination. Furthermore, rational pathway-centric combination regimens directed at hallmark signaling nodes may be directed to genetically defined patient subsets. For example, KRAS-mutated NSCLC exhibits mutually exclusive, co-occurring genetic alterations in STK11 and CDKN2A (43). The present data suggest that KRASG12C/STK11-mutated NSCLC could be readily addressed by combining a KRASG12C inhibitor with an RTK or mTOR inhibitor, whereas KRASG12C/CDKN2A-mutated NSCLC could be more effectively addressed by combination with a CDK4/6 inhibitor. Collectively, the present studies support the broad utility of covalent KRASG12C inhibitors in treating KRASG12C-mutated cancers and provide defining strategies to identify patients likely to benefit from single-agent therapy or rationally directed combinations.

  

UPDATED 02/07/2021

The November 1st issue of Science highlights a series of findings which give cancer researchers some hope in finally winning a thirty year war with the discovery of drugs that target KRAS, one of the most commonly mutated oncogenes  (25% of cancers), and thought to be a major driver of tumorigenesis. Once considered an undruggable target, mainly because of the smooth surface with no obvious pockets to fit a drug in, as well as the plethora of failed attempts to develop such an inhibitor, new findings with recently developed candidates, highlighted in this article and other curated within, are finally giving hope to researchers and oncologists who have been hoping for a clinically successful inhibitor of this once considered elusive target.

For a great review on development of G12C KRas inhibitors please see Dr. Hobb’s and Channing Der’s review in Cell Selective Targeting of the KRAS G12C Mutant: Kicking KRAS When It’s Down

Figure 1Mechanism of Action of ARS853 showing that the inhibitors may not need bind to the active conformation of KRAS for efficacy

Abstract: Two recent studies evaluated a small molecule that specifically binds to and inactivates the KRAS G12C mutant. The new findings argue that the perception that mutant KRAS is persistently frozen in its active GTP-bound form may not be accurate.

Although the development of the KRASG12C-specific inhibitor, compound 12 (Ostrem et al., 2013), was groundbreaking, subsequent studies found that the potency of compound 12 in cellular assays was limited (Lito et al., 2016, Patricelli et al., 2016). A search for more-effective analogs led to the development of ARS853 (Patricelli et al., 2016), which exhibited a 600-fold increase of its reaction rate in vitro over compound 12 and cellular activities in the low micromolar range.

A Summary and more in-depth curation of the Science article is given below:

After decades, progress against an ‘undruggable’ cancer target

Summary

Cancer researchers are making progress toward a goal that has eluded them for more than 30 years: shrinking tumors by shutting off a protein called KRAS that drives growth in many cancer types. A new type of drug aimed at KRAS made tumors disappear in mice and shrank tumors in lung cancer patients, two companies report in papers published this week. It’s not yet clear whether the drugs will extend patients’ lives, but the results are generating a wave of excitement. And one company, Amgen, reports an unexpected bonus: Its drug also appears to stimulate the immune system to attack tumors, suggesting it could be even more powerful if paired with widely available immunotherapy treatments.

Jocelyn Kaiser. After decades, progress against an ‘undruggable’ cancer target. Science  01 Nov 2019: Vol. 366, Issue 6465, pp. 561 DOI: 10.1126/science.366.6465.561

The article highlights the development of three inhibitors: by Wellspring Biosciences, Amgen, and Mirati Therapeutics.

Wellspring BioSciences

In 2013, Dr. Kevan Shokat’s lab at UCSF discovered a small molecule that could fit in the groove of the KRAS mutant G12C.  The G12C as well as the G12D is a common mutation found in KRAS in cancers. KRAS p.G12C mutations predominate in NSCLC comprising 11%–16% of lung adenocarcinomas (45%–50% of mutant KRAS is p.G12C) (Campbell et al., 2016; Jordan et al., 2017), as well as 1%–4% of pancreatic and colorectal adenocarcinomas, respectively (Bailey et al., 2016; Giannakis et al., 2016).  This inhibitor was effective in shrinking, in mouse studies conducted by Wellspring Biosciences,  implanted tumors containing this mutant KRAS.

See Wellspring’s news releases below:

March, 2016 – Publication – Selective Inhibition of Oncogenic KRAS Output with Small Molecules Targeting the Inactive State

 

February, 2016 – Publication – Allele-specific inhibitors inactivate mutant KRAS G12C by a trapping mechanism

 

Amgen

Amgen press release on AMG510 Clinical Trial at ASCO 2019

THOUSAND OAKS, Calif., June 3, 2019 /PRNewswire/ — Amgen (NASDAQ: AMGN) today announced the first clinical results from a Phase 1 study evaluating investigational AMG 510, the first KRASG12C inhibitor to reach the clinical stage. In the trial, there were no dose-limiting toxicities at tested dose levels. AMG 510 showed anti-tumor activity when administered as a monotherapy in patients with locally-advanced or metastatic KRASG12C mutant solid tumors. These data are being presented during an oral session at the 55th Annual Meeting of the American Society of Clinical Oncology (ASCO) in Chicago.

“KRAS has been a target of active exploration in cancer research since it was identified as one of the first oncogenes more than 30 years ago, but it remained undruggable due to a lack of traditional small molecule binding pockets on the protein. AMG 510 seeks to crack the KRAS code by exploiting a previously hidden groove on the protein surface,” said David M. Reese, M.D., executive vice president of Research and Development at Amgen. “By irreversibly binding to cysteine 12 on the mutated KRAS protein, AMG 510 is designed to lock it into an inactive state. With high selectivity for KRASG12C, we believe investigational AMG 510 has high potential as both a monotherapy and in combination with other targeted and immune therapies.”

The Phase 1, first-in-human, open-label multicenter study enrolled 35 patients with various tumor types (14 non-small cell lung cancer [NSCLC], 19 colorectal cancer [CRC] and two other). Eligible patients were heavily pretreated with at least two or more prior lines of treatment, consistent with their tumor type and stage of disease. 

Canon, J., Rex, K., Saiki, A.Y. et al. The clinical KRAS(G12C) inhibitor AMG 510 drives anti-tumour immunity. Nature 575, 217–223 (2019) doi:10.1038/s41586-019-1694-1

Besides blocking tumor growth, AMG510 appears to stimulate T cells to attack the tumor, thus potentially supplying a two pronged attack to the tumor, inhibiting oncogenic RAS and stimulating anti-tumor immunity.

Mirati Therapeutics

Mirati’s G12C KRAS inhibitor (MRTX849) is being investigated in a variety of solid malignancies containing the KRAS mutation.

For recent publication on results in lung cancer see Patricelli M.P., et al. Cancer Discov. 2016; (Published online January 6, 2016)

For more information on Mirati’s KRAS G12C inhibitor see https://www.mirati.com/pipeline/kras-g12c/

KRAS G12C Inhibitor (MRTX849)

Study 849-001 – Phase 1b/2 of single agent MRTX849 for solid tumors with KRAS G12C mutation

Phase 1b/2 clinical trial of single agent MRTX849 in patients with advanced solid tumors that have a KRAS G12C mutation.

See details for this study at clinicaltrials.gov

UPDATED 02/07/2021

Amgen scientists’ rapid work to challenge the undruggable KRAS G12C mutation in cancer

Inside a 40-year quest to challenge the KRAS G12C mutation in cancer
By Amgen Oncology
Amgen’s sotorasib, an investigational lung cancer treatment, has been submitted to the FDA for review

 

 

Nearly four decades have passed since researchers first identified the RAS gene family, which includes HRASNRAS and KRASRAS is the most frequently mutated family of oncogenes – or potentially cancerous genes – in human cancers.1,2 While research  efforts have been able to identify and develop treatments for other driver gene mutations that contribute to cancer growth, success with treating KRAS, the most frequently mutated variant of the RAS family, has remained elusive.2 But now there is hope.

Amgen, a leading biotechnology company, has taken on one of the toughest challenges of the last 40 years in cancer research.3 Chemical biologist Kevan Shokat’s lab at the University of California, San Francisco, identified a small molecule that could slip into a groove on a KRAS mutation called G12C in 2013.4 Building on their own research strategies and this new insight, scientists at Amgen used structural biology and medicinal chemistry to identify an adjacent groove, and by November 2017, made the initial decision to advance the molecule that would become investigational sotorasib.5

KRAS G12C is the most common KRAS mutation in NSCLC.6,7 In the U.S., about 13% of patients with NSCLC harbor the KRAS G12C mutation.8 There is a high unmet need and poor outcomes in the second-line treatment of KRAS G12C-driven non-small cell lung cancer (NSCLC) and, currently, there are no KRASG12C targeted therapies approved.

According to Amgen’s head of research and development David Reese, “the company’s scientists had an idea some time ago that the future of oncology would be led by the marriage of immuno-oncology and precision therapy. We wanted to go after high value targets, and RAS proteins are one of them.”

Because of this effort to rapidly accelerate the speed of innovation, investigational sotorasib entered clinical trials in humans less than 12 months.

 

 

At the same time that scientists discovered investigational sotorasib, the team was undertaking a project to map out every step it takes to progress a potential new treatment from an idea in a lab to being made available for patients. The goal was to shrink timelines and eliminate gaps to develop drugs more rapidly in order to reach patients with serious illnesses like NSCLC as quickly as possible.

Because of this effort to rapidly accelerate the speed of innovation, sotorasib entered clinical trials in humans less than 12 months.5 Sotorasib was the first investigational KRASG12C inhibitor to enter the clinic, and is now being studied in the broadest clinical program exploring 10 combinations with global investigational sites spanning five continents.9 In a little more than two years, the sotorasib clinical program has established a clinical data set of more than 700 patients studied across 13 tumor types.9

The investigational treatment was recently submitted to the FDA for review and was granted Breakthrough Therapy designation, a distinction designed to expedite the development and review of drugs.5 It was also accepted into the FDA’s Real-Time Oncology Review pilot program, which aims to explore a more efficient review process.5

To learn more about Amgen and how the speed of innovation is bringing new oncology treatments to patients with high unmet needs, visit Amgen.com/KnowKRAS.  

______________________

1 Ryan MB, et al. Nat Rev Clin Oncol. 2018;15:709-720.

2 Cox AD, et al. Nat Rev Drug Discov. 2014;13:828-851.

3 Kim D, et al. Cell. 2020. doi:10.1016/j.cell.2020.09.044.

4 Ostrem JM, et al. Nature. 2013 ; 503 :548-551.

5 AMGEN, 2020. Retrieved January 8, 2021, from https://www.amgen.com/stories/2020/12/rapidly-advancing-development-of-amgens-investigational-kras-g12c-inhibitor

6 Pakkala S, et al. JCI Insight. 2018;3:e120858.

7 Arbour KC, et al. Clin Cancer Res. 2018;24:334-340.

8 Amgen, Data on File. 2020.

9 ClinicalTrials.gov. NCT04185883, NCT04380753, NCT03600883, NCT04303780. https://clinicaltrials.gov/ct2/. Accessed January 20, 2020.

 
 
Members of the editorial and news staff of the USA TODAY Network were not involved in the creation of this content.
 
 
 

Additional References:

Allele-specific inhibitors inactivate mutant KRAS G12C by a trapping mechanism.

Lito P et al. Science. (2016)

Targeting KRAS Mutant Cancers with a Covalent G12C-Specific Inhibitor.

Janes MR et al. Cell. (2018)

Potent and Selective Covalent Quinazoline Inhibitors of KRAS G12C.

Zeng M et al. Cell Chem Biol. (2017)

Campbell, J.D., Alexandrov, A., Kim, J., Wala, J., Berger, A.H., Pedamallu, C.S., Shukla, S.A., Guo, G., Brooks, A.N., Murray, B.A., et al.; Cancer Genome Atlas Research Network (2016). Distinct patterns of somatic genome alterations in lung adenocarcinomas and squamous cell carcinomas. Nat. Genet.48, 607–616

Jordan, E.J., Kim, H.R., Arcila, M.E., Barron, D., Chakravarty, D., Gao, J., Chang, M.T., Ni, A., Kundra, R., Jonsson, P., et al. (2017). Prospective comprehensive molecular characterization of lung adenocarcinomas for efficient patient matching to approved and emerging therapies. Cancer Discov. 7, 596–609.

Bailey, P., Chang, D.K., Nones, K., Johns, A.L., Patch, A.M., Gingras, M.C., Miller, D.K., Christ, A.N., Bruxner, T.J., Quinn, M.C., et al.; Australian Pancreatic Cancer Genome Initiative (2016). Genomic analyses identify molecular subtypes of pancreatic cancer. Nature 531, 47–52.

Giannakis, M., Mu, X.J., Shukla, S.A., Qian, Z.R., Cohen, O., Nishihara, R., Bahl, S., Cao, Y., Amin-Mansour, A., Yamauchi, M., et al. (2016). Genomic correlates of immune-cell infiltrates in colorectal carcinoma. Cell Rep. 15, 857–865.

Read Full Post »

Single-cell RNA-seq helps in finding intra-tumoral heterogeneity in pancreatic cancer

Reporter and Curator: Dr. Sudipta Saha, Ph.D.

4.3.6

4.3.6  Single-cell RNA-seq helps in finding intra-tumoral heterogeneity in pancreatic cancer, Volume 2 (Volume Two: Latest in Genomics Methodologies for Therapeutics: Gene Editing, NGS and BioInformatics, Simulations and the Genome Ontology), Part 4: Single Cell Genomics

Pancreatic cancer is a significant cause of cancer mortality; therefore, the development of early diagnostic strategies and effective treatment is essential. Improvements in imaging technology, as well as use of biomarkers are changing the way that pancreas cancer is diagnosed and staged. Although progress in treatment for pancreas cancer has been incremental, development of combination therapies involving both chemotherapeutic and biologic agents is ongoing.

Cancer is an evolutionary disease, containing the hallmarks of an asexually reproducing unicellular organism subject to evolutionary paradigms. Pancreatic ductal adenocarcinoma (PDAC) is a particularly robust example of this phenomenon. Genomic features indicate that pancreatic cancer cells are selected for fitness advantages when encountering the geographic and resource-depleted constraints of the microenvironment. Phenotypic adaptations to these pressures help disseminated cells to survive in secondary sites, a major clinical problem for patients with this disease.

The immune system varies in cell types, states, and locations. The complex networks, interactions, and responses of immune cells produce diverse cellular ecosystems composed of multiple cell types, accompanied by genetic diversity in antigen receptors. Within this ecosystem, innate and adaptive immune cells maintain and protect tissue function, integrity, and homeostasis upon changes in functional demands and diverse insults. Characterizing this inherent complexity requires studies at single-cell resolution. Recent advances such as massively parallel single-cell RNA sequencing and sophisticated computational methods are catalyzing a revolution in our understanding of immunology.

PDAC is the most common type of pancreatic cancer featured with high intra-tumoral heterogeneity and poor prognosis. In the present study to comprehensively delineate the PDAC intra-tumoral heterogeneity and the underlying mechanism for PDAC progression, single-cell RNA-seq (scRNA-seq) was employed to acquire the transcriptomic atlas of 57,530 individual pancreatic cells from primary PDAC tumors and control pancreases. The diverse malignant and stromal cell types, including two ductal subtypes with abnormal and malignant gene expression profiles respectively, were identified in PDAC.

The researchers found that the heterogenous malignant subtype was composed of several subpopulations with differential proliferative and migratory potentials. Cell trajectory analysis revealed that components of multiple tumor-related pathways and transcription factors (TFs) were differentially expressed along PDAC progression. Furthermore, it was found a subset of ductal cells with unique proliferative features were associated with an inactivation state in tumor-infiltrating T cells, providing novel markers for the prediction of antitumor immune response. Together, the findings provided a valuable resource for deciphering the intra-tumoral heterogeneity in PDAC and uncover a connection between tumor intrinsic transcriptional state and T cell activation, suggesting potential biomarkers for anticancer treatment such as targeted therapy and immunotherapy.

References:

https://www.ncbi.nlm.nih.gov/pubmed/31273297

https://www.ncbi.nlm.nih.gov/pubmed/21491194

https://www.ncbi.nlm.nih.gov/pubmed/27444064

https://www.ncbi.nlm.nih.gov/pubmed/28983043

https://www.ncbi.nlm.nih.gov/pubmed/24976721

https://www.ncbi.nlm.nih.gov/pubmed/27693023

Read Full Post »

New Targeted Cancer Therapy may be ‘Possible Hope’ for Some Pancreatic Cancer Patients

Reporter: Irina Robu, PhD

 

UPDATED on 7/18/2019

BREAKTHROUGH PANCREATIC CANCER TREATMENT PHASE III TRIAL OPENS IN ISRAEL

Hope is that successful trials will allow Rafael Pharmaceuticals will receive expedited FDA approval by late 2020.

BY MAAYAN JAFFE-HOFFMAN  JULY 18, 2019 18:30

“What it does is feeds misinformation to these regulatory elements, making them feel that there is too much carbon flow through both of these complexes, causing them to be inhibited,” Pardee said. “It simultaneously inhibits both complexes so tumor cells that are primarily driven by glucose cannot utilize glucose in the TCA cycle. Tumor cells that are primarily driven by glutamine usage cannot use glutamine-derived carbons in the TCA cycle. And, importantly, tumors cannot switch from one source to the other in the presence of CPI-613,” he explained.

He said that hitting two complexes simultaneously has many advantages. One is that the carbon source the tumor is primarily dependent on does not matter; another is that evolved resistance for both complexes simultaneously is very unlikely to happen.

Pardee said CPI-613’s key differentiators are that it is highly selective on the uptake and target level in cancer cells, which leads to less toxicity to healthy cells. This allows for patients to receive extended treatment courses and for the drug to be used in combination with other drugs.

CPI-613 is being administered in this clinical trial with a chemotherapy combination of fluorouracil, leucovorin, irinotecan, and oxaliplatin, called FOLFIRINOX.

SOURCE

https://www.jpost.com/HEALTH-SCIENCE/Breakthrough-pancreatic-cancer-treatment-phase-III-trial-opens-in-Israel-596059

 

New Targeted Cancer Therapy may be ‘Possible Hope’ for Some Pancreatic Cancer Patients

Pancreatic cancer is the 12th maximum common cancer and the fourth leading cause of cancer death. The cancer is often difficult to diagnose as there is no cost-effective ways to screen for the illness. For over 52% of people who are diagnosed after the cancer has spread and with a 5-year survival rate.

Scientists at Sheba Medical Center in Israel developed a targeted cancer therapy drug together with AstraZeneca and Merck which can offer a possible new solution for patients with a specific kind of pancreatic cancer by delaying the progression of the disease. To evaluate the safety and test the efficacy of a new drug treatment regimen based on Lynparza tablets. The tablets are a pharmacological inhibitor of the enzyme poly (ADP-ribose) polymerase which inhibit the enzyme. They were developed for a number of indications, but most prominently for the treatment of cancer, as numerous forms of cancer are more dependent for their development on the enzyme than regular cells are. This makes poly (ADP-ribose) polymerase an attractive target for cancer therapy.

Their study included 154 patients who were randomly assigned to get the tablets at a dose of 300 mg twice a day with metastatic pancreatic cancer who carried the genetic mutation called BRCA 1 and BRCA 2. BRCA1 and BRCA2 are human genes that produce proteins accountable for repairing damaged DNA and play a substantial role in preserving the genetic stability of cells. Once either of these genes is mutated, DNA damage can’t be repaired properly and cells become unstable. As a result, cells are more likely to develop additional genetic alterations that can lead to cancer.

Patients with these mutations make up six to seven percent of the metastatic pancreatic cancer patients. The trial using the using the medicine Lynparza offers possible hope for those who suffer from metastatic pancreatic cancer and have a BRCA mutation and slows down the disease progression. According to the researchers this is the first Phase 3 biomarker that is positive in pancreatic cancer and the drug gives incredible hope for patients with the advanced stage of the cancer.

SOURCE
https://www.timesofisrael.com/israeli-researchers-find-potential-hope-for-some-pancreatic-cancer-patients/

Read Full Post »

Pancreatic cancer survival is determined by ratio of two enzymes, Volume 2 (Volume Two: Latest in Genomics Methodologies for Therapeutics: Gene Editing, NGS and BioInformatics, Simulations and the Genome Ontology), Part 1: Next Generation Sequencing (NGS)

Reporter and Curator: Dr. Sudipta Saha, Ph.D.

 

Protein kinase C (PKC) isozymes function as tumor suppressors in increasing contexts. These enzymes are crucial for a number of cellular activities, including cell survival, proliferation and migration — functions that must be carefully controlled if cells get out of control and form a tumor. In contrast to oncogenic kinases, whose function is acutely regulated by transient phosphorylation, PKC is constitutively phosphorylated following biosynthesis to yield a stable, autoinhibited enzyme that is reversibly activated by second messengers. Researchers at University of California San Diego School of Medicine found that another enzyme, called PHLPP1, acts as a “proofreader” to keep careful tabs on PKC.

 

The researchers discovered that in pancreatic cancer high PHLPP1 levels lead to low PKC levels, which is associated with poor patient survival. They reported that the phosphatase PHLPP1 opposes PKC phosphorylation during maturation, leading to the degradation of aberrantly active species that do not become autoinhibited. They discovered that any time an over-active PKC is inadvertently produced, the PHLPP1 “proofreader” tags it for destruction. That means the amount of PHLPP1 in patient’s cells determines his amount of PKC and it turns out those enzyme levels are especially important in pancreatic cancer.

 

This team of researchers reversed a 30-year paradigm when they reported evidence that PKC actually suppresses, rather than promotes, tumors. For decades before this revelation, many researchers had attempted to develop drugs that inhibit PKC as a means to treat cancer. Their study implied that anti-cancer drugs would actually need to do the opposite — boost PKC activity. This study sets the stage for clinicians to one day use a pancreatic cancer patient’s PHLPP1/PKC levels as a predictor for prognosis, and for researchers to develop new therapeutic drugs that inhibit PHLPP1 and boost PKC as a means to treat the disease.

 

The ratio — high PHLPP1/low PKC — correlated with poor prognoses: no pancreatic patient with low PKC in the database survived longer than five-and-a-half years. On the flip side, 50 percent of the patients with low PHLPP1/high PKC survived longer than that. While still in the earliest stages, the researchers hope that this information might one day aid pancreatic diagnostics and treatment. The researchers are next planning to screen chemical compounds to find those that inhibit PHLPP1 and restore PKC levels in low-PKC-pancreatic cancer cells in the lab. These might form the basis of a new therapeutic drug for pancreatic cancer.

 

References:

 

https://health.ucsd.edu/news/releases/Pages/2019-03-20-two-enzymes-linked-to-pancreatic-cancer-survival.aspx?elqTrackId=b6864b278958402787f61dd7b7624666

 

https://www.ncbi.nlm.nih.gov/pubmed/30904392

 

https://www.ncbi.nlm.nih.gov/pubmed/29513138

 

https://www.ncbi.nlm.nih.gov/pubmed/18511290

 

https://www.ncbi.nlm.nih.gov/pubmed/28476658

 

https://www.ncbi.nlm.nih.gov/pubmed/28283201

 

https://www.ncbi.nlm.nih.gov/pubmed/24231509

 

https://www.ncbi.nlm.nih.gov/pubmed/28112438

 

Read Full Post »

One blood sample can be tested for a comprehensive array of cancer cell biomarkers: R&D at WPI

Curator: Marzan Khan, B.Sc

 

A team of mechanical engineers at Worcester Polytechnic Institute (WPI) have developed a fascinating technology – a liquid biopsy chip that captures and detects metastatic cancer cells, just from a small blood sample of cancer patients(1). This device is a recent development in the scientific field and holds tremendous potential that will allow doctors to spot signs of metastasis for a variety of cancers at an early stage and initiate an appropriate course of treatment(1).

Metastasis occurs when cancer cells break away from their site of origin and spread to other parts of the body via the lymph or the bloodstream, where they give rise to secondary tumors(2). By this time, the cancer is at an advanced stage and it becomes increasingly difficult to fight the disease. The cells that are shed by primary and metastatic cancers are called circulating tumor cells (CTCs) and their numbers lie in the range of 1–77,200/m(3). The basis of the liquid biopsy chip test is to capture these circulating tumor cells in the patient’s blood and identify the cell type through specific interaction with antibodies(4).

The chip is comprised of individual test units or small elements, about 3 millimeters wide(4). Each small element contains a network of carbon nanotube sensors in a well which are functionalized with antibodies(4). These antibodies will bind cell-surface antigens or protein markers unique for each type of cancer cell. Specific interaction between a cell surface protein and its corresponding antibody is a thermodynamic event that causes a change in free energy which is transduced into electricity(3). This electrical signature is picked up by the semi-conducting carbon nanotubes and can be seen as electrical spikes(4). Specific interactions create an increase in electrical signal, whereas non-specific interactions cause a decrease in signal or no change at all(4). Capture efficiency of cancer cells with the chip has been reported to range between 62-100%(4).

The liquid biopsy chip is also more advanced than microfluidics for several reasons. Firstly, the nanotube-chip arrays can capture as well as detect cancer cells, while microfluidics can only capture(4). Samples do not need to be processed for labeling or fixation, so the cell structures are preserved(4). Unlike microfluidics, these nanotubes will also capture tiny structures called exosomes spanning the nanometer range that are produced from cancer cells and carry the same biomarkers(4).

Pancreatic cancer is the fourth leading cause of cancer-associated deaths in the United states, with a survival window of 5 years in only 6% of the cases with treatment(5). In most patients, the disease has already metastasized at the time of diagnosis due to the lack of early-diagnostic markers, affecting some of the major organs such as liver, lungs and the peritoneum(5,6). Despite surgical resection of the primary tumor, the recurrence of local and metastatic tumors is rampant(5). Metastasis is the major cause of mortality in cancers(5). The liquid biopsy chip, that identifies CTCs can thus become an effective diagnostic tool in early detection of cancer as well as provide information into the efficacy of treatment(3). At present, ongoing experiments with this device involve testing for breast cancers but Dr. Balaji Panchapakesan and his team of engineers at WPI are optimistic about incorporating pancreatic and lung cancers into their research.

REFERENCES

1.Nanophenotype. Researchers build liquid biopsy chip that detects metastatic cancer cells in blood: One blood sample can be tested for a comprehensive array of cancer cell biomarkers. 27 Dec 2016. Genesis Nanotechnology,Inc

https://genesisnanotech.wordpress.com/2016/12/27/researchers-build-liquid-biopsy-chip-that-detects-metastatic-cancer-cells-in-blood-one-blood-sample-can-be-tested-for-a-comprehensive-array-of-cancer-cell-markers/

2.Martin TA, Ye L, Sanders AJ, et al. Cancer Invasion and Metastasis: Molecular and Cellular Perspective. In: Madame Curie Bioscience Database [Internet]. Austin (TX): Landes Bioscience; 2000-2013.

https://www.ncbi.nlm.nih.gov/books/NBK164700/

3.F Khosravi, B King, S Rai, G Kloecker, E Wickstrom, B Panchapakesan. Nanotube devices for digital profiling of cancer biomarkers and circulating tumor cells. 23 Dec 2013. IEEE Nanotechnology Magazine 7 (4), 20-26

Nanotube devices for digital profiling of cancer biomarkers and circulating tumor cells

4.Farhad Khosravi, Patrick J Trainor, Christopher Lambert, Goetz Kloecker, Eric Wickstrom, Shesh N Rai and Balaji Panchapakesan. Static micro-array isolation, dynamic time series classification, capture and enumeration of spiked breast cancer cells in blood: the nanotube–CTC chip. 29 Sept 2016. Nanotechnology. Vol 27, No.44. IOP Publishing Ltd

http://iopscience.iop.org/article/10.1088/0957-4484/27/44/44LT03/meta

5.Seyfried, T. N., & Huysentruyt, L. C. (2013). On the Origin of Cancer Metastasis. Critical Reviews in Oncogenesis18(1-2), 43–73.

https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3597235/

6.Deeb, A., Haque, S.-U., & Olowoure, O. (2015). Pulmonary metastases in pancreatic cancer, is there a survival influence? Journal of Gastrointestinal Oncology6(3), E48–E51. http://doi.org/10.3978/j.issn.2078-6891.2014.114

https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4397254/

Other related articles published in this Open Access Online Scientific Journal include the following:

 

Liquid Biopsy Chip detects an array of metastatic cancer cell markers in blood – R&D @Worcester Polytechnic Institute, Micro and Nanotechnology Lab

Reporters: Tilda Barliya, PhD and Aviva Lev-Ari, PhD, RN

https://pharmaceuticalintelligence.com/2016/12/28/liquid-biopsy-chip-detects-an-array-of-metastatic-cancer-cell-markers-in-blood-rd-worcester-polytechnic-institute-micro-and-nanotechnology-lab/

 

Trovagene’s ctDNA Liquid Biopsy urine and blood tests to be used in Monitoring and Early Detection of Pancreatic Cancer

Reporters: David Orchard-Webb, PhD and Aviva Lev-Ari, PhD, RN

https://pharmaceuticalintelligence.com/2016/07/06/trovagenes-ctdna-liquide-biopsy-urine-and-blood-tests-to-be-used-in-monitoring-and-early-detection-of-pancreatic-cancer/

 

Liquid Biopsy Assay May Predict Drug Resistance

Curator: Larry H. Bernstein, MD, FCAP

https://pharmaceuticalintelligence.com/2015/11/06/liquid-biopsy-assay-may-predict-drug-resistance/


New insights in cancer, cancer immunogenesis and circulating cancer cells

Larry H. Bernstein, MD, FCAP, Curator

https://pharmaceuticalintelligence.com/2016/04/15/new-insights-in-cancer-cancer-immunogenesis-and-circulating-cancer-cells/

 

Prognostic biomarker for NSCLC and Cancer Metastasis

Larry H. Bernstein, MD, FCAP, Curato

https://pharmaceuticalintelligence.com/2016/03/24/prognostic-biomarker-for-nsclc-and-cancer-metastasis/

 

Monitoring AML with “cell specific” blood test

Larry H. Bernstein, MD, FCAP, Curator

https://pharmaceuticalintelligence.com/2016/01/23/monitoring-aml-with-cell-specific-blood-test/

 

Diagnostic Revelations

Larry H. Bernstein, MD, FCAP, Curator

https://pharmaceuticalintelligence.com/2015/11/02/diagnostic-revelations/

 

Circulating Biomarkers World Congress, March 23-24, 2015, Boston: Exosomes, Microvesicles, Circulating DNA, Circulating RNA, Circulating Tumor Cells, Sample Preparation

Reporter: Aviva Lev-Ari, PhD, RN

https://pharmaceuticalintelligence.com/2015/03/03/circulating-biomarkers-world-congress-march-23-24-2015-boston-exosomes-microvesicles-circulating-dna-circulating-rna-circulating-tumor-cells-sample-preparation/

 

 

 

Read Full Post »

Li -Fraumeni Syndrome and Pancreatic Cancer

Curator: Marzan Khan, B.Sc.

Li-Fraumeni syndrome (LFS) is a condition that makes individuals prone to developing a wide variety of cancers that occur early on in life, the most common types being- soft tissue sarcoma, osteosarcoma, breast cancer, brain tumors, adrenocortical carcinoma (ACC), and leukemia. (1) Pancreatic cancer is minimally associated with the condition. (2) A survey found the presence of pancreatic cancer in only 1% of 475 tumor samples collected from 91 families who were carriers of p53 mutations, with half of them having LFS. The incidence of breast cancer amongst them was the highest -24%. (2) Pancreatic carcinoma in LFS patients usually occurs in the later stages of life. (3)

The underlying cause of LFS is germline mutations in TP53 gene on chromosome 17p, that encodes the transcription factor p53, crucial in cell cycle regulation and the repair of damaged and/or abnormal cells. (4) In the majority of cases, this mutation is obtained by inheritance. (5) De-novo germline mutations in p53 occur in 7%-20% of the cases. (5)

A person showing symptoms of any type of cancer at an early age or having first or second-degree relatives with cancer are at risk of developing LFS. (5) That is why tracing family history is an important part of diagnosis in LFS patients. Genetic testing can confirm mutations present in the gene, however, there are controversial ethical issues regarding their use, particularly in children and fetuses.

In patients with LFS, it is important to control the manifestations of the disease. They should be monitored closely so that any new cancers that arise are diagnosed and treated during the early stages. (6) Patients are also at risk of developing radiation-induced second and third primary tumors. (6) Therefore, radiation and alkylating agents should be used minimally (6) People at risk can be cautioned to avoid exposure to carcinogens such as sunlight, cigarette smoke, and alcohol consumption. (5) Therapeutic approaches that are aimed at restoring wild-type p53 by gene therapy as well as reactivating non-functional p53 by the use of small-molecule drugs are currently being investigated in many cancers. (7) Unlike radiation therapy, these small-molecule drugs are non-toxic to healthy cells, thus eliminating the risk of forming new tumors.

So far, PRIMA-1 has proven to be quite effective at correcting non-functional p53. (8) PRIMA-1 is changed to its methylated form, PRIMA-1MET   that forms covalent adducts to thiol groups in the mutated protein and modifies them. (8) As a result, p53 regains its ability to destroy malignant cells. (8) A research study also found that PRIMA-1 induces apoptosis and increases the sensitivity of pancreatic cancer cells to various chemotherapeutic agents. (9)

  1. Magali Olivier, David E. Goldgar, Nayanta Sodha, Hiroko Ohgaki, Paul Kleihues, Pierre Hainaut and Rosalind A. Eeles. Li-Fraumeni and Related Syndromes. Cancer Res October 15 2003 63 (20) 6643-6650 http://cancerres.aacrjournals.org/content/63/20/6643.abstract
  2. Kleihues P, Schauble B, zur Hausen H, et al. Tumors associated with p53 germline mutations: a synopsis of 91 families. Am J Pathol 1997; 150:1-13 https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1858532/
  3. John P. Neoptolemos, Raul Urrutia, James L. Abbruzzese, Markus W. Buchler. Pancreatic Cancer. 2010.1st ed, pp-6, 2010, Springer, Verlag, New York
  4. Mishra B and Patel RR. Gene Therapy for Treatment of Pancreatic Cancer. Austin Therapeutics. 2014;1(1): 10. https://books.google.ca/books?id=NmBB5ZoKkk4C&pg=PA6&lpg=PA6&dq=connection+between+li+fraumeni+and+Pancreatic+cancer&source=bl&ots=H0iCeaPP0N&sig=pqJT1tPMR6C-NIig3S_NkFKFsD0&hl=en&sa=X&ved=0ahUKEwi4nLrgzuPQAhUUIWMKHS3wBoc4ChDoAQhNMAg#v=onepage&q=connection%20between%20li%20fraumeni%20and%20Pancreatic%20cancer&f=false
  5. Schneider K, Zelley K, Nichols KE, et al. Li-Fraumeni Syndrome. 1999 Jan 19 [Updated 2013 Apr 11]. In: Pagon RA, Adam MP, Ardinger HH, et al., editors. GeneReviews® [Internet]. Seattle (WA): University of Washington, Seattle; 1993-2016. https://www.ncbi.nlm.nih.gov/pubmed/20301488
  6. Elisa Becze BA, ELS, 2011 Mar 1. An introduction to Li-Fraumeni Syndrome, Five-Minute-In-Service. http://connect.ons.org/columns/five-minute-in-service/an-introduction-to-li-fraumeni-syndrome
  7. Sorrell, A. D., Espenschied, C. R., Culver, J. O., & Weitzel, J. N. (2013).TP53Testing and Li-Fraumeni Syndrome: Current Status of Clinical Applications and Future Directions. Molecular Diagnosis & Therapy17(1), 31–47. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3627545/
  8. Emily J. Lewis. PRIMA-1 as a cancer therapy restoring mutant p53: a reviewBioscience Horizons (2015) 8: hzv006 http://biohorizons.oxfordjournals.org/content/8/hzv006.full
  9. Izetti, Patricia, Agnes Hautefeuille, Ana Lucia Abujamra, Caroline Brunetto de Farias, Juliana Giacomazzi, Bárbara Alemar, Guido Lenz, et al. ‘PRIMA-1, a Mutant p53 Reactivator, Induces Apoptosis and Enhances Chemotherapeutic Cytotoxicity in Pancreatic Cancer Cell Lines’. Investigational New Drugs 32, no. 5 (October 2014): 783–94. https://www.ncbi.nlm.nih.gov/pubmed/24838627

Izetti, Patricia, Agnes Hautefeuille, Ana Lucia Abujamra, Caroline Brunetto de Farias, Juliana Giacomazzi, Bárbara Alemar, Guido Lenz, et al. ‘PRIMA-1, a Mutant p53 Reactivator, Induces Apoptosis and Enhances Chemotherapeutic Cytotoxicity in Pancreatic Cancer Cell Lines’. Investigational New Drugs 32, no. 5 (October 2014): 783–94

Other related articles published in this Online Scientific Journal include the following:

p53 mutation – Li-Fraumeni Syndrome – Likelihood of Genetic or Hereditary conditions playing a role in Intergenerational incidence of Cancer

Reporter: Aviva Lev-Ari, PhD, RN

https://pharmaceuticalintelligence.com/2016/12/01/p53-mutation-li-fraumeni-syndrome-likelihood-of-genetic-or-hereditary-conditions-playing-a-role-in-intergenerational-incidence-of-cancer/

Pancreatic Cancer: Articles of Note @PharmaceuticalIntelligence.com

Curator: Aviva Lev-Ari, PhD, RN

https://pharmaceuticalintelligence.com/2016/05/26/pancreatic-cancer-articles-of-note-pharmaceuticalintelligence-com/

Read Full Post »

Oncolytic Virotherapy for Pancreatic Cancer: Overcoming Obstacles in Oncolytic Virus Delivery

Reporter: Aviva Lev-Ari, PhD, RN

 

We covered MGH’s Innovation on Tumor targeted therapy in Pancreatic Cancer in

Pancreatic Cancer Targeted Treatment?

Curator: Larry H. Bernstein, MD, FCAP

https://pharmaceuticalintelligence.com/2016/05/18/pancreatic-cancer-targeted-treatment/

 

Below, we report on the State of the Science for Overcoming Obstacles in Oncolytic Virus Delivery and provide the source for all the references used

 

ONCOLYTIC VIROTHERAPY FOR PANCREATIC CANCER

Adenovirus

ONYX-015 was the first TOV used in a clinical trial for pancreatic cancer. ONYX-015 was administered intratumourally under endoscopic ultrasound-guidance into patients with locally advanced adenocarcinoma of the pancreas or metastatic disease in phase I/II trials[132]. The treatment was well-tolerated in most patients, however no objective responses were seen with ONYX-015 as a single agent and only 2/21 patients experienced mild responses when combined with gemcitabine[132]. A second adenovirus vector carries a deletion in the E1A gene[133]. E1A normally binds to the retinoblastoma protein, forcing cells to prematurely enter the S phase of the cell cycle. Since most pancreatic cancers harbor a mutation in CDKN2A[134], the E1A protein is unnecessary for entry of the TOV into cancer cells. Furthermore a double-deleted (E1A and E1B19) adenovirus demonstrated increase potency and selectivity in pancreatic cancer models[135,136]. This demonstrates that TOVs can be genetically engineered to increase selectivity and efficacy while maintaining their potency. Adenovirus selectivity has also been improved by engineering tumour-specific promoters such as a human CEA promoter[137] or by substituting the adenovirus serotype 5 fiber knob with the fiber knob from serotype 3[138]. The potency of TOVs can also be improved further by engineering them with therapeutic genes that stimulate the immune system and/or improve direct oncolysis. Adenovirus ZD55-IL-24 expressing IL-24 locally in pancreatic tumours in immune competent mice inhibited tumour growth and induced a stronger T cell response compared to its backbone virus, as measured by IL-6 and IFN-γ levels[139].

HSV

Two oncolytic HSV-1 vectors are currently in clinical trials for the treatment of pancreatic cancer. HF10 is a non-engineered, naturally occurring oncolytic HSV that demonstrated regression in 1/6 of the patients treated[140,141]. OncoVex GM-CSF is a ∆34.5 and ICP47-deleted mutant expressing GM-CSF, whereby the deletions allow for tumour-selective replication and inhibition of protein-kinase R activation, respectively[142]. Phase I/II trials in various solid tumours demonstrated OncoVex GM-CSF to be well-tolerated at high and repeated doses[143,144]. A phase I clinical trial with OncoVex GM-CSF in patients with unresectable pancreatic cancer is underway.

Poxviruses

The most widely studied poxvirus is VV, which is highly immunogenic and produces a strong cytotoxic T cell response[145] and circulating neutralizing antibodies which can be detected decades later[146]. For its crucial role in the eradication of smallpox, much has been learned about its potential role in immunotherapy today. The Lister strain of vaccinia remarkably showed no replication degradation even under the hypoxic conditions of PDAC[147]. A second Lister strain, thymidine kinase-deleted replicating VV armed with IL-10 demonstrated superior and long-lasting antitumour immunity in both a subcutaneous pancreatic cancer model and a Kras-p53 mutant-transgenic pancreatic cancer model after systemic delivery compared to its unarmed backbone virus[148]. Myxoma virus, a rabbit-specific poxvirus combined with gemcitabine resulted in 100% long-term survival in Pan02-engrafted immunocompetent intraperitoneal dissemination models of pancreatic cancer[149]. The only poxvirus to be tested in clinical trials is a non-replicative VV that expresses the pancreatic TAAs CEA and MUC-2[150]. The vaccine also includes a triad of costimulatory molecules, B7.1 (CD80), ICAM-1 (intra-cellular adhesion molecule-1) and LFA-3 (leukocyte function-associated antigen-3) (TRICOM) (PANVAC-VF)[150]. GM-CSF was also used as an adjuvant following each vaccination of PANVAC-VF. Phase I trials demonstrated antigen-specific antitumour responses in 62.5% of patients enrolled and antibody responses against VV was observed in all ten patients, which was associated with an increase in survival (15.1 mo vs 3.9 mo)[48]. A phase III clinical trial for the treatment of metastatic pancreatic cancer after failing treatment with gemcitabine, however, was terminated after failing to reach its primary efficacy endpoint[151].

Other pre-clinical TOVs for pancreatic cancer therapy

Parvovirus, measles virus and reovirus have also demonstrated pre-clinical activity in pancreatic cancer models. Parvoviruses particularly demonstrated enhanced IL-2-activated NK responses against PDAC cells[152,153]. An armed measles virus (MV), MV-purine nucleoside phosphorylase (PNP)-anti-prostate stem cell antigen, that expresses the prodrug convertase PNP, which then activates the prodrug fludarabine, was shown to enhance the oncolytic efficacy of the virus in gemcitabine-resistant PDAC cells[154]. Reovirus is another promising TOV for pancreatic cancer therapy, particularly because its selectivity depends on the cellular activity of Ras, which is constitutively active in pancreatic cancer[155]. Reolysin® (Oncolytics Biotech Inc., Calgary, AB, Canada) a reovirus administered intraportally resulted in decreased metastatic tumour volumes in the liver of immunocompetent animal models[156,157]. A phase II study of Reolysin® in combination with gemcitabine in patients with advanced PDAC has been completed (clinicaltrials.gov: NCT00998322). A two-armed randomized phase II study of carboplatin and paclitaxel plus Reolysin® vs carboplatin and paclitaxel alone in recurrent or metastatic pancreatic cancer is currently being conducted by the United States National Cancer Institute (NCI-8601/OSU-10045).

RATIONALIZING VIRO-IMMUNE-CHECKPOINT COMBINATION THERAPY

A understanding how antitumour immunity is regulated allows us to recognize barriers against effective immunotherapy delivery and furthermore, allow for the development of rational combination therapies aiming targeting these mechanisms[108,158,159]. This approach allows therapies to work synergistically and also has the potential to benefit a broader patient population[108]. Tumours have evolved to avoid immune recognition and/or destruction at every stage in the antitumour response, therefore targeting more than one immune resistance mechanism will enhance antitumour immunity.

An important immunological barrier in cancer immunotherapy is the tolerance towards self-antigens. Tumours downregulate their antigenicity through various mechanisms in response to selective pressure by the immune system, a process called “immunoediting”[37]. Therefore, in order to raise an effective antitumour response, the immunological tolerance must be broken to allow tumour antigen-specific cytotoxic T cell responses[158]. This can be achieved by increasing the tumour load and/or enhance antigen presentation[108]. TOVs can initiate selective infection and replication in the tumour bed, exposing TAA, disrupting the immunotolerance employed by the tumour while re-engaging adaptive immune effector responses[39]. Combining an agent that can cause disruption to the tumour bed i.e., an oncolytic virus, with a novel antitumour immunomodulating agent such as anti-PD-1/PD-L1 antibodies can maximize immune-stimulating and immune-recruiting inflammatory responses[39]. Specifically, TOV lysis induces the release of tumour antigens into the microenvironment, which are then cross-presented to T cells in the draining lymph nodes by APCs[159] (Figure (Figure1).1). This allows T cell infiltration to the tumour bed. Next, T cell dysfunction must be reversed[108,158]. Immune checkpoint inhibitors alleviate immunosuppression, allowing the elimination of the tumour by the adaptive immune system[70]. TOVs in combination with immune checkpoint inhibitors can therefore potentiate and activate the immune system synergistically, ultimately creating a pro-inflammatory environment. Pre-existing TILs are strong prognostic predictors in cancer[106]. This is extremely relevant for tumours with poor immune-cell infiltration, such as pancreatic cancer, which would depend on TOV-infection mediated lymphocyte infiltration for an enhanced response to immune checkpoint blockade. Zamarin et al[160] demonstrated constrained replication of an intratumoural-injected Newcastle disease virus in a B16 melanoma model. Lymphocytic infiltrates, however, were detected in both TOV-injected and non-TOV-injected tumours, and rendered the tumours sensitive to CTLA-4 blockade. The antitumour activity was dependent on CD8+ T cells, NK cells and type I and II IFNs[160]. Ipilimumab with or without talimogene laherparapvec, is in early clinical testing in patients with unresected melanoma (clinicaltrials.org: NCT01740297). Interestingly, an MV engineered to express CTLA-4 or PD-L1 antibodies delayed tumour progression and prolonged median OS in B16 melanoma models[161]. Finally, TOVs have demonstrated a tolerable toxicity profile, whereby flu-like symptoms are the most common adverse events, and in fact, most of the side effects seen so far in the combination regiment are related to the immune checkpoint blockade inhibitor[162]. Dias et al[163] suggested an oncolytic adenovirus expressing CTLA-4 locally might reduce systemic side effects normally induced with anti-CTLA-4 antibodies alone.

OVERCOMING OBSTACLES IN ONCOLYTIC VIRUS DELIVERY

The main issue with virotherapy is systemic delivery for targeting metastatic cancer cells. Intravenous administration is more practical, especially for treatment of a tumour in a hard-to-reach location such as the pancreas, and with the majority of patients presenting with advanced or metastatic disease. However, nonimmune human serum and existing anti-TOV antibodies may neutralize the TOV in the bloodstream. Furthermore, non-specific hepatic and splenic sequestration of the TOV and ineffective extravasation into the tumours are important issues[164]. Currently, studies in pre-clinical models aim to overcome these obstacles. These include chemical modification of viral coat proteins by conjugation of biocompatible polymers e.g. polyethylene glycosylation[165,166], using mesenchymal stem cell carrier systems to deliver the TOV to the tumour bed[167169], and increasing vessel permeabilization[170,171].

In PDAC, however, the biggest hurdle may not be the host immune system, but the TME. The TME has played a significant role in not only acting as a physical barrier to deliver treatments, but it also in the development of resistance to conventional drugs. The TME remains a problem for successful TOV treatment. The TOV must be able to spread in the hypoxic and densely stromal-rich TME in order to attract enough attention to induce antitumour immunity[172]. Breaching the stromal barrier in PDAC is needed for TOVs to access the cancer cells[173]. Paradoxically, a recent study by Ilkow et al[174] demonstrated that the cross-talk between CAFs and cancer cells actually lead to increased permissibility of TOV-based therapeutics. Tumour cells producing TGF-α reprogrammed CAFs, dampening levels of anti-viral transcripts. This allowed the cells to be more sensitive to VV, vesicular stomatitis virus and maraba MG1 TOVs. The reprogrammed CAFs produced fibroblast growth factor (FGF)-2 which suppressed levels of retinoic acid-inducible gene I and increased the susceptibility of the tumour cells to virus[175]. This study also demonstrated that an FGF2-expressing TOV has improved therapeutic efficacy by sensitizing the tumour cells to virotherapy and is particularly relevant to pancreatic cancers, where CAFs are a major component of the tumour stroma[175]. It is important to note that not only the patient’s existing immune system may impede successful TOV therapy, but that the enhanced antitumour response by combinatory approaches (e.g., the inclusion of immune-checkpoint inhibitors) may also impede successful TOV infection, spread and engagement of the immune system. This stresses the importance of determining strategic combinations, dosing and timing schedules in future studies.

CONCLUSION

The poor prognosis of pancreatic cancer due in part to the limited efficacy of conventional and targeted therapies, appeals for a novel strategy to treat this disease. It has become very clear that the immune system has the greatest potential to selectively destroy tumours, and when it is strategically induced, a durable benefit can be achieved. Past and present studies have defined means for tumour escape from immune surveillance and have developed immunotherapies to counteract these mechanisms. However, with the various escape strategies leading to low immunogenicity and highly immunosuppressive tumour beds, a successful control of tumour growth by immunotherapy does not come without various obstacles and challenges. Future steps include the development of immune-monitoring strategies for the identification of biomarkers, to establishment guidelines to assess clinical end points of immunotherapy and finally to evaluate combination therapeutic strategies to maximize clinical benefit[176]. The ability of TOVs to stimulate inflammation, deliver genes and immunomodulatory agents as well as reduce tumour burden by direct cell lysis, allows them to be important therapeutic vectors for a highly immunosuppressed tumour such as PDAC. Immune checkpoint blockade agents can then reverse T cell anergy and further boost OV-induced responses. As this combinatory approach may exist as a double-edged sword, it is crucial to determine appropriate timing, dosing and sequence schedules of each agent.

SOURCE & REFERENCES

Read Full Post »

Recent Research On SMAD4 In Pancreatic Cancer

Curator: David Orchard-Webb, PhD

 

Deleted in Pancreatic Cancer, locus 4 (DPC4) officially known as SMAD4 is a component of the Transforming Growth Factor Beta (TGFß) pathway with tumour suppressive properties. As its name suggests it is frequently lost in pancreatic cancer, although through a variety of mechanisms in addition to gene deletion. The loss of SMAD4 is important in the progression of pancreatic intraepithelial neoplasia (PanIN) towards pancreatic ductal adenocarcinoma (PDAC). The expression of SMAD4 can suppress metastasis, angiogenesis, and cancer stem-like cell generation. SMAD4 can promote cancer cell apoptosis through a recently described mechanism involving a lethal epithelial to mesenchymal transition (EMT). SMAD4 status has a predictive role in pancreatic cancer personalised medicine. This curation categorises recent publications of note regarding SMAD4.

 

Role of SMAD4 in neoplastic progression towards PDAC

 

Garcia-Carracedo, Dario, Chih-Chieh Yu, Nathan Akhavan, Stuart A. Fine, Frank Schönleben, Naoki Maehara, Dillon C. Karg, et al. ‘Smad4 Loss Synergizes with TGFα Overexpression in Promoting Pancreatic Metaplasia, PanIN Development, and Fibrosis’. Edited by Ilse Rooman. PLOS ONE 10, no. 3 (24 March 2015): e0120851. doi:10.1371/journal.pone.0120851.

 

Norris, A M, A Gore, A Balboni, A Young, D S Longnecker, and M Korc. ‘AGR2 Is a SMAD4-Suppressible Gene That Modulates MUC1 Levels and Promotes the Initiation and Progression of Pancreatic Intraepithelial Neoplasia’. Oncogene 32, no. 33 (15 August 2013): 3867–76. doi:10.1038/onc.2012.394.

 

Leung, Lisa, Nikolina Radulovich, Chang-Qi Zhu, Dennis Wang, Christine To, Emin Ibrahimov, and Ming-Sound Tsao. ‘Loss of Canonical Smad4 Signaling Promotes KRAS Driven Malignant Transformation of Human Pancreatic Duct Epithelial Cells and Metastasis’. Edited by Hidayatullah G Munshi. PLoS ONE 8, no. 12 (27 December 2013): e84366. doi:10.1371/journal.pone.0084366.

 

Mechanism of SMAD4 deactivation

 

Xia, Xiang, Kundong Zhang, Gang Cen, Tao Jiang, Jun Cao, Kejian Huang, Chen Huang, Qian Zhao, and Zhengjun Qiu. ‘MicroRNA-301a-3p Promotes Pancreatic Cancer Progression via Negative Regulation of SMAD4’. Oncotarget 6, no. 25 (28 August 2015): 21046–63. doi:10.18632/oncotarget.4124.

 

Murphy, Stephen J., Steven N. Hart, Geoffrey C. Halling, Sarah H. Johnson, James B. Smadbeck, Travis Drucker, Joema Felipe Lima, et al. ‘Integrated Genomic Analysis of Pancreatic Ductal Adenocarcinomas Reveals Genomic Rearrangement Events as Significant Drivers of Disease’. Cancer Research 76, no. 3 (1 February 2016): 749–61. doi:10.1158/0008-5472.CAN-15-2198.

 

Sawai, Yugo, Yuzo Kodama, Takahiro Shimizu, Yuji Ota, Takahisa Maruno, Yuji Eso, Akira Kurita, et al. ‘Activation-Induced Cytidine Deaminase Contributes to Pancreatic Tumorigenesis by Inducing Tumor-Related Gene Mutations’. Cancer Research 75, no. 16 (15 August 2015): 3292–3301. doi:10.1158/0008-5472.CAN-14-3028.

 

Demagny, Hadrien, and Edward M De Robertis. ‘Point Mutations in the Tumor Suppressor Smad4/DPC4 Enhance Its Phosphorylation by GSK3 and Reversibly Inactivate TGF-β Signaling’. Molecular & Cellular Oncology 3, no. 1 (2 January 2016): e1025181. doi:10.1080/23723556.2015.1025181.

 

Foster, David. ‘BxPC3 Pancreatic Cancer Cells Express a Truncated Smad4 Protein upon PI3K and mTOR Inhibition’. Oncology Letters, 28 January 2014. doi:10.3892/ol.2014.1833.

 

Hao, Jun, Shuyu Zhang, Yingqi Zhou, Cong Liu, Xiangui Hu, and Chenghao Shao. ‘MicroRNA 421 Suppresses DPC4/Smad4 in Pancreatic Cancer’. Biochemical and Biophysical Research Communications 406, no. 4 (25 March 2011): 552–57. doi:10.1016/j.bbrc.2011.02.086.

 

SMAD4 effects on cell motility

 

Zhang, Xueying, Junxia Cao, Yujun Pei, Jiyan Zhang, and Qingyang Wang. ‘Smad4 Inhibits Cell Migration via Suppression of JNK Activity in Human Pancreatic Carcinoma PANC‑1 Cells’. Oncology Letters, 7 April 2016. doi:10.3892/ol.2016.4427.

 

Kang, Ya ’an, Jianhua Ling, Rei Suzuki, David Roife, Xavier Chopin-Laly, Mark J. Truty, Deyali Chatterjee, et al. ‘SMAD4 Regulates Cell Motility through Transcription of N-Cadherin in Human Pancreatic Ductal Epithelium’. Edited by Neil A. Hotchin. PLoS ONE 9, no. 9 (29 September 2014): e107948. doi:10.1371/journal.pone.0107948.

 

Chen, Yu-Wen, Pi-Jung Hsiao, Ching-Chieh Weng, Kung-Kai Kuo, Tzu-Lei Kuo, Deng-Chyang Wu, Wen-Chun Hung, and Kuang-Hung Cheng. ‘SMAD4 Loss Triggers the Phenotypic Changes of Pancreatic Ductal Adenocarcinoma Cells’. BMC Cancer 14, no. 1 (2014): 1. https://bmccancer.biomedcentral.com/articles/10.1186/1471-2407-14-181.

 

SMAD4 effects on angiogenesis

 

Zhou, Zhichao, Juming Lu, Jingtao Dou, Zhaohui Lv, Xi Qin, and Jing Lin. ‘FHL1 and Smad4 Synergistically Inhibit Vascular Endothelial Growth Factor Expression’. Molecular Medicine Reports 7, no. 2 (February 2013): 649–53. doi:10.3892/mmr.2012.1202.

 

SMAD4 mediated repression of cancer stem-like cells

 

Hoshino, Yukari, Jun Nishida, Yoko Katsuno, Daizo Koinuma, Taku Aoki, Norihiro Kokudo, Kohei Miyazono, and Shogo Ehata. ‘Smad4 Decreases the Population of Pancreatic Cancer–Initiating Cells through Transcriptional Repression of ALDH1A1’. The American Journal of Pathology 185, no. 5 (2015): 1457–1470. http://www.sciencedirect.com/science/article/pii/S0002944015000802.

 

SMAD4 mediated growth inhibition/ apoptosis induction

 

David, Charles J., Yun-Han Huang, Mo Chen, Jie Su, Yilong Zou, Nabeel Bardeesy, Christine A. Iacobuzio-Donahue, and Joan Massagué. ‘TGF-β Tumor Suppression through a Lethal EMT’. Cell 164, no. 5 (February 2016): 1015–30. doi:10.1016/j.cell.2016.01.009.

 

Wang, Qi, Juanjuan Li, Wei Wu, Ruizhe Shen, He Jiang, Yuting Qian, Yanping Tang, et al. ‘Smad4-Dependent Suppressor Pituitary Homeobox 2 Promotes PPP2R2A-Mediated Inhibition of Akt Pathway in Pancreatic Cancer’. Oncotarget 7, no. 10 (8 March 2016): 11208–22. doi:10.18632/oncotarget.7158.

 

Poorly characterised targets of SMAD4

 

Fullerton, Paul T., Chad J. Creighton, and Martin M. Matzuk. ‘Insights Into SMAD4 Loss in Pancreatic Cancer From Inducible Restoration of TGF-β Signaling’. Molecular Endocrinology (Baltimore, Md.) 29, no. 10 (October 2015): 1440–53. doi:10.1210/me.2015-1102.

 

Li, Lei, Zhaoshen Li, Xiangyu Kong, Dacheng Xie, Zhiliang Jia, Weihua Jiang, Jiujie Cui, et al. ‘Down-Regulation of MicroRNA-494 via Loss of SMAD4 Increases FOXM1 and β-Catenin Signaling in Pancreatic Ductal Adenocarcinoma Cells’. Gastroenterology 147, no. 2 (August 2014): 485–497.e18. doi:10.1053/j.gastro.2014.04.048.

 

Drugs that restore SMAD4

 

Lin, Sheng-Zhang, Jin-Bo Xu, Xu Ji, Hui Chen, Hong-Tao Xu, Ping Hu, Liang Chen, et al. ‘Emodin Inhibits Angiogenesis in Pancreatic Cancer by Regulating the Transforming Growth Factor-Β/drosophila Mothers against Decapentaplegic Pathway and Angiogenesis-Associated microRNAs’. Molecular Medicine Reports 12, no. 4 (October 2015): 5865–71. doi:10.3892/mmr.2015.4158.

 

Predictive value of SMAD4 status in personalised medicine

 

Whittle, Martin C., Kamel Izeradjene, P. Geetha Rani, Libing Feng, Markus A. Carlson, Kathleen E. DelGiorno, Laura D. Wood, et al. ‘RUNX3 Controls a Metastatic Switch in Pancreatic Ductal Adenocarcinoma’. Cell 161, no. 6 (June 2015): 1345–60. doi:10.1016/j.cell.2015.04.048.

 

Boone, Brian A., Shirin Sabbaghian, Mazen Zenati, J. Wallis Marsh, A. James Moser, Amer H. Zureikat, Aatur D. Singhi, Herbert J. Zeh, and Alyssa M. Krasinskas. ‘Loss of SMAD4 Staining in Pre-Operative Cell Blocks Is Associated with Distant Metastases Following Pancreaticoduodenectomy with Venous Resection for Pancreatic Cancer’. Journal of Surgical Oncology 110, no. 2 (August 2014): 171–75. doi:10.1002/jso.23606.

 

Herman, Joseph M., Katherine Y. Fan, Aaron T. Wild, Laura D. Wood, Amanda L. Blackford, Ross C. Donehower, Manuel Hidalgo, et al. ‘Correlation of Smad4 Status With Outcomes in Patients Receiving Erlotinib Combined With Adjuvant Chemoradiation and Chemotherapy After Resection for Pancreatic Adenocarcinoma’. International Journal of Radiation Oncology*Biology*Physics 87, no. 3 (November 2013): 458–59. doi:10.1016/j.ijrobp.2013.06.2039.

 

Other Related Articles Published In This Open Access Online Journal Include The Following:

 

https://pharmaceuticalintelligence.com/2016/06/10/pancreatic-cancer-modeling-using-retrograde-viral-vector-delivery-and-in-vivo-crisprcas9-mediated-somatic-genome-editing/

https://pharmaceuticalintelligence.com/2015/04/10/wnt%CE%B2-catenin-signaling-7-10/

 

Read Full Post »

Pancreatic Cancer Targeted Treatment?

Curator: Larry H. Bernstein, MD, FCAP

 

 

MGH study identifies potential treatment target for pancreatic cancer

Molecular signature found in 30 percent of PDAC tumors, associated with more aggressive cancer

http://www.massgeneral.org/about/pressrelease.aspx?id=1933&

 

Massachusetts General Hospital (MGH) investigators have identified the first potential molecular treatment target for the most common form of pancreatic cancer, which kills more than 90 percent of patients. Along with finding that the tumor suppressor protein SIRT6 is inactive in around 30 percent of cases of pancreatic ductal adenocarcinoma (PDAC), the team identified the precise pathway by which SIRT6 suppresses PDAC development, a mechanism different from the way it suppresses colorectal cancer. The paper will appear in the June 2 issue of Cell and have been published online.

“With the advance of cancer genomics, it has become evident that alterations in epigenetic factors – those that control whether and when other genes are expressed – represent some of the most frequent alterations in cancer,” says Raul Mostoslavsky, MD, PhD, of the MGH Cancer Center, senior author of the report.  “Yet, not many of those factors have been described before, and those that have been identified have not been linked to specific downstream targets.  Not only did more than a third of analyzed PDAC patient samples exhibit the molecular signature we identified, those patients also turned out to have very poor prognoses.”

Among its other functions, SIRT6 is known to control how cells process glucose, and a 2012 study by Mostoslavsky’s team found that its ability to suppress colorectal cancer involves control of a process called glycolysis.  But while that study also found reduced SIRT6 expression in PDAC tumor cells, the current investigation indicated that SIRT6 deficiency promotes PDAC through a different mechanism. Experiments in cell lines and animal models revealed that low SIRT6 levels in PDAC were correlated with increased expression of Lin28b, an oncoprotein normally expressed during fetal development.

Lin28b expression proved to be essential to the growth and survival of SIRT6-deficient PDAC cells and acted by preventing a family of tumor-suppressing mRNAs called let-7 from blocking expression of three genes previously associated with increased aggressiveness and metastasis in pancreatic cancers.  All of these hallmarks – reduced SIRT6, increased Lin28b and reduced let-7 expression – were found in tumor samples from patients who died more quickly.

“A general message from these studies is that cancer cells benefit from modulating epigenetic factors like SIRT6 by acquiring the ability to override normal cellular growth control patterns,” says Mostoslavsky, an associate professor of Medicine at Harvard Medical School and an associate member at the Broad Institute.  “Each tumor type may acquire a unique set of capabilities that may provide tumor-specific growth and survival advantages, which may need to be determined for each kind of cancer.  In terms of our findings regarding PDAC, we are intrigued by the downstream pathways controlled by Lin28b and how they increase aggressiveness and metastasis, and we are hopeful that developing in the future Lin28b inhibitors could benefit this subset of PDAC patients, who currently have very few treatment options.”

 

SIRT6 Suppresses Pancreatic Cancer through Control of Lin28b

Sita Kugel, Carlos Sebastián, Julien Fitamant,…., Alon Goren, Vikram Deshpande, Nabeel Bardeesy, Raul Mostoslavsky

Figure thumbnail fx1
Highlights
  • Loss of SIRT6 cooperates with oncogenic Kras to drive pancreatic cancer
  • SIRT6 regulates the oncofetal protein Lin28b through promoter histone deacetylation
  • Lin28b drives the growth and survival of SIRT6-deficient pancreatic cancer
  • SIRT6 and Lin28b expression define prognosis in specific pancreatic cancer subsets

Chromatin remodeling proteins are frequently dysregulated in human cancer, yet little is known about how they control tumorigenesis. Here, we uncover an epigenetic program mediated by the NAD+-dependent histone deacetylase Sirtuin 6 (SIRT6) that is critical for suppression of pancreatic ductal adenocarcinoma (PDAC), one of the most lethal malignancies. SIRT6 inactivation accelerates PDAC progression and metastasis via upregulation of Lin28b, a negative regulator of the let-7 microRNA. SIRT6 loss results in histone hyperacetylation at theLin28b promoter, Myc recruitment, and pronounced induction of Lin28b and downstream let-7 target genes, HMGA2, IGF2BP1, and IGF2BP3. This epigenetic program defines a distinct subset with a poor prognosis, representing 30%–40% of human PDAC, characterized by reduced SIRT6 expression and an exquisite dependence on Lin28b for tumor growth. Thus, we identify SIRT6 as an important PDAC tumor suppressor and uncover the Lin28b pathway as a potential therapeutic target in a molecularly defined PDAC subset.

 

The multifaceted functions of sirtuins in cancer

Angeliki Chalkiadaki & Leonard Guarente  Affiliations  Corresponding author
Nature Reviews Cancer (2015); 15:608–624     http://dx.doi.org:/10.1038/nrc3985

The sirtuins (SIRTs; of which there are seven in mammals) are NAD+-dependent enzymes that regulate a large number of cellular pathways and forestall the progression of ageing and age-associated diseases. In recent years, the role of sirtuins in cancer biology has become increasingly apparent, and growing evidence demonstrates that sirtuins regulate many processes that go awry in cancer cells, such as cellular metabolism, the regulation of chromatin structure and the maintenance of genomic stability. In this article, we review recent advances in our understanding of how sirtuins affect cancer metabolism, DNA repair and the tumour microenvironment and how activating or inhibiting sirtuins may be important in preventing or treating cancer.

 

Figure 1: Overview of the role of sirtuins in the regulation of cancer metabolism

http://www.nature.com/nrc/journal/v15/n10/carousel/nrc3985-f1.jpg

The inhibitory effects of sirtuin 3 (SIRT3), SIRT4 and SIRT6 on metabolic pathways that drive cancer cells are depicted. In normal cells, SIRT6 functions as a co-repressor for the transcription factors hypoxia-inducible factor 1α (HIF1α…

 

Tumor suppressor p53 cooperates with SIRT6 to regulate gluconeogenesis by promoting FoxO1 nuclear exclusion

Ping Zhanga,1, Bo Tua,1, Hua Wangb , Ziyang Caoa , Ming Tanga , … , Bin Gaob , Robert G. Roederd,2, and Wei-Guo Zhua,e,2
PNAS | July 22, 2014;111(29): 10684–10689 |   http://www.pnas.org/content/111/29/10684.full.pdf  http://www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1411026111/-/DCSupplemental.

In mammalian cells, tumor suppressor p53 plays critical roles in the regulation of glucose metabolism, including glycolysis and oxidative phosphorylation, but whether and how p53 also regulates gluconeogenesis is less clear. Here, we report that p53 efficiently down-regulates the expression of phosphoenolpyruvate carboxykinase (PCK1) and glucose-6-phosphatase (G6PC), which encode rate-limiting enzymes in gluconeogenesis. Cell-based assays demonstrate the p53-dependent nuclear exclusion of forkhead box protein O1 (FoxO1), a key transcription factor that mediates activation of PCK1 and G6PC, with consequent alleviation of FoxO1- dependent gluconeogenesis. Further mechanistic studies show that p53 directly activates expression of the NAD+-dependent histone deacetylase sirtuin 6 (SIRT6), whose interaction with FoxO1 leads to FoxO1 deacetylation and export to the cytoplasm. In support of these observations, p53-mediated FoxO1 nuclear exclusion, down-regulation of PCK1 and G6PC expression, and regulation of glucose levels were confirmed in C57BL/J6 mice and in liver-specific Sirt6 conditional knockout mice. Our results provide insights into mechanisms of metabolism-related p53 functions that may be relevant to tumor suppression.

As the “guardian of the genome,” tumor suppressor p53 has been reported to coordinate diverse cellular responses to a broad range of environment stresses (1) and to play antineoplastic roles by activating downstream target genes involved in DNA damage repair, apoptosis, and cell-cycle arrest (2). Recent studies have indicated broader roles for p53 in mediating metabolic changes in cells under various physiological and pathological conditions (3–7). For example, p53 was reported to influence the balance between glycolysis and oxidative phosphorylation by inducing the p53-induced glycolysis and apoptosis regulator (TIGAR) and by regulating the synthesis of cytochrome c oxidase 2 (SCO2) (3), respectively, thus promoting the switch from glycolysis to oxidative phosphorylation. p53 also may impede metabolism by reducing glucose import (4) or by inhibiting the pentose phosphate pathway (PPP) (5). More recently, context-dependent inhibitory (6) or stimulatory (7, 8) effects of p53 on gluconeogenesis have been reported. It thus is clear that p53 plays important roles in glucose regulation in mammalian cells. Glucose homeostasis is maintained by a delicate balance between intestinal absorption of sugar, gluconeogenesis, and the utilization of glucose by the peripheral tissues, irrespective of feeding or fasting (9). The gluconeogenesis pathway is catalyzed by several key enzymes that include the first and last rate-limiting enzymes of the process, phosphoenolpyruvate carboxykinase (PCK1) and glucose-6-phosphatase (G6PC), respectively. The expression of both PCK1 and G6PC is controlled mainly at the transcription level. For example, the transcription factor forkhead box protein O1 (FoxO1) activates gluconeogenesis through direct binding to the promoters of G6PC and PCK1 (10). A dominant negative FoxO1 lacking its transactivation domain significantly decreases gluconeogenesis (11) whereas FoxO1 ablation impairs fasting- and cAMP-induced PCK1 and G6PC expression (12). Therefore, factors influencing expression of FoxO1 or its binding activity to the PCK1 and G6PC promoters are potential targets for gluconeogenesis regulation. The transcription activity of FoxO family members is regulated by a sophisticated signaling network. Various environmental stimuli cause different posttranslational modifications of FoxO proteins, including phosphorylation, acetylation, ubiquitination, and methylation (13–15). The phosphorylation of FoxO proteins is known to be essential for their shuttling between the nucleus and cytoplasm. For example, kinase Akt/PKB phosphorylates FoxO1 at threonine 24 and at serines 256 and 319, which in turn leads to 14-3-3 binding and subsequent cytoplasmic sequestration. The acetylation of FoxO proteins also affects their trafficking and DNA-binding activities (15–17). Sirtuin (SIRT)1, a homolog of the yeast silent information regulator-2 (Sir2), has been identified as a deacetylase for FoxO proteins (15, 17, 18). Of the seven mammalian sirtuins, SIRT1, SIRT6, and SIRT7 are localized to the nucleus (19), and SIRT6 was recently reported to act as a central player in regulating the DNA damage response, glucose metabolism, and aging (20–26). Using a knockout mouse model, it was found that SIRT6 functions as a corepressor of the transcription factor Hif1α to suppress glucose uptake and glycolysis.

Significance: Beyond its canonical functions in processes such as cell-cycle arrest, apoptosis, and senescence, the tumor suppressor p53 has been increasingly implicated in metabolism. Here, in vitro and in vivo studies establish a role for p53 in gluconeogenesis through a previously unidentified mechanism involving (i) direct activation of the gene encoding the NAD-dependent deacetylase sirtuin 6 (SIRT6), (ii) SIRT6-dependent deacetylation and nuclear exclusion of forkhead box protein O1 (FoxO1), and (iii) downregulation of FoxO1-activated genes (G6PC and PCK1) that are rate-limiting for gluconeogenesis. These results have implications for proposed tumor-suppressor functions of p53 through regulation of metabolic pathways.

Among a variety of other functions, SIRT6 was previously connected to glucose metabolism. For example, SIRT6 acts as a corepressor of the transcription factor Hif1-α to suppress glycolysis (23). Conversely, the deletion of Sirt6 in mice results in severe hypoglycemia (33) whereas the liver-specific deletion of Sirt6 leads to increased glycolysis and triglyceride synthesis (23, 34). Our study adds further evidence that SIRT6 plays an important role in glucose metabolism by connecting p53 transcription activity and gluconeogenesis. Our data also reemphasize a previously established role for SIRT6 in regulating the acetylation state and nuclear localization of FoxO proteins, albeit in a divergent manner. Thus, the Caenorhabditis elegans SIRT6/7 homolog SIR-2.4 was implicated in DAF-16 deacetylation and consequent nuclear localization and function in stress responses (35); and the effect was reported to be indirect and to involve a stress-induced inhibition by SIR-2.4 of CBP-mediated acetylation of DAF-16 that is independent of its deacetylase activity. These results, emphasizing context-dependent SIRT6 mechanisms, contrast with the SIRT6 deacetylase activity requirement for FoxO1 nuclear exclusion in the present study and a likely direct effect of SIRT6 on FoxO1 deacetylation based on their direct interaction, the SIRT6 deacetylase activity requirement, and precedent (15, 17, 18) from direct SIRT1-mediated deacetylation of FoxO proteins.

Despite a high genetic diversity, cancer cells exhibit a common set of functional characteristics, one being the “Warburg effect”: i.e., continuous high glucose uptake and a higher rate of glycolysis than that in normal cells (36). To favor the rapid proliferation requirement for high ATP/ADP and ATP/AMP ratios, cancer cells use large amounts of glucose. p53, as one of the most important tumor suppressors, exerts its antineoplastic function through diverse pathways that include the regulation of glucose metabolism. Thus, p53 regulates glucose metabolism by activation of TIGAR (3), which lowers the intracellular concentrations of fructose-2,6-bisphosphate and decreases glycolysis. On the other hand, p53 activation causes down-regulation of several glycolysisrelated factors such as phosphoglycerate mutase (PGM) (37) and the glucose transporters (4). Expression of p53 also can limit the activity of IκBα and IκBβ, thereby restricting the activation of NFκB and dampening the expression of glycolysis-promoting genes such as GLUT3 (38). As a reverse glycolysis pathway, gluconeogenesis generates glucose from noncarbohydrate precursors and is conceivably essential for tumor cell growth. However, the current study further supports the notion (6) that p53 is also involved in a gluconeogenesis inhibition pathway, which in this case is executed by enhanced SIRT6 expression and subsequent FoxO1 nuclear exclusion. These results raise the interesting possibility that an inhibition of gluconeogenesis may contribute to the tumorsuppressive function of p53.

 

http://www.frontiersin.org/Journal/DownloadFile/1/2496/20065/1/21/fphar-03-00022_pdf

http://www.jcpjournal.org/journal/DOIx.php?id=10.15430/JCP.2013.18.3.221
Keywords : Oncogenes, Tumor suppressors, Glutamine metabolism, Cancer cells … p53 is a well-known protein which is involved in many cellular functionsincluding cell … deprivation activates p53 by regulating protein phosphatase 2A ( PP2A). …. and tumor suppressors may affect glutamine metabolism in cancercells

 

Protein controlling glucose metabolism also a tumor suppressor

Finding supports metabolic strategies to control tumor growth

http://www.massgeneral.org/about/pressrelease.aspx?id=1530    December 6, 2012

A protein known to regulate how cells process glucose also appears to be a tumor suppressor, adding to the potential that therapies directed at cellular metabolism may help suppress tumor growth.  In their report in the Dec. 7 issue of Cell, a multi-institutional research team describes finding that cells lacking the enzyme SIRT6, which controls how cells process glucose, quickly become cancerous.  They also found evidence that uncontrolled glycolysis, a stage in normal glucose metabolism, may drive tumor formation in the absence of SIRT6 and that suppressing glycolysis can halt tumor formation.

“Our study provides solid evidence that SIRT6 may function as a tumor suppressor by regulating glycolytic metabolism in cancer cells,” says Raul Mostoslavsky, MD, PhD, of the Massachusetts General Hospital (MGH) Cancer Center, senior author of the report.  “Critically, our findings indicate that, in tumors driven by low SIRT6 levels, drugs that may inhibit glycolysis – currently a hot research topic among biotechnology companies – could have therapeutic benefits.”

The hypothesis that a switch in the way cells process glucose could set off tumor formation was first proposed in the 1920s by German researcher Otto Warburg, who later received the Nobel Prize for discoveries in cellular respiration.  He observed that, while glucose metabolism is normally a two-step process involving glycolysis in the cellular cytoplasm followed by cellular respiration in the mitochondria, in cancer cells rates of glycolysis are up to 200 times higher.  Warburg’s proposition that this switch in glucose processing was a primary cause of cancer did not hold up, as subsequent research supported the role of mutations in oncogenes, which can spur tumor growth if overexpressed, and tumor suppressors, which keep cell proliferation under control.  But recent studies have suggested that alterations in cellular metabolism may be part of the process through which activated oncogenes or inactivated tumor suppressors stimulate cancer formation.

A 2010 study led by Mostoslavsky found that the absence of SIRT6 – one of a family of proteins called sirtuins that regulate many important biological pathways – appears to “flip the switch” from normal glucose processing to the excess rates of glycolysis seen in cancer cells. The current study was specifically designed to investigate whether SIRT6’s control of glucose metabolism also suppresses tumor formation.  The research team first showed that cultured skin cells from embryonic mice lacking SIRT6 proliferated rapidly and quickly formed tumors when injected into adult mice.  They also confirmed elevated glycolysis levels in both cells lacking SIRT6 and tumor cells and found that formation of tumors through SIRT6 deficiency did not appear to involve oncogene activation.

Analysis of tumor samples from patients found reduced SIRT6 expression in many – particularly in colorectal and pancreatic tumors.  Even among patients whose tumors appeared to be more aggressive, higher levels of SIRT6 expression may have delayed or, for some, prevented relapse.   In a mouse model programmed to develop numerous colon polyps, the researchers showed that lack of intestinal SIRT6 expression tripled the formation of polyps, many of which became invasive tumors.  Treating the animals with a glycolytic inhibitor significantly reduced tumor formation, even in the absence of SIRT6.

“Our results indicate that, at least in certain cancers, inhibiting glycolytic metabolism could provide a strong alternative way to halt cancer growth, possibly acting synergistically with current anti-tumor therapies,” says Mostoslavsky, an assistant professor of Medicine at Harvard Medical School.  “Cancer metabolism has only recently emerged as a hallmark of tumorigenesis, and the field is rapidly expanding.  With the current pace of research and the speed at which some basic discoveries are moving into translational studies, it is likely that drugs targeting cancer metabolism may be available to patients in the near future.”

 

THE HISTONE DEACETYLASE SIRT6 IS A NOVEL TUMOR SUPPRESSOR THAT CONTROLS CANCER METABOLISM

Reprogramming of cellular metabolism is a key event during tumorigenesis. Despite being known for decades (Warburg effect), the molecular mechanisms regulating this switch remained unexplored. Here, we identify SIRT6 as a novel tumor suppressor that regulates aerobic glycolysis in cancer cells. Importantly, loss of SIRT6 leads to tumor formation without activation of known oncogenes, while transformed SIRT6-deficient cells display increased glycolysis and tumor growth, suggesting that SIRT6 plays a role in both establishment and maintenance of cancer. Using a conditional SIRT6 allele, we show that SIRT6 deletion in vivoincreases the number, size and aggressiveness of tumors. SIRT6 also functions as a novel regulator of ribosome metabolism by co-repressing MYC transcriptional activity. Lastly, SIRT6 is selectively downregulated in several human cancers, and expression levels of SIRT6 predict prognosis and tumor-free survival rates, highlighting SIRT6 as a critical modulator of cancer metabolism. Our studies reveal SIRT6 to be a potent tumor suppressor acting to suppress cancer metabolism.

Cancer cells are characterized by the acquisition of several characteristics that enable them to become tumorigenic (Hanahan and Weinberg, 2000). Among them, the ability to sustain uncontrolled proliferation represents the most fundamental trait of cancer cells. This hyperproliferative state involves the deregulation of proliferative signaling pathways as well as loss of cell cycle regulation. In addition, tumor cells need to readjust their energy metabolism to fuel cell growth and division. This metabolic adaptation is directly regulated by many oncogenes and tumor suppressors, and is required to support the energetic and anabolic demands associated with cell growth and proliferation (Lunt and Vander Heiden, 2011).

Alteration in glucose metabolism is the best-known example of metabolic reprogramming in cancer cells. Under aerobic conditions, normal cells convert glucose to pyruvate through glycolysis, which enters the mitochondria to be further catabolized in the tricarboxylic acid cycle (TCA) to generate adenosine-5’-triphosphate (ATP). Under anaerobic conditions, mitochondrial respiration is abated; glucose metabolism is shifted towards glycolytic conversion of pyruvate into lactate. This metabolic reprogramming is also observed in cancer cells even in the presence of oxygen and was first described by Otto Warburg several decades ago (Warburg, 1956; Warburg et al., 1927). By switching their glucose metabolism towards “aerobic glycolysis”, cancer cells accumulate glycolytic intermediates that will be used as building blocks for macromolecular synthesis (Vander Heiden et al., 2009). Most cancer cells exhibit increased glucose uptake, which is due, in part, to the upregulation of glucose transporters, mainly GLUT1 (Yamamoto et al., 1990; Younes et al., 1996). Moreover, cancer cells display a high expression and activity of several glycolytic enzymes, including phospho-fructose kinase (PFK)-1, pyruvate kinase M2, lactate dehydrogenase (LDH)-A and pyruvate dehydrogenase kinase (PDK)-1 (Lunt and Vander Heiden, 2011), leading to the high rate of glucose catabolism and lactate production characteristic of these cells. Importantly, downregulation of either LDH-A or PDK1 decreases tumor growth (Bonnet et al., 2007; Fantin et al., 2006; Le et al., 2010) suggesting an important role for these proteins in the metabolic reprogramming of cancer cells.

Traditionally, cancer-associated alterations in metabolism have been considered a secondary response to cell proliferation signals. However, growing evidence has demonstrated that metabolic reprogramming of cancer cells is a primary function of activated oncogenes and inactivated tumor suppressors (Dang et al., 2012;DeBerardinis et al., 2008; Ward and Thompson, 2012). Despite this evidence, whether the metabolic reprogramming observed in cancer cells is a driving force for tumorigenesis remains as yet poorly understood.

Sirtuins are a family of NAD+-dependent protein deacetylases involved in stress resistance and metabolic homeostasis (Finkel et al., 2009). In mammals, there are seven members of this family (SIRT1-7). SIRT6 is a chromatin-bound factor that was first described as a suppressor of genomic instability (Mostoslavsky et al., 2006). SIRT6 also localizes to telomeres in human cells and controls cellular senescence and telomere structure by deacetylating histone H3 lysine 9 (H3K9) (Michishita et al., 2008). However, the main phenotype SIRT6 deficient mice display is an acute and severe metabolic abnormality. At 20 days of age, they develop a degenerative phenotype that includes complete loss of subcutaneous fat, lymphopenia, osteopenia, and acute onset of hypoglycemia, leading to death in less than ten days (Mostoslavsky et al., 2006). Recently, we have demonstrated that the lethal hypoglycemia exhibited by SIRT6 deficient mice is caused by an increased glucose uptake in muscle and brown adipose tissue (Zhong et al., 2010). Specifically, SIRT6 co-represses HIF-1α by deacetylating H3K9 at the promoters of several glycolytic genes and, consequently, SIRT6 deficient cells exhibit increased glucose uptake and upregulated glycolysis even under normoxic conditions (Zhong et al., 2010). Such a phenotype, reminiscent of the “Warburg Effect” in tumor cells, prompted us to investigate whether SIRT6 may protect against tumorigenesis by inhibiting glycolytic metabolism.

Here, we demonstrate that SIRT6 is a novel tumor suppressor that regulates aerobic glycolysis in cancer cells. Strikingly, SIRT6 acts as a first hit tumor suppressor and lack of this chromatin factor leads to tumor formation even in non-transformed cells. Notably, inhibition of glycolysis in SIRT6 deficient cells completely rescues their tumorigenic potential, suggesting that enhanced glycolysis is the driving force for tumorigenesis in these cells. Furthermore, we provide new data demonstrating that SIRT6 regulates cell proliferation by acting as a co-repressor of c-Myc, inhibiting the expression of ribosomal genes. Finally, SIRT6 expression is downregulated in human cancers, strongly reinforcing the idea that SIRT6 is a novel tumor suppressor.

…..

In addition to controlling glucose metabolism in cancer cells, our current work unravels a novel function of SIRT6 as a regulator of ribosomal gene expression. One of the main features of cancer cells is their high proliferative potential. In order to proliferate, cancer cells readjust their metabolism to generate biosynthetic precursors for macromolecular synthesis (Deberardinis et al., 2008). However, protein synthesis also requires the activation of a transcriptional program leading to ribosome biogenesis and mRNA translation (van Riggelen et al., 2010). As a master regulator of cell proliferation, MYC regulates ribosome biogenesis and protein synthesis by controlling the transcription and assembly of ribosome components as well as translation initiation (Dang et al., 2012; van Riggelen et al., 2010). Our results show that SIRT6 specifically regulates the expression of ribosomal genes. In keeping with this, SIRT6-deficient tumor cells exhibit high levels of ribosomal protein gene expression. Beyond ribosome biosynthesis, MYC regulates glucose and glutamine metabolism (Dang et al., 2012). Our results show that glutamine – but not glucose – metabolism is rescued in SIRT6-deficient/MYC knockdown cells, suggesting that SIRT6 and MYC might have redundant roles in regulating glucose metabolism.

Overall, our results indicate that SIRT6 represses tumorigenesis by inhibiting a glycolytic switch required for cancer cell proliferation. Inhibition of glycolysis in SIRT6-deficient cells abrogates tumor formation, providing proof of concept that inhibition of glycolytic metabolism in tumors with low SIRT6 levels could provide putative alternative approaches to modulate cancer growth. Furthermore, we uncover a new role for SIRT6 as a regulator of ribosome biosynthesis by co-repressing MYC transcriptional activity. Our results indicate that SIRT6 sits at a critical metabolic node, modulating both glycolytic metabolism and ribosome biosynthesis (Figure 7L). SIRT6 deficiency deregulates both pathways, leading to robust metabolic reprogramming that is sufficient to promote tumorigenesis bypassing major oncogenic signaling pathway activation.

 

Lack of cellular enzyme triggers switch in glucose processing

Understanding mechanism underlying SIRT6 activity may help treat diabetes, cancer

http://www.massgeneral.org/about/pressrelease.aspx?id=1196   January 21, 2010

A study investigating how a cellular enzyme affects blood glucose levels in mice provides clues to pathways that may be involved in processes including the regulation of longevity and the proliferation of tumor cells. In their report in the January 22 issue of Cell, a Massachusetts General Hospital (MGH)-based team of researchers describes the mechanism by which absence of the enzyme SIRT6 induces a fatal drop in blood sugar in mice by triggering a switch between two critical cellular processes.

“We found that SIRT6 functions as a master regulator of glucose levels by maintaining the normal processes by which cells convert glucose into energy,” says Raul Mostoslavsky, MD, PhD, of the MGH Cancer Center, who led the study. “Learning more about how this protein controls the way cells handle glucose could lead to new approaches to treating type 2 diabetes and even cancer.”

SIRT6 belongs to a family of proteins called sirtuins, which regulate important biological pathways in organisms from bacteria to humans. Originally discovered in yeast, sirtuins in mammals have been shown to have important roles in metabolic regulation, programmed cell death and adaptation to stress. SIRT6 is one of seven mammalian sirtuins, and Mostoslavsky’s team previously showed that mice lacking the protein die in the first month of life from acute hypoglycemia. The current study was designed to investigate exactly how lack of SIRT6 causes this radical drop in blood sugar.

Normally cells convert glucose into energy through a two-step process. The first stage called glycolysis takes place in the cytoplasm, where glucose is broken down into an acid called pyruvate and a few molecules of ATP, the enzyme that provides the energy to power most biological processes. Pyruvate is taken into cellular structures called mitochondria, where it is further processed to release much greater amounts of ATP through a process called cellular respiration.

In a series of experiments in mouse cells, the researchers showed that SIRT6-deficiency hypoglycemia is caused by increased cellular uptake of glucose and not by elevated insulin levels or defects in the absorption of glucose from food. They then found increased levels of glycolysis and reduced mitochondrial respiration in SIRT6-knockout cells, something usually seen when cells are starved for oxygen or glucose, and showed that activation of the switch from cellular respiration to glycolysis is controlled through SIRT6’s regulation of a protein called HIF1alpha. Normally, SIRT6 represses glycolytic genes through its role as a compactor of chromatin – the tightly wound combination of DNA and a protein backbone that makes up chromosomes. In the absence of SIRT6, this structure is opened, causing activation of these glycolytic genes. The investigators’ finding increased expression of glycolytic genes in living SIRT6-knockout mice – which also had elevated levels of lactic acid, characteristic of a switch to glycolytic glucose processing – supported their cellular findings.

Studies in yeast, worms and flies have suggested a role for sirtuins in aging and longevity, and while much of the enzymes’ activity in mammals is unclear, SIRT6’s control of critical glucose-metabolic pathways could signify a contribution to lifespan regulation. Elevated glycolysis also is commonly found in tumor cells, suggesting that a lack of SIRT6 could contribute to tumor growth. Conversely, since knocking out SIRT6 causes blood sugar to drop, limited SIRT6 inhibition could be a novel strategy for treating type 2 diabetes.

“There’s a lot we still don’t know about SIRT6,” adds Mostoslavsky, who is an assistant professor of Medicine at Harvard Medical School. “We need to identify the factors that interact with SIRT6 and determine how it is regulated; investigate whether it acts as a tumor suppressor and how it might help lower glucose levels in diabetes; and determine its target organs in living animals, all of which we are investigating.”

 

A tale of metabolites: the crosstalk between chromatin and energy metabolism

Mitochondrial metabolism influences histone and DNA modifications by retrograde signaling and activation of transcriptional programs. Considering the high number of putative sites for acetylation and methylation in chromatin, we propose in this Perspective that epigenetic modifications might impinge on cellular metabolism by affecting the pool of acetyl-CoA and SAM.

Metabolism can be defined as the sum of chemical reactions that occur within a cell to sustain life. It is also the way that a cell interacts with energy sources: in other words, it is the coordination of energy intake, its utilization and storage that ultimately allows growth and cell division. In animal cells, mitochondria have evolved to become the most efficient system to generate energy. This organelle consumes carbon sources via oxidative phosphorylation to produce ATP, the energy currency of the cell. Additionally, the mitochondria produces intermediate metabolites for the biosynthesis of DNA, proteins and lipids.

Under basic dividing conditions, uptake of nutrients is tightly regulated through growth signaling pathways, thus differentiated cells engage in oxidative metabolism, the most efficient mechanism to produce energy from nutrients. Cells metabolize glucose to pyruvate through glycolysis in the cytoplasm, and this pyruvate is then oxidized into CO2 through the mitochondrial TCA cycle. The electrochemical gradient generated across the inner mitochondrial membrane facilitates ATP production in a highly efficient manner. Studies in recent years indicate that under conditions of nutrient excess, cells increase their nutrient uptake, adopting instead what is known as aerobic glycolysis, an adaptation that convert pyruvate into lactate, enabling cells to produce intermediate metabolites to sustain growth (anabolic metabolism) (1). Interestingly, most cancer cells undergo the same metabolic switch (Warburg Effect), a unique evolutionary trait that allows them to grow unabated. Although aerobic glycolysis generates much less ATP from glucose compared to oxidative phosphorylation, it provides critical intermediate metabolites that are used for anaplerotic reactions, and therefore is an obligatory adaptation among highly proliferative cells. In response to variations in nutrient availability, cells regulate their metabolic output, coordinating biochemical reactions and mitochondrial activity by altering transcription of mitochondrial genes through both activation of transcription factors, such as PGC1α, and chromatin modulators that exert epigenetic changes on metabolic genes.

Mitochondrial dysfunction has been implicated in aging, degenerative diseases and cancer. Proper mitochondrial function can be compromised by the accumulation of mutations in mitochondrial DNA that occur during aging. In addition, reactive oxygen species (ROS) produced during oxidative phosphorylation can promote oxidative damage to DNA, protein and lipids, in turn adversely affecting global cellular functions. In recent years, several studies have illustrated a novel unexpected link between metabolism and gene activity: fluctuations in mitochondrial and cytoplasmic metabolic reactions can reprogram global metabolism by means of impacting epigenetic dynamics. These studies will be briefly summarized in the first part of this article. In the second part, we will propose a provocative novel hypothesis: the crosstalk between metabolism and epigenetics is a two-way street, and defects in chromatin modulators may affect availability of intermediate metabolites, in turn influencing energy metabolism.

Metabolism impacts epigenetics

A regulated crosstalk between metabolic pathways in the mitochondria and epigenetic mechanisms in the nucleus allows cellular adaptations to new environmental conditions. Fine-tuning of gene expression is achieved by changes in chromatin dynamics, including methylation of DNA and posttranslational modifications of histones: acetyl, methyl and phosphate groups can be added by acetyltransferases, methyltransferases and kinases, respectively, to different residues on histones. Given the number of residues that can potentially undergo modifications in histone tails and in the DNA, it is reasonable to consider that metabolic changes affecting the availability of these metabolites will impact epigenetics (as discussed below).

Recently, acetylation of proteins was revealed to be as abundant as phosphorylation (2). This posttranslational modification involves the covalent binding of an acetyl group obtained from acetyl-CoA to a lysine. In histones, acetylation can modify higher order chromatin structure and serve as a docking site for histone code readers. Recent mass spectrometry studies have uncovered the complete acetylome in human cells and revealed that protein acetylation occurs broadly in the nucleus, cytoplasm and mitochondria, affecting more than 1700 proteins (2). Acetylation of proteins depends on the availability of acetyl-CoA in each cellular compartment, but this metabolite is produced in the mitochondria and cannot cross the mitochondrial membrane. In single cell eukaryotes, the pool of acetyl groups required for histone acetylation comes from the production of acetyl-CoA by the enzyme acetyl-CoA synthetase (Acs2p), which is responsible of converting acetate into acetyl-CoA. In mammalian cells, although they have a homolog enzyme to Acs2p, AceCS1, the majority of acetyl-CoA is produced from mitochondrion-derived citrate by the enzyme adenosine triphosphate (ATP)-citrate lyase (ACL) (3). ACL is present in the cytoplasm and in the nucleus, and is responsible for the production of acetyl-CoA from citrate in both compartments. Citrate is generated in the metabolism of glucose and glutamine in the TCA cycle. In contrast to acetyl-CoA, citrate can cross the mitochondrial membrane and diffuse through the nuclear pores into the nucleus, where it can be converted into acetyl-CoA by ACL. Wellen and colleagues found that ACL is required for acetylation of histones under normal growth conditions; knockdown of ACL decreases the pool of acetyl-CoA in the nucleus and reduces the level of histone acetylation (3). Strikingly, reduction in histone acetylation occurs preferentially around glycolytic genes, leading to downregulation of their transcription and therefore inhibition of glycolysis. These observations reveal a process where glucose metabolism dictates histone acetylation that in a feedback mechanism controls the rate of glycolysis.

Notably, deacetylation of histones also exhibits a metabolic influence. Deacetylation of histones is achieved by class I and class II histone deacetylases (HDACs) and by a separate class (class III), also known as sirtuins. Sirtuins use NAD+ as a cofactor for deacetylation, and the ratio of NAD+/NADH regulates their activity. In diets rich in carbohydrates, growth factors stimulate cellular glucose uptake and the production of energy is carried out through glycolysis. In this context, the NAD+/NADH ratio decreases, in turn inhibiting, in theory, sirtuins in the cytoplasm (Sirt2) and nucleus (Sirt1, Sirt6 and Sirt7). In fact, low Sirt1 and Sirt6 activity generates a global increase in protein acetylation. Interestingly, Sirt6, which is exclusively nuclear, deacetylates H3K9 Hif1α target genes, repressing their transcription. Since most of these genes are glycolytic, deacetylation of histones by Sirt6 modulates glycolysis. Indeed, SIRT6-deficient mice experience a dramatic increase in glucose uptake for glycolysis, triggering a fatal hypoglycemia in few weeks (4).

In animal cells, both histone acetylation and deacetylation are under the control of glucose metabolism through the availability of acetyl-CoA and NAD+, respectively. However, is this metabolic control restricted to acetylation, or can other reactions in the nucleus be influenced by the energy status of the cell?

Histone methyltransferases (HMTs) use S-adenosylmethionine (SAM) to transfer a methyl group onto lysine and arginine residues on histone tails. SAM is produced from methionine by the enzyme S-adenosyl methionine transferase (MAT) in a reaction that uses ATP. The recent finding of MAT in the nucleus suggests that the SAM pool could also be controlled locally in this compartment (5). The reverse reaction catalyzed by histone demethylases (HDMs) uses flavin adenine dinucleotide (FAD+) and α-ketoglutarate as coenzymes. FAD is a common redox coenzyme that exists in two different redox states. In its reduced state, FADH2 is a carrier of energy and when oxidized, FAD+ is consumed in the oxidation of succinate to fumarate by the enzyme succinate dehydrogenase (complex II) in one of the last steps of the TCA cycle. On the other hand, α-ketoglutarate is an intermediate in the TCA cycle. It is generated from isocitrate by the enzymes isocitrate dehydrogenase 1 and 2 (IDH1-cytosolic and IDH2-mitochondrial) (Figure 1A–B). Based on these findings, it is easy to infer that the amount of coenzymes used for histone methylation and demethylation could also be controlled by metabolic reactions. Moreover, the different cellular compartments compete for the same metabolites. Indeed, changes in diet that affect the biosynthesis of SAM, FAD and α-ketoglutarate in the mitochondria and cytoplasm have been shown to impact histone methylation (6).

An external file that holds a picture, illustration, etc. Object name is nihms447752f1.jpg

Figure 1   A) Diagram depicting two-way crosstalk between metabolites in cytoplasm/mitochondria and chromatin.

More recently, some of the metabolic enzymes responsible for producing cofactors for nuclear biochemical reactions have been found mutated in cancer. For instance, IDH1 and IDH2 somatic mutations are recurrent in gliomas and acute myeloid leukemias (AML). These mutations lead not only to a decreased production of α-ketoglutarate but also to a new activity: α-ketoglutarate is in fact converted into 2-hydroxyglutarate (2-HG), a metabolite rarely found in normal cells. The new metabolite is a competitive inhibitor of α-ketoglutarate-dependent dioxygenase enzymes, including the Jumonji C (JmjC) domain containing histone demethylases and the recently discovered TET family of 5-methylcytosine (5mC) hydroxylases involved in DNA demethylation (7). By inhibiting JmjC and TET enzymes, the aberrant production of 2-HG generates a genome-wide histone and DNA hypermethylation phenotype. This is considered to be, at least in part, at the origin of tumorigenesis in IDH1 and IDH2 mutated cells and for this reason, 2-HG may earn its place as an oncometabolite. The discovery that mutations in metabolic enzymes may influence tumorigenesis by means of controlling genome-wide epigenetic changes caused a paradigm shift, indicating that such metabolic abnormalities may affect cancer beyond the Warburg Effect.  ….

Chromatin modifications and cellular metabolism are tightly connected. Thus far the only aspects that have been considered are the retrograde signaling, with mitochondrial metabolites affecting histone modifications, and the anterograde transcriptional regulation of metabolism. A third aspect of the link between nucleus and metabolism has been, in our opinion, omitted so far: a direct influence of chromatin on acetyl-CoA and SAM availability, which may have an essential role also in cancer establishment and development (Figure 1A–B). Notably, a shift towards glycolytic metabolism is now considered a hallmark of cancer cells. It is also true that multiple tumors carry mutations in chromatin modifiers. However, new studies suggest that those two processes may be much more intertwined that previously appreciated, further blurring the limits on their respective roles in tumorigenesis. There is no doubt that changes in metabolite availability can drastically impact chromatin modifications. We believe that the opposite may be true as well. At least in mouse models, deficiency in two chromatin modifiers, SIRT6 and Jhdm2, causes drastic metabolic abnormalities. Even though some of those phenotypes depend on changes in gene-expression, we would like to propose that severe attrition of metabolite pools might as well play a role, a possibility that awaits experimental proof.

……

 

Investigators at UC San Diego say that when they blocked a well known signaling molecule that plays a major role in driving colorectal cancer, an escape pathway emerged that allowed tumors to continue to grow.

The pathway they explored, ERK1/2, is a problem for about a third of all colorectal cancer patients, says Petrus R. de Jong, MD, PhD, a co-first author on the paper.

“Since we were genetically deleting the ERK1/2 pathway, we expected to see less cell proliferation,” said de Jong. “Instead, the opposite occurred. There was more cell growth and loss of organization within the cells.”

The problem was ERK5, the investigators add. And when that was blocked as well in animal models and cell lines for the disease, the combination approach proved more effective in halting cancer growth.

“If you block one pathway, cancer cells usually mutate and find another pathway that ultimately allows for a recurrence of cancer growth,” said co-first author Koji Taniguchi. “Usually, mutations occur over weeks or months. But other times, as in this case, the tumor does not need to develop mutations to find an escape route from targeted therapy. When you find the compensatory pathway and block both, there is no more escape.”

 

GEN News Highlights    May 18, 2016   http://www.genengnews.com/gen-news-highlights/blocking-cancer-signaling-leads-to-discovery-of-new-tumor-promoting-pathway/81252738/
Blocking Cancer Signaling Leads to Discovery of New Tumor-Promoting Pathway

http://www.genengnews.com/Media/images/GENHighlight/thumb_115705_web4813301741.jpg

 Immunofluorescent staining of intestinal epithelium tissue shows cell growth (green). In a normal mouse model (left), cell growth is controlled, but in a mouse model with the ERK1/2 pathway blocked (right) increased cell proliferation and loss of organization occurred. [UC San Diego Health]

An international research team lead by scientists at the University of California San Diego School of Medicine uncovered some surprising results while investigating a potential therapeutic target for the ERK1 and two pathways. These signaling pathways are widely expressed and known to drive cancer growth in one-third of patients with colorectal cancer (CRC). The UCSD team found that an alternative pathway immediately emerges when ERK1/2 is halted, thus allowing tumor cell proliferation to continue.

“Since we were genetically deleting the ERK1/2 pathway, we expected to see less cell proliferation,” explained co-lead study author Petrus R. de Jong, M.D., Ph.D., translational scientist at Sanford Burnham Prebys Medical Discovery Institute. “Instead, the opposite occurred. There was more cell growth and loss of organization within the cells.”

The exciting part of this new study is investigators found that treating both ERK1/2 and the compensatory pathway ERK5 concomitantly with a combination of drug inhibitors halted CRC growth more effectively in both mouse models and human CRC cell lines.

“We show that loss of Erk1/2 in intestinal epithelial cells results in defects in nutrient absorption, epithelial cell migration, and secretory cell differentiation,” the authors wrote. “However, intestinal epithelial cell proliferation is not impeded, implying compensatory mechanisms. Genetic deletion ofErk1/2 or pharmacological targeting of MEK1/2 results in supraphysiological activity of the ERK5 pathway. Furthermore, targeting both pathways causes a more effective suppression of cell proliferation in murine intestinal organoids and human CRC lines.”

The findings from this study were published recently in Nature Communications in an article entitled “ERK5 Signalling Rescues Intestinal Epithelial Turnover and Tumour Cell Proliferation upon ERK1/2 Abrogation.”

The ERK pathway plays a critical role in embryonic development and tissue repair because it instructs cells to multiply and start dividing, but when overactivated cancer growth often occurs.

“Therapies aimed at targeting ERK1/2 likely fail because this mechanism is allowing proliferation through a different pathway,” noted senior study author Eyal Raz, M.D., professor of medicine at UC San Diego School of Medicine. “Previously, ERK5 didn’t seem important in colorectal cancer. This is an underappreciated escape pathway for tumor cells. Hence, the combination of ERK1/2 and ERK5 inhibitors may lead to more effective treatments for colorectal cancer patients.”

Currently, there are 1.2 million people living with CRC in the United States, making it the third most common cancer among men and women. In 2016 alone, an estimated 134,490 new cases are expected to be diagnosed, so understanding the molecular mechanisms that drive tumor promotion are paramount to treating this disease effectively.

“If you block one pathway, cancer cells usually mutate and find another pathway that ultimately allows for a recurrence of cancer growth,” remarked co-lead study author Koji Taniguchi, M.D., Ph.D., senior researcher at the Keio University School of Medicine in Tokyo. “Usually, mutations occur over weeks or months. But other times, as in this case, the tumor does not need to develop mutations to find an escape route from targeted therapy. When you find the compensatory pathway and block both, there is no more escape.”

The researchers were excited by their findings but urged caution at over interpretation of their initial findings and suggested that other classes of inhibitors be tested in combination with ERK5 inhibitors in human CRC cells in preclinical mouse models before any patient trial can begin.

 

ERK5 signalling rescues intestinal epithelial turnover and tumour cell proliferation upon ERK1/2 abrogation

Petrus R. de JongKoji TaniguchiAlexandra R. HarrisSamuel BertinNaoki TakahashiJen DuongAlejandro D. CamposGarth PowisMaripat CorrMichael Karin & Eyal Raz
Nature Communications 7, Article number:11551  doi:10.1038/ncomms11551

The ERK1/2 MAPK signalling module integrates extracellular cues that induce proliferation and differentiation of epithelial lineages, and is an established oncogenic driver, particularly in the intestine. However, the interrelation of the ERK1/2 module relative to other signalling pathways in intestinal epithelial cells and colorectal cancer (CRC) is unclear. Here we show that loss of Erk1/2in intestinal epithelial cells results in defects in nutrient absorption, epithelial cell migration and secretory cell differentiation. However, intestinal epithelial cell proliferation is not impeded, implying compensatory mechanisms. Genetic deletion of Erk1/2 or pharmacological targeting of MEK1/2 results in supraphysiological activity of the ERK5 pathway. Furthermore, targeting both pathways causes a more effective suppression of cell proliferation in murine intestinal organoids and human CRC lines. These results suggest that ERK5 provides a common bypass route in intestinal epithelial cells, which rescues cell proliferation upon abrogation of ERK1/2 signalling, with therapeutic implications in CRC.

The extracellular signal-regulated kinases 1 and 2 (ERK1/2) are part of the classical family of mammalian mitogen-activated protein kinases (MAPKs), which also include three c-Jun amino-terminal kinases (JNK1/2/3), four p38 isoforms and its lesser-known counterpart, ERK5. The serine/threonine kinases ERK1 (MAPK3, also known as p44 MAPK) and ERK2 (MAPK1, also known as p42 MAPK) show 83% amino acid identity, are ubiquitously expressed and typically activated by growth factors and phorbol esters, whereas the p38 and JNK pathways are mainly activated by inflammatory cytokines and stress1. MAPKs are involved in regulation of mitosis, gene expression, cell metabolism, cell motility and apoptosis. ERK1/2 are activated by MEK1 and MEK2, which themselves are activated by Raf-1, A-Raf or B-Raf1, 2. Ras proteins (K-Ras, H-Ras or N-Ras) are small GTPases that can be activated by receptor tyrosine kinases (RTKs) or G-protein coupled receptors (GPCRs), which recruit Raf proteins to the plasma membrane where they are activated. Together, these modules constitute the Ras–Raf–MEK–ERK pathway3.

The activation of ERK1/2 results in their nuclear translocation where they can phosphorylate a variety of nuclear targets such as Elk-1, c-Fos and c-Myc1, in addition to p90 ribosomal S6 kinases (p90RSKs) and mitogen- and stress-activated protein kinases, MSK1/2. The full repertoire of substrates for ERK1/2 consists of at least 160 cellular proteins4. These proteins are typically involved in the regulation of cell proliferation—more specifically, G1/S-phase cell cycle progression—and differentiation. However, their cellular effects are context-dependent and determined by the spatial and temporal dynamics of ERK1/2 activity5, which are highly regulated by scaffolding proteins and phosphatases3, 6, 7.

Despite vast literature on the role of ERK1/2 in cell proliferation, the absolute requirement of this signalling module in rapidly dividing tissues relative to other signalling pathways is unknown. The small intestinal epithelium is particularly suitable to address this question given the short (4–8 days) and dynamic life cycle of intestinal epithelial cells (IECs). Lgr5+ intestinal stem cells at the intestinal crypt base produce transit-amplifying cells, which then undergo a number of proliferative cycles before terminal differentiation into absorptive enterocytes at the crypt–villus border. Enterocytes then migrate to the villus tip where they undergo anoikis and are shed into the gut lumen8. All of these cellular events are tightly coordinated by the Wnt, Notch, bone morphogenetic protein (BMP) and Hedgehog pathways9, whereas the roles of ERK1/2 remain to be charted. In the intestines, the ERK1/2 pathway is likely activated by autocrine and paracrine factors downstream of RTKs, such as epidermal growth factor receptor (EGFR)10, and by exogenous microbial-derived substrates that signal through the Toll-like receptor (TLR)/MyD88 pathway11.

To study the effects of ERK1/2 in the adult intestinal epithelium, we generated mice with a conditional (IEC-specific) and tamoxifen-inducible deletion of Erk2 on the Erk1−/− background, which completely abrogates this pathway. We show that the ERK1/2 signalling module, surprisingly, is dispensable for IEC proliferation. Genetic deletion of Erk1/2 in primary IEC or treatment of colorectal cancer (CRC) cell lines with MEK1/2 inhibitors results in compensatory activation of the ERK5 pathway. Moreover, the treatment of human CRC lines with a combination of MEK1/2 and ERK5 inhibitors is more efficacious in the inhibition of cancer cell growth. Thus, compensatory signalling by ERK5 suggests a potential rescue pathway that has clinical implications for targeted therapy in colorectal cancer.

….

Figure 1: Wasting disease associated with malabsorption in Erk1/2ΔIEC mice.

http://www.nature.com/ncomms/2016/160517/ncomms11551/images_article/ncomms11551-f1.jpg

ERK1/2ΔIEC causes wasting and enterocyte dysfunction

Here we show that ERK1/2 signalling in mouse intestinal epithelium is dispensable for cell proliferation, while it resulted in abnormal differentiation of enterocytes, wasting disease and ultimately lethality (Fig. 1). Consistent with these findings, ERK1/2 MAPKs were shown to be associated with the enterocyte brush border and activated upon RTK stimulation or feeding27 or electrical field stimulation in polarized epithelium28. This seems at odds with literature that suggest that maintained ERK1/2 signalling precludes enterocyte differentiation29, 30. A possible explanation for this discrepancy could be that cycling IEC in the transit amplifying zone of the crypt require relatively high levels of active ERK1/2, which is readily blocked by pharmacological intervention, whereas a transition to low level ERK1/2 activity in IEC migrating into the villus compartment promotes the absorptive enterocyte differentiation program that is only perturbed upon complete genetic deletion of Erk1/2. Little is known about the role of ERK1/2 signalling in the life cycle of secretory cells in the gut. A recent report by Heuberger et al.15 described that IEC-specific deletion of non-receptor tyrosine phosphatase, Shp2, resulted in the loss of p-ERK1/2 levels in the small intestine. This coincided with an increased number of Paneth cells at the expense of goblet cells in the small intestine, as well as shortening of villi. They also observed the strongest staining for epithelial p-ERK1/2 in the TA zone. This p-ERK1/2+ staining pattern and the architectural organization of the TA zone were lost in Shp2 knockout mice. Interestingly, the deleterious effects of Shp2 deficiency were rescued by expression of constitutively active MEK1. A model was proposed in which the balance between Wnt/β-catenin and MAPK signalling determines Paneth cell versus goblet cell differentiation, respectively15. This proposed crucial role for ERK1/2 MAPK signalling in intestinal secretory cell differentiation is consistent with our observations inERK1/2ΔIECmice.

Migration and differentiation are functionally intertwined in the intestines, as demonstrated by the immature phenotype of mislocalized Paneth cells observed in ΔIEC mice (Fig. 2). Critical to epithelial cell migration is proper cytoskeleton reorganization mediated by the small GTPases of the Rho family, cell polarization regulated by Cdc42 and dynamic adhesion through cell–matrix and cell–cell interaction via integrin/FAK/Src signalling31. The ERK1/2 module is used as a downstream effector of many of these pathways in the intestine, including Rho GTPases32, FAK33and Src34, and has been suggested to promote cell motility33, 35. RTK signalling also contributes to cell migration, for example, Eph–Ephrin receptor interactions are crucial for correct positioning of Paneth cells36. Ephrin receptor-induced epithelial cell migration has been shown to be mediated by Src and ERK1/2 activation37, 38, which may explain the Paneth cell mislocalization observed in ΔIEC mice. In summary, the ERK1/2 module is indispensable for full maturation of both absorptive enterocytes and the secretory lineage (Fig. 7a), confirming its crucial role in the integration of cellular cues required for determination of epithelial cell fate.

Figure 7: Roles of ERK1/2 and ERK5 in intestinal homeostasis and tumorigenesis.

Roles of ERK1/2 and ERK5 in intestinal homeostasis and tumorigenesis.

(a) When the ERK1/2 pathway is intact, extracellular cues that are transduced via RTKs or GPCRs activate Ras under physiological conditions, or alternatively, Ras is constitutively active in colorectal cancer (RasΔ*), which preferentially activates the Raf–MEK1/2–ERK1/2 module. The nuclear and transcriptional targets of ERK1/2 are crucial for enterocyte and secretory cell differentiation, IEC migration, as well as cell proliferation under homeostatic and oncogenic conditions. Importantly, ERK1/2 activation also results in the activation of negative feedback mechanisms that suppress its upstream kinases (for example, RTKs, son of sevenless, Raf) and activate dual specificity phosphatases (DUSPs), resulting in the silencing of the ERK5 module. (b) Upon abrogation of MEK1/2 or genetic knockout ofErk1/2, the lack of negative feedback mechanisms (that is, feedback activation) results in upregulation of the Ras–Raf–MEK5–ERK5 module, which maintains cell proliferation under physiological conditions, or results in continued tumour cell proliferation in colorectal cancer, respectively. However, the lack of activation of ERK1/2-specific targets results in differentiation and migration defects of intestinal epithelial cells culminating in malabsorption, wasting disease and mortality. Compensatory upregulation of the ERK5 pathway in CRC can be reversed by targeted treatment with its specific inhibitor, XMD8-92.

An unexpected finding was the redundancy of ERK1/2 in the gut with regard to cell proliferation.Erk1/2 deletion was compensated by upregulated ERK5 signalling. Genetic targeting of ERK1/2 in vitro previously showed that Erk2 knockdown is more effective than Erk1 knockdown in suppressing cell proliferation, although this may be related to higher expression levels of the former39. The effect of gene dosage was demonstrated in vivo by the observations that whileErk1−/− mice are viable12 and Erk2−/− mice die in utero13, Erk2+/− mice are only viable when at least one copy of Erk1 is present. However, mice heterozygous (+/−) for both Erk1 and Erk2 alleles were born at lower than Mendelian ratio39. More recently, it was reported that transgenic expression of ERK1 can compensate for Erk2 deletion40, demonstrating functional redundancy between both family members. Deletion of Erk1/2 in adult skin tissue resulted in hypoplasia, which was associated with G2/M cell cycle arrest, without notable differentiation defects of keratinocytes41. These data differ from our observations in the intestines, which might be explained by incomplete and transient siRNA-mediated knockdown of ERK1/2 in primary keratinocyte cultures41, compared with more efficient genomic deletion of Erk1 and Erk2 that is typically achieved by the Villin-Cre-ERT2 system14, possibly resulting in different outcomes.

Both ERK1/2 and ERK5 have been described to promote cell cycle progression, although they have different upstream signalling partners, MEK1/2 and MEK5, respectively1. Furthermore, ERK2 and ERK5 proteins share only about 66% sequence identity, and MEK5 is phosphorylated by MEKK2/3, which can also activate the p38 and JNK pathways42. The ERK5 pathway is classically activated by stress stimuli, in addition to mitogens; thus, it shares features of both the ERK1/2, and p38 and JNK pathways, respectively43. ERK5 induces expression of cyclin D1 (refs 44, 45), and suppresses expression of cyclin dependent kinase inhibitors46, thereby promoting G1/S-phase cell cycle progression. Importantly, the role of ERK5 in IEC differentiation and intestinal homeostasis is currently unknown. Knockout of Erk1/2 in IEC induced activity of ERK5, which was not detectable in naive mice (Fig. 3). These data suggest that the ERK1/2 and ERK5 modules may share proximal signalling components. Although EGFR is a likely candidate in this context19, 20, we found that abrogation of EGFR signalling did not prevent enhanced ERK5 activity upon MEK1/2 inhibition. Although it was originally suggested that ERK5 signalling is independent of Ras20, other groups established that Ras, either through physiological signalling47, or by its oncogenic activity48,49, activates the MEK5–ERK5 signalling axis. Thus, rewiring of signalling networks downstream of Ras could explain the supraphysiological activity of ERK5 upon conditional deletion of Erk1/2 in the intestines. In fact, it has been shown that ERK1/2 signalling mediates negative feedback on ERK5 activity50, possibly through transcriptional activation of dual specificity phosphatases (DUSPs)51. Alternatively, ERK1/2-induced FOS-like antigen 1 (Fra-1) may negatively regulate MEK5 (ref. 52). These data suggest that ERK5 is a default bypass route downstream of RTK-Ras and activated upon loss of ERK1/2-mediated repression, thereby ensuring the transduction of mitogenic signals to the nucleus (Fig. 7b). Consistent with this concept, we found that ERK5 inhibition induces atrophy of ΔIEC intestinal organoids (Fig. 4). In addition, important downstream transcriptional targets of ERK5 and ERK1/2 overlap, such as immediate-early gene Fra1 and oncogene c-Myc, whereas c-Fos and Egr1 were specifically induced by ERK1/2 (Fig. 6 and Supplementary Fig. 7). Specificity of ERK1/2 over ERK5 and other MAPK family members for the activation of c-Fos has been previously described53, demonstrating their differential biological output despite the shared ability to transduce potent mitogenic signals.

Our findings may be relevant for the use of MAPK inhibitors in the treatment of colorectal cancer. Although there was only a mild phenotype in the colons of ΔIEC mice under homeostatic conditions, the Ras–RAF–MEK–ERK pathway is generally upregulated in malignant cells including CRC54. Targeted therapy typically results in feedback activation of upstream players of the targeted kinase, which are then able to reactivate the same pathway or utilize bypass signalling routes55. For example, on activation, ERK1/2 phosphorylates EGFR, son of sevenless56, and Raf57, thereby terminating upstream signalling activity. Knockout of Erk1/2 eliminates this negative feedback. Our data suggest that ERK5 is a putative resistance pathway in the context of targeted treatment with MEK1/2 or ERK1/2 inhibitors (Fig. 7b). Different classes of MEK1/2 inhibitors display various modes of resistance to therapy (innate, adaptive and acquired)58. Since we have only used one MEK1/2 inhibitor (PD0325901) in our studies, it will be necessary to evaluate other classes of inhibitors in combination with ERK5 inhibitors. Importantly, while treatment with either the MEK1/2 or ERK5 inhibitor suppressed tumour growth in murine Apc−/− organoids, only the latter was able to inhibit the proliferation of Apc−/−;KRASG12V organoids (Fig. 6), which are more representative of human CRC. In line with this, suppression of ERK5 expression by forced expression of miR-143/145 inhibited intestinal adenoma formation in the ApcMin/+ model59, and activated MEK5 correlated with more invasive CRC in human60. ERK5 has been previously reported to mediate resistance to cytotoxic chemotherapy-induced apoptosis61. The highly specific and bioavailable ERK5 inhibitor, XMD8-92, has shown antitumour effects in multiple preclinical cancer models by inhibiting tumour angiogenesis, metastasis and chemo-resistance62. Furthermore, ERK5 inhibition does not induce feedback activation of upstream or parallel signalling pathways62. In conclusion, the ERK1/2 and ERK5 MAPK modules display a high degree of signalling plasticity in the intestinal epithelium, which has implications for targeted treatment of colorectal cancer.

 

Researchers Reveal Role of Transcription Factor Isoforms in Colon Diseases

http://www.genengnews.com/gen-news-highlights/researchers-reveal-role-of-transcription-factor-isoforms-in-colon-diseases/81252735/

http://www.genengnews.com/Media/images/GENHighlight/thumb_May18_2016_PoonamjotDeol_ColonicCrypt2521803322.jpg


 

 

 

 

 

 

 

 

Balance between the two isoforms, P1 and P2, of nuclear receptor HNF4a in the colonic crypt influences susceptibility to colitis and colon cancer. P1 is seen here in green. P2 is seen in red. [Poonamjot Deol, Sladek lab, UC Riverside]

Scientists at the University of California, Riverside have determined the distribution of the P1 and P2 isoforms of hepatocyte nuclear factor 4α (HNF4α) in the colons of mice. They report (“Opposing Roles of Nuclear Receptor HNF4α Isoforms in Colitis and Colitis-Associated Colon Cancer”) in eLife that maintaining a balance of P1 and P2 is crucial for reducing risk of contracting colon cancer and colitis.

What is already known in the field of cell biology is that the HNF4α transcription factor plays a key role in both diseases. HNF4α comes in two major isoforms, P1-HNF4α and P2-HNF4α (P1 and P2), but just how each isoform is involved in colitis and colon cancer is not understood.

“P1 and P2 have been conserved between mice and humans for 70 million years,” said Frances M. Sladek, Ph.D., professor of cell biology, who led the research project. “Both isoforms are important and we want to keep an appropriate balance between them in our gut by avoiding foods that would disrupt this balance and consuming foods that help preserve it. What these foods are is our next focus in the lab.”

The intestine is the only adult tissue in the body that expresses both P1 and P2. Dr. Sladek and her team have shown for the first time that these isoforms perform nonredundant functions in the intestine and are relevant to colitis and colitis-associated colon cancer.

“Our study also suggests that finding a drug to stabilize one isoform should be more effective than targeting both isoforms for treating colitis and colon cancer,” said Karthikeyani Chellappa, Ph.D., the first author of the research paper and a former postdoctoral researcher in Sladek’s lab.

Dr. Sladek explained that the colonic epithelial surface has finger-like invaginations (into the colonic wall) called colonic crypts that house stem cells at their base. These stem cells help regenerate new epithelial cells that continuously migrate up toward the surface, thus ensuring complete renewal of the intestinal lining every 3–5 days.

The researchers observed that the P1-positive cells were found in the surface lining and the top portion of the crypt (green in the accompanying image) whereas P2-positive cells were mostly in the proliferative compartment in the lower half of the crypt (the proliferation marker is red in the image.) Furthermore, when transgenic mice genetically engineered to have only either P1 or P2 were subjected to a carcinogen and, subsequently, to an irritant to stress the epithelial lining of the colon, the researchers found that the P1 mice showed fewer tumors than wild-type control mice. When treated with irritant alone, these mice were resistant to colitis. In sharp contrast, mice with only P2 showed more tumors and were much more susceptible to colitis.

The researchers explain these findings by invoking the “barrier function,”  a mucosal barrier generated by the colon’s epithelial cells that prevents bacteria in the gut from entering the body. In the case of P1 mice, this barrier function was enhanced. The P2 mice, on the other hand, showed a compromised barrier function, presumably allowing bacteria to pass through.

Next, the researchers examined genes expressed in the P1 and P2 mice. They found that resistin-like molecule (RELM)-beta, a cytokine (a signaling molecule of the immune system) expressed in the gastrointestinal tract and implicated in colitis, was expressed far more in the P2 mice than the P1 mice.

“This makes sense since a reduced barrier function means bacteria can go across the barrier, which activates RELM-beta,” Dr. Sladek said. “We also found that the P2 protein transcribes RELM-beta more effectively than the P1 protein.”

Next, Poonamjot Deol, Ph.D.,  an assistant project scientist in Dr. Slaked’s lab and the second author of the eLife study, will lead a project aimed at understanding how diet affects the distribution of P1 and P2 in the gut. She and others in the lab also plan to investigate how obesity and colitis may be linked. (Diet studies performed in Dr. Sladek’s lab in the past illustrated soybean oil’s adverse effect on obesity.)

“In the case of colitis, could soybean oil be playing a part in allowing bacteria to get across the barrier function?” Dr. Deol said. “We do not know. We know its detrimental effect on obesity. But more research needs to be done where colitis is concerned.”

Opposing roles of nuclear receptor HNF4α isoforms in colitis and colitis-associated colon cancer

 Karthikeyani Chellappa, 

HNF4α has been implicated in colitis and colon cancer in humans but the role of the different HNF4α isoforms expressed from the two different promoters (P1 and P2) active in the colon is not clear. Here, we show that P1-HNF4α is expressed primarily in the differentiated compartment of the mouse colonic crypt and P2-HNF4α in the proliferative compartment. Exon swap mice that express only P1- or only P2-HNF4α have different colonic gene expression profiles, interacting proteins, cellular migration, ion transport and epithelial barrier function. The mice also exhibit altered susceptibilities to experimental colitis (DSS) and colitis-associated colon cancer (AOM+DSS). When P2-HNF4α-only mice (which have elevated levels of the cytokine resistin-like β, RELMβ, and are extremely sensitive to DSS) are crossed with Retnlb-/- mice, they are rescued from mortality. Furthermore, P2-HNF4α binds and preferentially activates the RELMβ promoter. In summary, HNF4α isoforms perform non-redundant functions in the colon under conditions of stress, underscoring the importance of tracking them both in colitis and colon cancer.

 

Read Full Post »

Aduro Biotech Phase II Pancreatic Cancer Trial CRS-207 plus cancer vaccine GVAX Fails

Reporter: Stephen J. Williams, Ph.D

From Biospace News

May 16, 2016
By Alex Keown, BioSpace.com Breaking News Staff

BERKELEY, Calif. – Shares of Aduro Biotech (ADRO) have fallen more than 25 percent this morning following news that the company’s Phase II trial for its combination pancreatic cancer drug, CRS-207 did not meet its primary endpoint of survivability.

Aduro said its Eclipse trial of CRS-207 failed to show an improvement in overall survival for patients with pancreatic cancer who had failed at least two prior therapies in the metastatic setting. Median overall survival was 3.8 months for patients treated with the immunotherapy regimen of CRS-207 and the cancer vaccine GVAX pancreas, 5.4 months for patients treated with CRS-207 alone and 4.6 months for patients administered chemotherapy. Aduro said there were no reported safety concerns during the trial and full study findings will be presented at a later date.

Stephen T. Isaacs, chairman, president and chief executive officer of Aduro, called the findings a disappointing and “unexpected outcome.’

“While we are well aware of the very difficult-to-treat nature of late-stage metastatic pancreatic cancer, we are surprised by the divergence of these data from the results of our Phase IIa study. At the same time, we continue to look forward to the interim results later this year from our ongoing Stellar trial, which is evaluating CRS-207 and GVAX Pancreas with and without the anti-PD1 checkpoint inhibitor nivolumab as a second-line therapy for patients with metastatic pancreatic cancer,” Isaacs said in a statement.

For full story please see http://www.biospace.com/News/aduro-biotechs-stock-craters-after-pancreatic/419628/source=TopBreaking

Also from FierceBiotech

UPDATED: Aduro combo fails in a key pancreatic cancer study

Read Full Post »

Older Posts »

%d bloggers like this: