Feeds:
Posts
Comments

Posts Tagged ‘health’

Early Diagnosis

Reporter: Stephen J. Williams, Ph.D.

This post contains a curation of all Early Diagnosis posts on this site as well as a curation of the Early Detection Research Network.

Early Research Detection Network (EDRN)

Welcome to EDRN

The Early Detection Research Network (EDRN), an initiative of the National Cancer Institute (NCI), brings together dozens of institutions to help accelerate the translation of biomarker information into clinical applications and to evaluate new ways of testing cancer in its earliest stages and for cancer risk.

Getting Started…

Check out the EDRN Highlights — a listing of our accomplishments and milestones.

 

► Scientific Components ► For Public, Patients, Advocates
► Collaborative Opportunities (how to join EDRN) ► For Researchers

Highlights

Highlights of the accomplishments of the Early Detection Research Network.

A brief list of major EDRN-developed assays that have been adapted for clinical use is described in the table below:

Detection/Biomarker Assay Discovery Refine/Adapt for Clin Use Clinical Validation Clinical Translation
Blood proPSA FDA approved
Urine PCA3 FDA approved
OVA1™ for Ovarian Cancer FDA approved
ROMA Algorithm for CA125 and HE4 Tests for Pelvic Mass Malignancies FDA approved
Blood/DCP and AFP-L3 for Hepatocellular Carcinoma FDA approved
Blood GP73 Together with AFP-L3 used  for monitoring cirrhotic patients for HCC in China
MiPS (Mi Prostate Score Urine test), Multiplex analysis of T2-ERG gene fusion, PCA3 and serum PSA In CLIA Lab
FISH to detect T2S:Erg fusion for Prostate Cancer In CLIA Lab
GSTP1 methylation for repeat biopsies in prostate cancer In CLIA Lab
Mitochondrial deletion for detection of prostate cancer In CLIA Lab
Somalogic 12-marker panel for Lung Cancer In CLIA Lab
80-gene panel for Lung Cancer In CLIA Lab
Vimentin Methylation Marker for Colon Cancer In CLIA Lab
Galectin-3 ligand for detection of adenomas and colon cancer In CLIA Lab
8-gene panel for Barrett’s Esophagus In CLIA Lab
SOPs for Blood (Serum, Plasma), Urine, Stool Frequently used by biomarker research community
EDRN Pre/Validation Specimen Reference Sets (specimens from well characterized and matched cases and controls from specific disease spectra) Frequently used by biomarker research community

Since its inception in 1999 EDRN has achieved several key milestones, summarized below:

1998 through 2000: Inception and Inauguration of EDRN

2001 to 2003: Meeting the Challenges to Harness and Share Emerging Scientific Knowledge

  • EDRN Second Report, Translational Research to Identify Early Cancer and Cancer Risk, October 2002, http://edrn.nci.nih.gov/docs.) published.
  • EDRN joined the Gordon Research Conferences to co-host the New Frontiers in Cancer detection and Diagnosis in 2002.

 

  • Guidelines Set for Studies Measuring Biomarker Predictive Power Journal of National Cancer Institute (Vol. 93, No. 14, July 18, 2001).
  • EDRN Associate Membership Program Initiated: This novel approach to make EDRN inclusive has been extremely successful. EDRN has now more than 120 Associate Members who are significantly contributing to EDRN efforts in biomarker discovery, development and validation.

2003 to 2004: Network Surges Ahead in Real-time

  • Collaborative Discovery and Validation Projects:  More than 100 collaborative projects spanned the various organ sites. These projects are monitored through the EDRN’s electronic System Information System (eSIS).
  • EDRN Virtual Specimen Bank and Validation Management System Launched: The EDRN Virtual Specimen Bank, also known as ERNE knowledge system, was deployed to 10 institutions in early 2003, allowing a common web-based query to search for available specimens across the EDRN Clinical Epidemiology and Validation Centers https://ginger.fhcrc.org/edrn/imp/GateServlet?pwd. VSIMS was created to allow multiple studies to be administered efficiently by minimizing development time with standardization of information and data management across multiple activities and research sites. This system encompasses all the security features of Food and Drug Administration (FDA)-required auditing systems.
  • Partnership on the Plasma Proteome Project (PPP) Initiative of the Human Proteome Organization (HUPO): PPP project was initiated to evaluate multiple technology platforms, develop bioinformatic tools and standards for protein identification, and create a database of the plasma proteome. The entire study was published in the August issue of the journal Proteomics August 2005, Volume 4 (4), pp 1045-1450.

2005 to 2008: An Investment in Prevention

  • In late 2006, EDRN’s Program for Rapid, Independent Diagnostic Evaluation (PRIDE), was established (http://grants.nih.gov/grants/guide/notice-files/NOT-CA-07-003.html ) as an administrative means to assist extramural investigators in successfully conducting cross-laboratory validation of biomarkers. Ten applications have been reviewed and five are being supported.
  • EDRN underwent external reviews in 2007 and 2008.
  • The Canary Foundation, Palo Alto, CA signed a Memorandum of Understanding with EDRN, NCI on supporting prostate cancer surveillance network of investigators from seven institutions. The tissue and serum will be collected during a three-year period and will be made available to extramural scientists for discovery and validation research.
  • The Lustgarten Foundation, N.Y., funded 6 institutions to generate monoclonal antibodies and associated hybridoma cell lines for pancreatic cancer antigens (biomarkers) identified by EDRN and non-EDRN investigators. These resources will be stored at the NCI-Frederick Facility for distribution to extramural investigators.

2009 to 2011: Realizing Investment for Clinical Use

  • Two biomarker tests approved by FDA and two IVDs pending FDA review.
  • Six biomarker tests offered by CLIA labs.
  • One biomarker test approved for clinical use outside the USA

A Curation of Posts on Early Detection of Cancer and Other Early Detection Networks is Included Below

 

BRCA 1 and 2 and Early Detection of Cancer

Early Detection of Prostate Cancer: American Urological Association (AUA) Guideline

Mechanism involved in Breast Cancer Cell Growth: Function in Early Detection & Treatment

Warning signs may lead to better early detection of ovarian cancer

Cancer Detection

Biomarker tool development for Early Diagnosis of Pancreatic Cancer: Van Andel Institute and Emory University

China, India, and Russia account for 46% of all new cancer cases globally, as well as 52% of cancer-related mortality per 4/2014 Lancet Oncology article

 

Read Full Post »

New Generation of Platinated Compounds to Circumvent Resistance

Curator/Writer: Stephen J. Williams, Ph.D.

Resistance to chemotherapeutic drugs continues to be a major hurdle in the treatment of neoplastic disorders, irregardless if the drug is a member of the cytotoxic “older” drugs or the cytostatic “newer” personalized therapies like the tyrosine kinase inhibitors.  For the platinatum compounds such as cisplatin and carboplatin, which are mainstays in therapeutic regimens for ovarian and certain head and neck cancers, development of resistance is often regarded as the final blow, as new options for these diseases have been limited.

Although there are many mechanisms by which resistance to platinated compounds may develop the purpose of this posting is not to do an in-depth review of this area except to refer the reader to the book   Ovarian Cancer and just to summarize the well accepted mechanisms of cisplatin resistance including:

  • Decreased cellular cisplatin influx
  • Increased cellular cisplatin efflux
  • Increased cellular glutathione and subsequent conjugation, inactivation
  • Increased glutathione-S-transferase activity (GST) and subsequent inactivation, conjugation
  • Increased γ-GGT
  • Increased metallothionenes with subsequent conjugation, inactivation
  • Increased DNA repair: increased excision repair
  • DNA damage tolerance: loss of mismatch repair (MMR)
  • altered cell signaling activities and cell cycle protein expression

Williams, S.J., and Hamilton, T.C. Chemotherapeutic resistance in ovarian cancer. In: S.C. Rubin, and G.P. Sutton (eds.), Ovarian Cancer, pp.34-44. Lippincott, Wilkins, and Williams, New York, 2000.

Also for a great review on clinical platinum resistance by Drs. Maritn, Hamilton and Schilder please see the following Clinical Cancer Research link here.

This curation represents the scientific rationale for the development of a new class of platinated compounds which are meant to circumvent mechanisms of resistance, in this case the loss of mismatch repair (MMR) and increased tolerance to DNA damage.

An early step in the production of cytotoxicity by the important anticancer drug cisplatin and its analog carboplatin is the formation of intra- and inter-strand adducts with tumor cell DNA 1-3. This damage triggers a cascade of events, best characterized by activation of damage-sensing kinases (reviewed in 4), p53 stabilization, and induction of p53-related genes involved in apoptosis and cell cycle arrest, such as bax and the cyclin-dependent kinase inhibitor p21waf1/cip1/sdi1 (p21), respectively 5,6. DNA damage significantly induces p21 in various p53 wild-type tumor cell lines, including ovarian carcinoma cells, and this induction is responsible for the cell cycle arrest at G1/S and G2/M borders, allowing time for repair 7,8.  DNA lesions have the ability of  to result in an opening of chromatin structure, allowing for transcription factors to enter 56-58.  Therefore the anti-tumoral ability of cisplatin and other DNA damaging agents is correlated to their ability to bind to DNA and elicit responses, such as DNA breaks or DNA damage responses which ultimately lead to cell cycle arrest and apoptosis.  Therefore either repair of such lesions, the lack of recognition of such lesions, or the cellular tolerance of such lesions can lead to resistance of these agents.

resistmech2

Mechanisms of Cisplatin Sensitivity and Resistance. Red arrows show how a DNA lesion results in chemo-sensitivity while the beige arrow show common mechanisms of resistance including increased repair of the lesion, effects on expression patterns, and increased inactivation of the DNA damaging agent by conjugation reactions

 

 

 

 

 

 

 

 

 

 

 

 

 

 

mechPtresistance

 

 

Increased DNA Repair Mechanisms of Platinated Lesion Lead to ChemoResistance

 

DNA_repair_pathways

Description of Different Types of Cellular DNA Repair Pathways. Nucleotide Excision Repair is commonly up-regulated in highly cisplatin resistant cells

 

 

 

 

 

 

 

 

 

 

 

Loss of Mismatch Repair Can Lead to DNA Damage Tolerance

dnadamage tolerance

 

 

 

 

 

 

 

 

In the following Cancer Research paper Dr. Vaisman in the lab of Dr. Steve Chaney at North Carolina (and in collaboration with Dr. Tom Hamilton) describe how cisplatin resistance may arise from loss of mismatch repair and how oxaliplatin lesions are not recognized by the mismatch repair system.
Cancer Res. 1998 Aug 15;58(16):3579-85.

The role of hMLH1, hMSH3, and hMSH6 defects in cisplatin and oxaliplatin resistance: correlation with replicative bypass of platinum-DNA adducts.

Abstract

Defects in mismatch repair are associated with cisplatin resistance, and several mechanisms have been proposed to explain this correlation. It is hypothesized that futile cycles of translesion synthesis past cisplatin-DNA adducts followed by removal of the newly synthesized DNA by an active mismatch repair system may lead to cell death. Thus, resistance to platinum-DNA adducts could arise through loss of the mismatch repair pathway. However, no direct link between mismatch repair status and replicative bypass ability has been reported. In this study, cytotoxicity and steady-state chain elongation assays indicate that hMLH1 or hMSH6 defects result in 1.5-4.8-fold increased cisplatin resistance and 2.5-6-fold increased replicative bypass of cisplatin adducts. Oxaliplatin adducts are not recognized by the mismatch repair complex, and no significant differences in bypass of oxaliplatin adducts in mismatch repair-proficient and -defective cells were found. Defects in hMSH3 did not alter sensitivity to, or replicative bypass of, either cisplatin or oxaliplatin adducts. These observations support the hypothesis that mismatch repair defects in hMutL alpha and hMutS alpha, but not in hMutS beta, contribute to increased net replicative bypass of cisplatin adducts and therefore to drug resistance by preventing futile cycles of translesion synthesis and mismatch correction.

 

 

The following are slides I had co-prepared with my mentor Dr. Thomas C. Hamilton, Ph.D. of Fox Chase Cancer Center on DNA Mismatch Repair, Oxaliplatin and Ovarina Cancer.

edinborough2mmrtranslesion1

 

 

 

 

 

 

Multiple Platinum Analogs of Cisplatin (like Oxaliplatin )Had Been Designed to be Sensitive in MMR Deficient Tumors

edinborough2diffptanalogs

 

 

 

 

 

 

mmroxaliplatin

 

 

 

 

 

 

edinborough2ptanalogsresist

 

 

 

 

 

 

edinborough2relresistptanalogsdifflines

 

 

 

 

 

 

edinborough2msimlmh2refract

 

 

 

 

 

 

edinborough2gogoxaliplatintrial

 

 

 

 

 

 

 

Please see below video on 2015 Nobel Laureates and their work to elucidate the celluar DNA repair mechanisms.

Clinical genetics expert Kenneth Offit gives an overview of Lynch syndrome, a genetic disorder that can cause colon (HNPCC) and other cancers by defects in the MSH2 DNA mismatch repair gene. (View Video)

 

 

References

  1. Johnson, S. W. et al. Relationship between platinum-DNA adduct formation, removal, and cytotoxicity in cisplatin sensitive and resistant human ovarian cancer cells. Cancer Res 54, 5911-5916 (1994).
  2. Eastman, A. The formation, isolation and characterization of DNA adducts produced by anticancer platinum complexes. Pharmacology and Therapeutics 34, 155-166 (1987).
  3. Zhen, W. et al. Increased gene-specific repair of cisplatin interstrand cross-links in cisplatin-resistant human ovarian cancer cell lines. Molecular and Cellular Biology 12, 3689-3698 (1992).
  4. Durocher, D. & Jackson, S. P. DNA-PK, ATM and ATR as sensors of DNA damage: variations on a theme? Curr Opin Cell Biol 13, 225-231 (2001).
  5. el-Deiry, W. S. p21/p53, cellular growth control and genomic integrity. Curr Top Microbiol Immunol 227, 121-37 (1998).
  6. Ewen, M. E. & Miller, S. J. p53 and translational control. Biochim Biophys Acta 1242, 181-4 (1996).
  7. Gartel, A. L., Serfas, M. S. & Tyner, A. L. p21–negative regulator of the cell cycle. Proc Soc Exp Biol Med 213, 138-49 (1996).
  8. Chang, B. D. et al. p21Waf1/Cip1/Sdi1-induced growth arrest is associated with depletion of mitosis-control proteins and leads to abnormal mitosis and endoreduplication in recovering cells. Oncogene 19, 2165-70 (2000).
  9. Davies, N. P., Hardman, L. C. & Murray, V. The effect of chromatin structure on cisplatin damage in intact human cells. Nucleic Acids Res 28, 2954-2958 (2000).
  10. Vichi, P. et al. Cisplatin- and UV-damaged DNA lure the basal transcription factor TFIID/TBP. Embo J 16, 7444-7456 (1997).
  11. Xiao, G. et al. A DNA damage signal is required for p53 to activate gadd45. Cancer Res 60, 1711-9 (2000).

Other articles in this Open Access Journal on ChemoResistance Include:

Cancer Stem Cells as a Mechanism of Resistance

An alternative approach to overcoming the apoptotic resistance of pancreatic cancer

Mutation D538G – a novel mechanism conferring acquired Endocrine Resistance causes a change in the Estrogen Receptor and Treatment of Breast Cancer with Tamoxifen

Can IntraTumoral Heterogeneity Be Thought of as a Mechanism of Resistance?

Nitric Oxide Mitigates Sensitivity of Melanoma Cells to Cisplatin

Heroes in Medical Research: Barnett Rosenberg and the Discovery of Cisplatin

Read Full Post »

Cancer Stem Cells as a Mechanism of Resistance

 

Curator: Stephen J. Williams, Ph.D.

The cancer stem-cell hypothesis proposes the existence of a subset of cells within a heterogeneous tumor cell population that have stem-cell like properties [1], and may be essential for the progression and metastases of epithelial malignancies, by providing a reservoir of cells that self-renew and differentiate into the bulk of the tumor [2]. The stem-cell hypothesis implies that similar genetic regulatory pathways might define critical stem-cell like functions, such as self-renewal and pluripotency, in both normal and cancer stem-cells. Indeed, cancer stem-cells have been identified in many tumor types, such as breast [3], pancreas [4] and ovarian [5], based on screening with cellular markers typically found in normal stem-cells such as CD44, ALDH1, and CD133 (reviewed in [2]). A number of studies have suggested that the expression of these stem-cell markers is correlated with poor prognosis [6-9]. The ability to identify and isolate these populations may have a significant impact on design of individualized therapies.

Great general posts and good review on this site about Cancer Stem Cells, their markers, and ability to target them with chemotherapy can be seen here.

In Focus: Identity of Cancer Stem Cells

In Focus: Targeting of Cancer Stem Cells

Stem Cells and Cancer

 

However, there has been growing acknowledgement of the ability of cancer stem cell populations to resist the cytotoxic effects of most chemotherapeutic agents, including cisplatin, topoisomerase inhibitors, DNA damaging agents, and even tyrosine kinase inhibitors (TKI). Indeed, some feel that intrinsic resistance to cytotoxic drugs may be a biological feature of cancer stem cells.

Definitions:

Acquired resistance: a resistance to a particular drug which results following continued exposure to said drug. Can take days (in cases of some TKIs) or months to develop. Acquired resistant cells lines are developed by exposure to increasing drug concentration over a time period (either intermittent exposure or continuous exposure)

Intrinsic resistance: a pre-existing resistance usually termed refractory where cancer cells THAT HAVE NOT BEEN EXPOSED to drug, do not respond to initial drug exposure. Can be seen experimentally in panels of unrelated cancer cells lines isolated from untreated patients which show no cytotoxicity to drug exposure in vitro.

Below is one of the first reports which described the drug resistant phenotype of cancer stem cells in an in vivo (mouse) model of breast cancer with videos.

Cancer Res. 2008 May 1;68(9):3243-50. doi: 10.1158/0008-5472.CAN-07-5480.

Cancer stem cells contribute to cisplatin resistance in Brca1/p53-mediated mouse mammary tumors.

Shafee N1, Smith CR, Wei S, Kim Y, Mills GB, Hortobagyi GN, Stanbridge EJ, Lee EY.

Author information

Abstract

The majority of BRCA1-associated breast cancers are basal cell-like, which is associated with a poor outcome. Using a spontaneous mouse mammary tumor model, we show that platinum compounds, which generate DNA breaks during the repair process, are more effective than doxorubicin in Brca1/p53-mutated tumors. At 0.5 mg/kg of daily cisplatin treatment, 80% primary tumors (n = 8) show complete pathologic response. At greater dosages, 100% show complete response (n = 19). However, after 2 to 3 months of complete remission following platinum treatment, tumors relapse and become refractory to successive rounds of treatment. Approximately 3.8% to 8.0% (mean, 5.9%) of tumor cells express the normal mammary stem cell markers, CD29(hi)24(med), and these cells are tumorigenic, whereas CD29(med)24(-/lo) and CD29(med)24(hi) cells have diminished tumorigenicity or are nontumorigenic, respectively. In partially platinum-responsive primary transplants, 6.6% to 11.0% (mean, 8.8%) tumor cells are CD29(hi)24(med); these populations significantly increase to 16.5% to 29.2% (mean, 22.8%; P < 0.05) in platinum-refractory secondary tumor transplants. Further, refractory tumor cells have greater colony-forming ability than the primary transplant-derived cells in the presence of cisplatin. Expression of a normal stem cell marker, Nanog, is decreased in the CD29(hi)24(med) populations in the secondary transplants. Top2A expression is also down-regulated in secondary drug-resistant tumor populations and, in one case, was accompanied by genomic deletion of Top2A. These studies identify distinct cancer cell populations for therapeutic targeting in breast cancer and implicate clonal evolution and expansion of cancer stem-like cells as a potential cause of chemoresistance.

Please Watch Videos

 

Below is a curation of talks and abstracts from the 2015 Annual AACR Meeting in Philadelphia, PA.

The Talk by Dr. Cheresh is an example of this school of thought; that inducing cancer cell stemness can result in development of drug resistance, in this case to a TKI. (For a press release on this finding see here.)

SY27-04: Induction of cancer stemness and drug resistance by EGFR blockade
Tuesday, Apr 21, 2015, 12:00 PM -12:15 PM
David A. Cheresh. UCSD Moores Cancer Center, La Jolla, CA

SY27-04  
 
Presentation Title: Induction of cancer stemness and drug resistance by EGFR blockade
Presentation Time: Tuesday, Apr 21, 2015, 12:00 PM -12:15 PM
Abstract Body: Tumor drug resistance is often accompanied by genetic and biological changes in the tumor cell population reflecting the acquisition of a stem-like state. However, it is not clear whether cancer therapies select for the growth of drug resistance cancer stem cells and/or directly induce the reprograming of tumor cells to a cancer stem-like, drug resistance state. We provide evidence that breast, pancreas and lung carcinomas in the presence of prolonged exposure to EGFR inhibitors undergo an epigenetic reprogramming resulting in a drug resistant stem-like tumor population expressing the cell surface marker CD61 (b3 integrin). In fact, CD61 in the context of KRAS, is necessary and sufficient to account for drug resistance, tumor initiation, self-renewal and expression of the pluripotent genes Oct 4 and Nanog. Once expressed, CD61 in the unligated state recruits KRAS to the plasma membrane leading to the activation of RalB, TBK1 and c-Rel driving both stemness and EGFR inhibitor resistance. Pharmacological targeting this pathway with drugs such as bortezomib or revlimid not only reverses stemness but resensitizes these epithelial tumors to EGFR inhibition. This epigenetic pathway can also be initiated by range of cellular stresses found within the tumor microenvironment such as hypoxia, nutrient deprivation, low pH, and oxidative stress. In normal tissues CD61 is induced during tissue remodeling and repair. For example, CD61 was found to be critical for mammary gland remodeling during pregnancy and as a mediator of pathological neovascularization. Together these findings reveal a stress-induced epigenetic pathway characterized by the upregulation of CD61 that promotes the remodeling of normal tissues but in tumors contributes to EGFR inhibitor resistance and tumor progression.

 

http://cancerres.aacrjournals.org/gca?gca=canres%3B75%2F15_Supplement%2F4&gca=canres%3B75%2F15_Supplement%2F6&gca=canres%3B75%2F15_Supplement%2F19&gca=canres%3B75%2F15_Supplement%2F24&gca=canres%3B75%2F15_Supplement%2F48&gca=canres%3B75%2F15_Supplement%2F54&gca=canres%3B75%2F15_Supplement%2F57&gca=canres%3B75%2F15_Supplement%2F88&gca=canres%3B75%2F15_Supplement%2F90&gca=canres%3B75%2F15_Supplement%2F97&allch=&submit=Go

Selected Abstracts

  1. Abstract 1
  2. Molecular and Cellular Biology – Poster Presentations – Proffered Abstracts – Poster Presentations – Cell Death Mechanisms: Abstract 4: ABT-263 is effective in a subset of non-small cell lung cancer cell lines
    • Aoi Kuroda,
    • Keiko Ohgino,
    • Hiroyuki Yasuda,
    • Junko Hamamoto,
    • Daisuke Arai,
    • Kota Ishioka,
    • Tetsuo Tani,
    • Shigenari Nukaga,
    • Ichiro Kawada,
    • Katsuhiko Naoki,
    • Kenzo Soejima,
    • and Tomoko Betsuyaku

Cancer Res August 1, 2015 75:4; doi:10.1158/1538-7445.AM2015-4

  1. Abstract 2
  2. Molecular and Cellular Biology – Poster Presentations – Proffered Abstracts – Poster Presentations – Cell Death Mechanisms: Abstract 6: Quantitative assessment of BCL-2:BIM complexes as a pharmacodynamic marker for venetoclax (ABT-199)
    • Sha Jin,
    • Paul Tapang,
    • Donald J. Osterling,
    • Wenqing Gao,
    • Daniel H. Albert,
    • Andrew J. Souers,
    • Joel D. Leverson,
    • Darren C. Phillips,
    • and Jun Chen

Cancer Res August 1, 2015 75:6; doi:10.1158/1538-7445.AM2015-6

  1. Molecular and Cellular Biology – Poster Presentations – Proffered Abstracts – Poster Presentations – Cell Death Mechanisms: Abstract 24: The phosphorylation of p53 at serine 46 is essential to induce cell death through palmdelphin in response to DNA damage
    • Nurmaa Khund Dashzeveg and
    • Kiyotsugu Yoshida

Cancer Res August 1, 2015 75:24; doi:10.1158/1538-7445.AM2015-24

  1. Abstract 5
  2. Molecular and Cellular Biology – Poster Presentations – Proffered Abstracts – Poster Presentations – Cell Signaling in Cancer 1: Abstract 48: Identification of a novel binding protein playing a critical role in HER2 activation in lung cancer cells
    • Tomoaki Ohtsuka,
    • Masakiyo Sakaguchi,
    • Katsuyoshi Takata,
    • Shinsuke Hashida,
    • Mototsugu Watanabe,
    • Ken Suzawa,
    • Yuho Maki,
    • Hiromasa Yamamoto,
    • Junichi Soh,
    • Hiroaki Asano,
    • Kazunori Tsukuda,
    • Shinichiro Miyoshi,
    • and Shinichi Toyooka

Cancer Res August 1, 2015 75:48; doi:10.1158/1538-7445.AM2015-48

  1. Abstract 1 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Death Mechanisms

Abstract 4: ABT-263 is effective in a subset of non-small cell lung cancer cell lines

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

Rationale:

ABT-263 (Navitoclax) is one of the BH3 mimetics targeting anti-apoptotic B-cell lymphoma-2 (Bcl-2) family proteins such as Bcl-2, Bcl-XL, and Bcl-w, thereby inducing apoptosis. It has been reported that the response to ABT-263 is associated with expressions of myeloid cell leukemia-1 (Mcl-1), an anti-apoptotic protein. Given its effectiveness as a single agent in preclinical studies, ABT-263 is currently being evaluated in clinical trials for small cell lung cancer (SCLC) and leukemia. However, the efficacy of ABT-263 in non-small cell lung cancer (NSCLC) has not been fully evaluated. We examined the effect of ABT-263 on cell proliferation of NSCLC cell lines and investigated the underlying mechanisms.

Methods:

The following 9 NSCLC cell lines were examined: SK-LU-1, A549, H358, Calu3, H3122, H1975, H460, H441, and BID007. The effects of ABT-263 in NSCLC cell lines were evaluated by MTS assay. Apoptosis was examined by flowcytometry using staining for annexin V and propidium iodide (PI), and also western blotting for cleaved PARP. Quantitative RT-PCR was carried out to assess the mRNA expression levels of anti-apoptotic genes and pro-apoptotic genes. Immunoprecipitation and western blotting were performed to compare the levels of anti-apoptotic and pro-apoptotic proteins between the sensitive and resistant cell lines. In addition, knockdown of Mcl-1 was performed by siRNA.

Results:

By screening 9 NSCLC cell lines using MTS assay, we found Calu3 and BID007were sensitive to ABT-263. We also confirmed that apoptosis was induced only in the ABT-263 sensitive lines but not in the ABT-263 resistant cell lines after ABT-263 treatment. However, the expression levels of Bcl-2 family proteins, including Mcl-1, did not differ significantly among the ABT-263 sensitive and resistant cell lines. Unlike the results in previous reports regarding SCLC, Mcl-1 was not decreased in the sensitive cell lines. The ABT-263 resistant cell lines became sensitive to ABT-263 after knockdown of Mcl-1 by siRNA, while the ABT-263 sensitive cell lines maintained the same sensitivity.

Conclusion:

We found that Calu3 and BID007 were sensitive to ABT-263. In the sensitive NSCLC cell lines, ABT-263 induces apoptosis irrespective of Mcl-1 expression levels.

Citation Format: Aoi Kuroda, Keiko Ohgino, Hiroyuki Yasuda, Junko Hamamoto, Daisuke Arai, Kota Ishioka, Tetsuo Tani, Shigenari Nukaga, Ichiro Kawada, Katsuhiko Naoki, Kenzo Soejima, Tomoko Betsuyaku. ABT-263 is effective in a subset of non-small cell lung cancer cell lines. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 4. doi:10.1158/1538-7445.AM2015-4

    • ©2015 American Association for Cancer Research.
  1. Abstract 2 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Death Mechanisms

Abstract 6: Quantitative assessment of BCL-2:BIM complexes as a pharmacodynamic marker for venetoclax (ABT-199)

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

The BCL-2-selective inhibitor venetoclax (ABT-199) binds with high affinity to the BH3-binding groove of BCL-2, thereby competing for binding with the BH3-only protein BIM (Souers et al., 2013). Venetoclax is currently being evaluated in clinical trials for CLL, AML, multiple myeloma and NHL. To facilitate these studies, we developed and validated a 384-well electrochemiluminescent ELISA (MSD, Gaithersburg, MD,USA) that quantifies expression of BCL-2, BCL-XL, and MCL-1protein alone or in complex with BIM. We subsequently quantified expression of BCL-2 and BCL-2:BIM complexes in 16 hematologic tumor cell lines. We found the EC50 of venetoclax in these tumor cell lines to correlate strongly with baseline BCL-2:BIM complex levels. This correlation was superior to the correlation between venetoclax EC50 and absolute BCL-2 expression. We also applied the assay to measure disruption of BCL-2:BIM complexes in vivo. Treatment of the Non-Hodgkin’s Lymphoma (NHL) xenograft model SU-DHL-4 with a BCL-2-selective inhibitor resulted in disruption of tumor BCL-2:BIM complexes that aligned with serum and tumor concentrations of inhibitor. Collectively, these data demonstrate that quantifying BCL-2:BIM complexes offers an accurate means of assessing target engagement by venetoclax and, potentially, predicting its efficacy. The utility of this assay is currently being assessed in clinical trials.

Citation Format: Sha Jin, Paul Tapang, Donald J. Osterling, Wenqing Gao, Daniel H. Albert, Andrew J. Souers, Joel D. Leverson, Darren C. Phillips, Jun Chen. Quantitative assessment of BCL-2:BIM complexes as a pharmacodynamic marker for venetoclax (ABT-199). [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 6. doi:10.1158/1538-7445.AM2015-6

    • ©2015 American Association for Cancer Research.
  1. Abstract 3 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Death Mechanisms

Abstract 19: Antitumor activity of selective inhibitors of XPO1/CRM1-mediated nuclear export in diffuse malignant peritoneal mesothelioma: the role of survivin

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

Survivin, which is highly expressed and promotes cell survival in diffuse malignant peritoneal mesothelioma (DMPM), exclusively relies on the nuclear exportin 1 (XPO1/CRM1) to be released in the cytoplasm and perform its anti-apoptotic function. Here, we explored the efficacy of selective inhibitors of nuclear export (SINEs) in patient-derived DMPM preclinical models. Exposure to individual SINE (KPT-251, KPT-276, KPT-330) was able to induce a time- and dose-dependent inhibition of the growth of two DMPM cell lines without affecting normal cell proliferation. Such a cell growth inhibition was preceded by a decline in the nuclear XPO1/CRM1 levels and an increase in the nuclear accumulation of its cargo proteins p53 and p21, which led to a cell cycle arrest at G1-phase. Our results also indicated that survivin is an essential component of the downstream signaling pathway of XPO1/CRM1 inhibition in DMPM cells. In fact, in both cell lines, exposure to SINEs led to a time-dependent reduction of cytoplasmic survivin levels and, after an initial survivin nuclear accumulation, also to a progressive decrease in the nuclear protein abundance, through the ubiquitin-proteasomal degradation pathway, leading to the complete depletion of total survivin levels. In both DMPM cell models, according to survivin anti-apoptotic activity, drug-induced reduction of cytoplasmic survivin levels correlated with the onset of caspase-dependent apoptosis. We further observed that SINEs can be combined with other survivin inhibitors, such as the survivin suppressant YM155 to achieve enhanced growth inhibition in DMPM cells. Initial in vivo experiments with orally administered KPT-251, KPT-276 and the orally available, clinical stage KPT-330 (selinexor) indicated that each compound was able to significantly reduce the growth of early-stage subcutaneous DMPM xenografts. Interestingly, additional experiments carry out with selinexor demonstrated that the compound was also able to inhibit the growth of late-stage subcutaneous DMPM xenografts in nude mice. Most importantly, oral administration of selinexor to SCID mice reduced the growth of orthotopic DMPM xenografts, which properly recapitulate the dissemination pattern in the peritoneal cavity of human DMPM and, for this reason, represent a valuable model for investigating novel therapeutic approaches for the disease. Consistent with an important role of survivin as a determinant of anti-cancer activity of SINE compounds, a reduction of the protein expression was observed in tumor specimens obtained from selinexor treated mice. Overall, our results (i) demonstrate a marked efficacy of SINEs in DMPM preclinical models, which is, at least in part, dependent on the interference with survivin intracellular distribution and function, and (ii) suggest SINE-mediated XPO1/CRM1 inhibition as a novel therapeutic option for the disease.

Citation Format: Nadia Zaffaroni, Michelandrea De Cesare, Denis Cominetti, Valentina Doldi, Alessia Lopergolo, Marcello Deraco, Paolo Gandellini, Yosef Landesman, Sharon Friedlander, Michael G. Kauffman, Sharon Shacham, Marzia Pennati. Antitumor activity of selective inhibitors of XPO1/CRM1-mediated nuclear export in diffuse malignant peritoneal mesothelioma: the role of survivin. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 19. doi:10.1158/1538-7445.AM2015-19

    • ©2015 American Association for Cancer Research.
  1. Abstract 4 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Death Mechanisms

Abstract 24: The phosphorylation of p53 at serine 46 is essential to induce cell death through palmdelphin in response to DNA damage

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

Tumor suppressor p53 plays a pivotal role in cell cycle arrest, DNA repair, and apoptosis in response to DNA damage. Promoter selectivity of p53 depends mainly on post-translational modification. Notably, the apoptotic function of p53 is related to its phosphorylation at serine-46 (ser46) to promote pro-apoptotic genes. However, little is known about the pro-apoptotic genes induced by Ser46 phosphorylation. Our research achieved to investigate the pro-apoptotic genes induced by p53 in a phospho-ser46-specific manner using microarray and ChIP sequencing in human cancer cell lines. As a result, palmdelphin (PALMD), an isoform of paralemmin protein, was strongly elicited from the phosphorylation of ser46. The mRNA and protein expression of PALMD increased only in wild type p53 transfected cells, but not in ser46-mutated cells. Importantly, PALMD moved to the nucleus in response to DNA damage and the apoptotic function of PALMD was tightly exerted with localization into nucleus. Interestingly, down-regulation of PALMD by siRNA resulted in necroptosis-like cell death through ATP depletion. Moreover, we found vimentin as a PALMD interacting protein and the depletion of vimentin increased PALMD level to accelerate apoptosis. These results demonstrate that p53 regulates cell death fate (apoptosis or necroptosis-like cell death) through promoting PALMD expression in a phospho-ser46-specific manner in response to DNA damage.

Citation Format: Nurmaa Khund Dashzeveg, Kiyotsugu Yoshida. The phosphorylation of p53 at serine 46 is essential to induce cell death through palmdelphin in response to DNA damage. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 24. doi:10.1158/1538-7445.AM2015-24

    • ©2015 American Association for Cancer Research.
  1. Abstract 5 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Signaling in Cancer 1

Abstract 48: Identification of a novel binding protein playing a critical role in HER2 activation in lung cancer cells

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

Human epidermal growth factor receptor 2 (HER2) is a member of epidermal growth factor receptor (EGFR) family. Previous studies have revealed that many kinds of malignant tumors have genetic mutations or amplification of HER2, indicating that HER2 alterations are oncogenic. Many kinds of HER2 targeted therapies are effective to HER2 positive tumors, but those treated tumors often get resistance to drugs. Thus, to elucidate HER2 related pathway in cancer biology is important to develop new therapeutic strategy for cancers.

Recently, we newly identified a protein X (a temporary name) as a novel binding protein to HER2 with immunoprecipitation and following LC-Ms/Ms analysis. The protein generally expressed in lung and breast cancers at remarkable level.

We constructed plasmid vectors carrying wild type HER2 and gene X. These vectors were simultaneously introduced to HEK293T cells to examine the binding ability of protein X and HER2 as well as the effect of gene X on HER2-mediated signal-transduction pathway. The approach clearly showed that the expression of gene X, resulted in phosphorylation of HER2 and subsequent activation of oncogenic effector molecules.

We next constructed several kinds of gene X-truncated variants and subjected to the binding assay to look for the binding domain of gene X to HER2. The analysis showed that N-terminal head domain of gene X was essential for the HER2 binding. This domain has an ability to induce HER2 phosphorylation and subsequent activation of the effector kinase, ERK.

In conclusion, we found that gene X is a novel binding protein to HER2 and has a role in HER2 activation.

Citation Format: Tomoaki Ohtsuka, Masakiyo Sakaguchi, Katsuyoshi Takata, Shinsuke Hashida, Mototsugu Watanabe, Ken Suzawa, Yuho Maki, Hiromasa Yamamoto, Junichi Soh, Hiroaki Asano, Kazunori Tsukuda, Shinichiro Miyoshi, Shinichi Toyooka. Identification of a novel binding protein playing a critical role in HER2 activation in lung cancer cells. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 48. doi:10.1158/1538-7445.AM2015-48

    • ©2015 American Association for Cancer Research.
  1. Abstract 6 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Signaling in Cancer 1

Abstract 54: Ezrin enhances signaling and nuclear translocation of the epidermal growth factor receptor in non-small cell lung cancer cells

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

The cytoskeletal cross linker protein ezrin is a member of the ezrin-radixin-moesin (ERM) family and plays important roles not only in cell motility, cell adhesion, and apoptosis, but also in various cell-signaling pathways. Ezrin interacts with EGFR in the cell membrane and involves in cell motility events, but little is known about the effects of this interaction on the EGFR signaling pathway. We investigated the role of Ezrin in EGFR signaling and nuclear trafficking in non-small cell lung cancer (NSCLC) cell lines. The ligand induced interaction between Ezrin and EGFR was evaluated by immunoprecipitation (IP) and immunofluorescence (IF) in H292 and A549 cells. Ezrin levels were reduced using siRNA in these two cell lines. Downstream signaling protein phosphorylation and nuclear localization of EGFR were detected after EGF treatment. Expressions of nuclear EGFR target genes were evaluated by qPCR. Endogenous Ezrin was found in a complex with EGFR in IP and IF. When Ezrin protein expression was inhibited, phosphorylation levels of EGFR at Y1068, Y1101 and Y845 were reduced as well as phosphorylation levels of downstream signaling pathway proteins ERK and STAT3. Cell fractionation revealed that EGFR nuclear translocation after EGF treatment significantly reduced in Ezrin-knockdown cells. Further, mRNA levels of EGFR target genes AuroraK-A, COX2, Cyclin D1 and iNOS were decreased in Ezrin-knockdown A549 cells. Small molecule ezrin inhibitors showed strong synergy with EGFR inhibitors in cytotoxicity assays. These results suggest that Ezrin has a role as an enhancer in the EGFR pathway and targeting ezrin may potentiate anti-EGFR based therapies in NSCLC.

Citation Format: Yasemin Saygideger Kont, Haydar Celik, Hayriye V. Erkizan, Tsion Minas, Jenny Han, Jeffrey Toretsky, Aykut Uren. Ezrin enhances signaling and nuclear translocation of the epidermal growth factor receptor in non-small cell lung cancer cells. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 54. doi:10.1158/1538-7445.AM2015-54

    • ©2015 American Association for Cancer Research.
  1. Abstract 7 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Cell Signaling in Cancer 1

Abstract 57: Substrates of protein kinase C drive cell rac1-dependent motility

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

This laboratory has identified and/or characterized substrates of PKC that upon phosphorylation give rise to motility, an aspect of metastasis. By use of the traceable kinase method, we discovered that alpha-tubulin and Cdc42 effector protein-4 (CEP4) are PKC substrates. Phosphorylation of alpha-tubulin stimulates its incorporation into microtubules (MTs), consequently increasing the stability and prolonged growth of MTs and leading to the activation of the small GTPase Rac1. CEP4 undergoes phosphorylation by PKC that results in its release from Cdc42, whereupon CEP4 binds a guanine nucleotide exchange factor (GEF) that in turn activates Rac1 GTPase. These results imply that Rac1 acts as a node in pathways driven by phosphorylated PKC substrates. Since translocation of IQGAP to the membrane is known to be promoted by Rac1, a role is explored in non-transformed human MCF-10A cells that express a specific phospho-mimetic mutant substrate. In addition, the phospho-mimetic mutant for each substrate expressed in human metastatic MDA-MB-231 cells produces different morphologies in 3-D growth assays. This research is being supported by NIH CA125632.

Citation Format: Susan A. Rotenberg, Xin Zhao, Shatarupa De. Substrates of protein kinase C drive cell rac1-dependent motility. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 57. doi:10.1158/1538-7445.AM2015-57

    • ©2015 American Association for Cancer Research.
  1. Abstract 8 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Deregulation of Gene Expression in Prostate Cancer and Sarcoma

Abstract 88: The Nkx3.1 homeobox gene maintains prostatic identity while its loss leads to prostate cancer initiation

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

Background

Maintenance of epithelial cell identity is tightly coordinated by tissue-specific gene expression programs, which are often deregulated during tumorigenesis. The homeodomain-containing transcription factor, Nkx3.1, is a key regulator of normal prostatic development and is frequently lost at early stages of prostate cancer initiation. In this study, we aim to elucidate detailed mechanisms governing Nkx3.1-driven maintenance of prostate identity and how deregulation of such can lead to prostate tumorigenesis.

Models and Methods

We evaluated the consequences of Nkx3.1 loss or gain of function in vivo using genetically-engineered mouse models and cell-recombination assays. RNA sequencing was performed to generate gene expression profiles, which were analyzed using Gene Set Enrichment analysis (GSEA), and validated by quantitative real-time PCR. In parallel, protein expression was assessed by immunofluorescence and western blot. Immunoprecipitation (IP) and chromatin-immunoprecipitation (ChIP) assays were performed using RWPE1 prostate epithelial cells.

Results

Here, we show that loss of function of Nkx3.1 leads to the progressive down-regulation of a prostate-specific gene expression program and to aberrant expression of genes that are not typically expressed in the prostate epithelium. Conversely, gain of function of Nkx3.1 in non-prostatic epithelium leads to the acquisition of a prostate-like morphology and expression of prostate-related genes. Our findings indicate that the underlying mechanism by which Nkx3.1 promotes prostatic identity is via epigenetic regulation of gene expression. In particular, we show that Nkx3.1 interacts with the histone methyl-transferase complex G9a/Glp. Finally, we demonstrate that this interaction is necessary for maintenance of prostate identity in vivo and that Nkx3.1 and G9a cooperate to control expression of genes that coordinate prostatic epithelial integrity.

Conclusions

Our results suggest that Nkx3.1 promotes prostatic identity by interacting with histone modifying enzymes to coordinate the expression of prostate-specific genes and that the loss of this function results in a failure to maintain prostate identity associated with early stages of prostate tumorigenesis.

Citation Format: Clémentine Le Magnen, Aditya Dutta, Cory Abate-Shen. The Nkx3.1 homeobox gene maintains prostatic identity while its loss leads to prostate cancer initiation. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 88. doi:10.1158/1538-7445.AM2015-88

    • ©2015 American Association for Cancer Research.
  1. Abstract 9 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Deregulation of Gene Expression in Prostate Cancer and Sarcoma

Abstract 90: K63-linked JARID1B ubiquitination by TRAF6 contributes to aberrant elevation of JARID1B in prostate cancer

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

Aberrant elevation of JARID1B and histone H3 Lys4 trimethylations (H3K4me3) is frequently observed in many diseases including prostate cancer (PCa), yet the mechanisms on the regulations of JARID1B and H3K4me3 through epigenetic modifications still remain poorly understood. In this study we performed immunohistochemistry staining, immunofluorescence imaging, immunoprecipitation, shRNA and Western blotting analysis in mouse embryonic fibroblasts (MEFs), mouse models, and cultured human prostate cancer cells. As a result, we discovered that SKP2 modulates JARID1B and H3K4me3 levels in vitro in PTEN null prostate cancer cells and in vivo in Pten/Trp53 mouse models. We demonstrated that levels of SKP2, JARID1B and H3K4me3 are strikingly elevated in vitro and in vivo when both PTEN and P53 are inactivated. Importantly, SKP2 inactivation resulted in a reduction of cell growth, cell migration and malignant transformation of Pten/Trp53 double null MEFs, and further restrained prostate tumorigenesis of Pten/Trp53 mutant mice. Mechanistically, JARID1B is ubiquitinated by E3 ligase TRAF6 through the K63-linkage in prostate cancer cells. Interestingly, SKP2 contributes to JARID1B ubiquitination machinery as a non-E3 ligase regulator by decreasing TRAF6-mediated ubiquitination of JARID1B. SKP2 deficiency resulted in an increase of JARID1B ubiquitination and in turn a reduction of H3K4me3, and induced senescence through JARID1B accumulation in nucleoli of PCa cells and prostate tumors of mice. Furthermore, we showed that the aberrant levels of SKP2, JARID1B, and H3K4me3 are associated with malignant features of castration-resistant prostate cancer (CRPC) in mice. Overall, our findings reveal a novel network of SKP2- JARID1B, and targeting SKP2 and JARID1B may be a potential strategy for PCa control.

Citation Format: Wenfu Lu, Shenji Liu, Bo Li, Yingqiu Xie, Christine Adhiambo, Qing Yang, Billy R. Ballard, Keiichi I. Nakayama, Robert J. Matusik, Zhenbang Chen. K63-linked JARID1B ubiquitination by TRAF6 contributes to aberrant elevation of JARID1B in prostate cancer. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 90. doi:10.1158/1538-7445.AM2015-90

    • ©2015 American Association for Cancer Research.
  1. Abstract 10 of 10Molecular and Cellular Biology / Poster Presentations – Proffered Abstracts / Poster Presentations – Histone Methylation and Acetylation

Abstract 97: CARM1 preferentially methylates H3R17 over H3R26 through a random kinetic mechanism

Proceedings: AACR 106th Annual Meeting 2015; April 18-22, 2015; Philadelphia, PA

CARM1 (PRMT4) is a type I arginine methyltransferase involved in the regulation of transcription, pre-mRNA splicing, cell cycle progression and the DNA damage response. Overexpression of CARM1 has been implicated in breast, prostate, and colorectal cancers. Since CARM1 appears to be a good target for the development of therapies against these cancers, we studied the substrate specificity and kinetic mechanism of the full-length human enzyme. CARM1 has been shown to methylate both residues R17 and R26 of histone H3. Substrate specificity was examined by testing CARM1 activity with several H3-based peptide substrates using a radiometric assay. Comparison of kcat/KM values reveal that methylation of H3R17 is preferred over H3R26. An R17/R26K peptide produced 8-fold greater kcat/KM value compared to the corresponding R17K/R26 peptide. These effects are KM-driven as kcat values remain relatively constant for the peptides tested. Shortening the peptide at the C-terminus by 5 amino acid residues greatly reduced the specificity (16-24-fold), demonstrating the contribution of distal residues to substrate binding. In contrast, adding residues to the N-terminus of the shortened peptide had a negative effect on activity. CARM1 displays little preference for monomethylated over unmethylated H3R17 (2-5-fold by kcat/KM) suggesting that it operates through a distributive mechanism. Previous crystallographic studies with mouse CARM1 showed that part of the substrate binding groove was formed by cofactor binding, thereby suggesting an ordered kinetic mechanism (Yue et al., EMBO J., 2007). Our results from dead-end and product inhibition studies performed with human CARM1, however, are consistent with a random kinetic mechanism. SAH and sinefungin demonstrate competitive inhibition with respect to SAM and produced noncompetitive inhibition patterns with respect to peptide. Both dimethylated R17 product peptide and dead-end R17K peptide exhibited noncompetitive inhibition patterns with respect to SAM. Furthermore, binding of SAM and peptide substrates were shown to be independent of each other in initial velocity experiments where both substrates were varied. Together, these results elucidate the kinetic mechanism of CARM1 and highlight elements important for binding affinity.

Citation Format: Suzanne L. Jacques, Katrina P. Aquino, Jodi Gureasko, P Ann Boriack-Sjodin, Robert A. Copeland, Thomas V. Riera. CARM1 preferentially methylates H3R17 over H3R26 through a random kinetic mechanism. [abstract]. In: Proceedings of the 106th Annual Meeting of the American Association for Cancer Research; 2015 Apr 18-22; Philadelphia, PA. Philadelphia (PA): AACR; Cancer Res 2015;75(15 Suppl):Abstract nr 97. doi:10.1158/1538-7445.AM2015-97

    • ©2015 American Association for Cancer Research.

References

 

  1. Bonnet D, Dick JE: Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med 1997, 3(7):730-737.
  2. Al-Hajj M, Clarke MF: Self-renewal and solid tumor stem cells. Oncogene 2004, 23(43):7274-7282.
  3. Hughes L, Malone C, Chumsri S, Burger AM, McDonnell S: Characterisation of breast cancer cell lines and establishment of a novel isogenic subclone to study migration, invasion and tumourigenicity. Clin Exp Metastasis 2008, 25(5):549-557.
  4. Li C, Lee CJ, Simeone DM: Identification of human pancreatic cancer stem cells. Methods Mol Biol 2009, 568:161-173.
  5. Zhang S, Balch C, Chan MW, Lai HC, Matei D, Schilder JM, Yan PS, Huang TH, Nephew KP: Identification and characterization of ovarian cancer-initiating cells from primary human tumors. Cancer Res 2008, 68(11):4311-4320.
  6. Kakarala M, Wicha MS: Implications of the cancer stem-cell hypothesis for breast cancer prevention and therapy. J Clin Oncol 2008, 26(17):2813-2820.
  7. Ginestier C, Hur MH, Charafe-Jauffret E, Monville F, Dutcher J, Brown M, Jacquemier J, Viens P, Kleer CG, Liu S et al: ALDH1 is a marker of normal and malignant human mammary stem cells and a predictor of poor clinical outcome. Cell Stem Cell 2007, 1(5):555-567.
  8. Dontu G: Breast cancer stem cell markers – the rocky road to clinical applications. Breast Cancer Res 2008, 10(5):110.
  9. Ferrandina G, Bonanno G, Pierelli L, Perillo A, Procoli A, Mariotti A, Corallo M, Martinelli E, Rutella S, Paglia A et al: Expression of CD133-1 and CD133-2 in ovarian cancer. Int J Gynecol Cancer 2008, 18(3):506-514.

 

Additional Articles on this Open Access Journal on Cancer Stem Cells Include

Nonhematologic Cancer Stem Cells [11.2.3]

In Focus: Identity of Cancer Stem Cells

In Focus: Targeting of Cancer Stem Cells

Stem Cells and Cancer

Positron Emission Tomography (PET) and Near-Infrared Fluorescence Imaging: Noninvasive Imaging of Cancer Stem Cells (CSCs) monitoring of AC133+ glioblastoma in subcutaneous and intracerebral xenograft tumors

“To Die or Not To Die” – Time and Order of Combination drugs for Triple Negative Breast Cancer cells: A Systems Level Analysis

Can IntraTumoral Heterogeneity Be Thought of as a Mechanism of Resistance?

 

 

Read Full Post »

Are Cyclin D and cdk Inhibitors A Good Target for Chemotherapy?

 

Curator: Stephen J. Williams, Ph.D.

UPDATED 7/12/2022

see below for great review

 

 

CDK4 and CDK6 kinases: From basic science to cancer therapy

SCIENCE
14 Jan 2022
Vol 375Issue 6577

Targeting cyclin-dependent kinases

Cyclin-dependent kinases (CDKs), in complex with their cyclin partners, modulate the transition through phases of the cell division cycle. Cyclin D–CDK complexes are important in cancer progression, especially for certain types of breast cancer. Fassl et al. discuss advances in understanding the biology of cyclin D–CDK complexes that have led to new concepts about how drugs that target these complexes induce cancer cell cytostasis and suggest possible combinations to widen the types of cancer that can be treated. They also discuss progress in overcoming resistance to cyclin D–CDK inhibitors and their possible application to diseases beyond cancer. —GKA

Structured Abstract

BACKGROUND

Cyclins and cyclin-dependent kinases (CDKs) drive cell division. Of particular importance to the cancer field are D-cyclins, which activate CDK4 and CDK6. In normal cells, the activity of cyclin D–CDK4/6 is controlled by the extracellular pro-proliferative or inhibitory signals. By contrast, in many cancers, cyclin D–CDK4/6 kinases are hyperactivated and become independent of mitogenic stimulation, thereby driving uncontrolled tumor cell proliferation. Mouse genetic experiments established that cyclin D–CDK4/6 kinases are essential for growth of many tumor types, and they represent potential therapeutic targets. Genetic and cell culture studies documented the dependence of breast cancer cells on CDK4/6. Chemical CDK4/6 inhibitors were synthesized and tested in preclinical studies. Introduction of these compounds to the clinic represented a breakthrough in breast cancer treatment and will likely have a major impact on the treatment of many other tumor types.

ADVANCES

Small-molecule CDK4/6 inhibitors (palbociclib, ribociclib, abemaciclib) showed impressive results in clinical trials for patients with hormone receptor–positive breast cancers. Addition of CDK4/6 inhibitors to standard endocrine therapy substantially extended median progression-free survival and prolonged median overall survival. Consequently, all three CDK4/6 inhibitors have been approved for treatment of women with advanced or metastatic hormone receptor–positive breast cancers. In the past few years, the renewed interest in CDK4/6 biology has yielded several surprising discoveries. The emerging concept is that CDK4/6 kinases regulate a much wider set of cellular functions than anticipated. Consequently, CDK4/6 inhibitors, beyond inhibiting tumor cell proliferation, affect tumor cells and the tumor environment through mechanisms that are only beginning to be elucidated. For example, inhibition of CDK4/6 affects antitumor immunity acting both on tumor cells and on the host immune system. CDK4/6 inhibitors were shown to enhance the efficacy of immune checkpoint blockade in preclinical mouse cancer models. These new concepts are now being tested in clinical trials.

OUTLOOK

Palbociclib, ribociclib, and abemaciclib are being tested in more than 300 clinical trials for more than 50 tumor types. These trials evaluate CDK4/6 inhibitors in combination with a wide range of therapeutic compounds that target other cancer-relevant pathways. Several other combination treatments were shown to be efficacious in preclinical studies and will enter clinical trials soon. Another CDK4/6 inhibitor, trilaciclib, is being tested for its ability to shield normal cells of the host from cytotoxic effects of chemotherapy. New CDK4/6 inhibitors have been developed and are being assessed in preclinical and clinical trials. The major impediment in the therapeutic use of CDK4/6 inhibitors is that patients who initially respond to treatment often develop resistance and eventually succumb to the disease. Moreover, a substantial fraction of tumors show preexisting, intrinsic resistance to CDK4/6 inhibitors. One of the main challenges will be to elucidate the full range of resistance mechanisms. Even with the current, limited knowledge, one can envisage the principles of new, improved approaches to overcome known resistance mechanisms. Another largely unexplored area for future study is the possible involvement of CDK4/6 in other pathologic states beyond cancer. This will be the subject of intense studies, and it may extend the utility of CDK4/6 inhibitors to the treatment of other diseases.
Targeting cyclin D–CDK4/6 for cancer treatment.
D-cyclins (CycD) activate CDK4 and CDK6 in G1 phase of the cell cycle and promote cell cycle progression by phosphorylating the retinoblastoma protein RB1. RB1 inhibits E2F transcription factors; phosphorylation of RB1 activates E2F-driven transcription. In many cancers, CycD-CDK4/6 is constitutively activated and drives uncontrolled cell proliferation. The development of small-molecule CDK4/6 inhibitors provided a therapeutic tool to repress constitutive CycD-CDK4/6 activity and to inhibit cancer cell proliferation. As with several targeted therapies, tumors eventually develop resistance and resume cell proliferation despite CDK4/6 inhibition. New combination treatments, involving CDK4/6 inhibitors plus inhibition of other pathways, are being tested in the clinic to delay or overcome the resistance.
OPEN IN VIEWER

 

Abstract

Cyclin-dependent kinases 4 and 6 (CDK4 and CDK6) and their activating partners, D-type cyclins, link the extracellular environment with the core cell cycle machinery. Constitutive activation of cyclin D–CDK4/6 represents the driving force of tumorigenesis in several cancer types. Small-molecule inhibitors of CDK4/6 have been used with great success in the treatment of hormone receptor–positive breast cancers and are in clinical trials for many other tumor types. Unexpectedly, recent work indicates that inhibition of CDK4/6 affects a wide range of cellular functions such as tumor cell metabolism and antitumor immunity. We discuss how recent advances in understanding CDK4/6 biology are opening new avenues for the future use of cyclin D–CDK4/6 inhibitors in cancer treatment.
Cyclin D1, the activator of CDK4 and CDK6, was discovered in the early 1990s (12). The role of cyclin D1 in oncogenesis was already evident at the time of its cloning, as it was also identified as the protein product of the PRAD1 oncogene, which is rearranged and overexpressed in parathyroid adenomas (3), and of the BCL1 oncogene, which is rearranged in B-lymphocytic malignancies (4). Subsequently, the remaining two D-type cyclins, D2 and D3, were discovered on the basis of their homology to cyclin D1 (1).
Cyclins serve as regulatory subunits of cyclin-dependent kinases (CDKs) (5). Shortly after the discovery of D-cyclins, CDK4 and CDK6 were identified as their kinase partners (6). Mouse gene knockout studies revealed that CDK4 and CDK6 play redundant roles in development, and combined ablation of CDK4 and CDK6 was found to result in embryonic lethality (7). The essentially identical phenotype was seen in cyclin D–knockout mice, thereby confirming the role of D-cyclins as CDK4/6 activators in vivo (8). Surprisingly, these analyses revealed that many normal nontransformed mammalian cell types can proliferate without any cyclin D–CDK4/6 activity (78).
CDK4 and CDK6 are expressed at constant levels throughout the cell cycle. By contrast, D-cyclins are labile proteins that are transcriptionally induced upon stimulation of cells with growth factors. For this reason, D-cyclins are regarded as links between the cellular environment and the cell cycle machinery (6).
Cell cycle inhibitors play an important role in regulating the activity of cyclin D–CDK4/6 (Fig. 1). The INK inhibitors (p16INK4A, p15INK4B, p18INK4C, p19INK4D) bind to CDK4 or CDK6 and prevent their interaction with D-type cyclins, thereby inhibiting cyclin D–CDK4/6 kinase activity. By contrast, KIP/CIP inhibitors (p27KIP1, p57KIP2, p21CIP1), which inhibit the activity of CDK2-containing complexes, serve as assembly factors for cyclin D–CDK4/6 (69). This was demonstrated by the observation that mouse fibroblasts devoid of p27KIP1 and p21CIP1 fail to assemble cyclin D–CDK4/6 complexes (10).
Fig. 1. Molecular events governing progression through the G1 phase of the cell cycle.
The mammalian cell cycle can be divided into G1, S (DNA synthesis), G2, and M (mitosis) phases. During G1 phase, cyclin D (CycD)–CDK4/6 kinases together with cyclin E (CycE)–CDK2 phosphorylate the retinoblastoma protein RB1. This activates the E2F transcriptional program and allows entry of cells into S phase. Members of the INK family of inhibitors (p16INK4A, p15INK4B, p18INK4C, and p19INK4D) inhibit cyclin D–CDK4/6; KIP/CIP proteins (p21CIP1, p27KIP1, and p57KIP2) inhibit cyclin E–CDK2. Cyclin D–CDK4/6 complexes use p27KIP1 and p21CIP1 as “assembly factors” and sequester them away from cyclin E–CDK2, thereby activating CDK2. Proteins that are frequently lost or down-regulated in cancers are marked with green arrows, overexpressed proteins with red arrows.
OPEN IN VIEWER
p27KIP1 can bind cyclin D–CDK4/6 in an inhibitory or noninhibitory mode, depending on p27KIP1 phosphorylation status. Cyclin D–p27KIP1-CDK4/6 complexes are catalytically inactive unless p27KIP1 is phosphorylated on Tyr88 and Tyr89 (11). Two molecular mechanisms may explain this switch. First, Tyr88/Tyr89 phosphorylation may dislodge the helix of p27KIP1 from the CDK active site and allow adenosine triphosphate (ATP) binding (12). Second, the presence of tyrosine-unphosphorylated p27KIP1 within the cyclin D–CDK4 complex prevents the activating phosphorylation of CDK4’s T-loop by the CDK-activating kinase (CAK) (12). Brk has been identified as a physiological kinase of p27KIP1 (13); Abl and Lyn can phosphorylate p27KIP1 in vitro, but their in vivo importance remains unclear (1114).
The activity of cyclin D–CDK4/6 is also regulated by proteolysis. Cyclin D1 is an unstable protein with a half-life of less than 30 min. At the end of G1 phase, cyclin D1 is phosphorylated at Thr286 by GSK3β (15). This facilitates association of cyclin D1 with the nuclear exportin CRM1 and promotes export of cyclin D1 from the nucleus to the cytoplasm (16). Subsequently, phosphorylated cyclin D1 becomes polyubiquitinated by E3 ubiquitin ligases, thereby targeting it for proteasomal degradation. Several substrate receptors of E3 ubiquitin ligases have been implicated in recognizing phosphorylated cyclin D1, including F-box proteins FBXO4 (along with αB crystallin), FBXO31, FBXW8, β-TrCP1/2, and SKP2 (17). The anaphase-promoting complex/cyclosome (APC/C) was also proposed to target cyclin D1 while F-box proteins FBXL2 and FBXL8 target cyclins D2 and D3 (1718). Surprisingly, the level and stability of cyclin D1 was unaffected by depletion of several of these proteins, indicating that some other E3 plays a rate-limiting role in cyclin D1 degradation (19). Indeed, recent studies reported that D-cyclins are ubiquitinated and targeted for proteasomal degradation by the E3 ubiquitin ligase CRL4, which uses AMBRA1 protein as its substrate receptor (2022).

Cyclin D–CDK4/6 in cancer

Genomic aberrations of the cyclin D1 gene (CCND1) represent frequent events in different tumor types. The t(11;14)(q13;q32) translocation juxtaposing CCND1 with the immunoglobulin heavy-chain (IGH) locus represents the characteristic feature of mantle-cell lymphoma and is frequently observed in multiple myeloma or plasma cell leukemia (2324). Amplification of CCND1 is seen in many other malignancies—for example, in 13 to 20% of breast cancers (2324), more than 40% of head and neck squamous cell carcinomas, and more than 30% of esophageal squamous cell carcinomas (23). A higher proportion of cancers (e.g., up to 50% of mammary carcinomas) overexpress cyclin D1 protein (24). Also, cyclins D2 and D3, CDK4, and CDK6 are overexpressed in various tumor types (59). Cyclin D–CDK4/6 can also be hyperactivated through other mechanisms such as deletion or inactivation of INK inhibitors, most frequently p16INK4A (5923). Altogether, a very large number of human tumors contain lesions that hyperactivate cyclin D–CDK4/6 (5).
An oncogenic role for cyclin D–CDK4/6 has been supported by mouse cancer models. For example, targeted overexpression of cyclin D1 in mammary glands of transgenic mice led to the development of mammary carcinomas (25). Also, overexpression of cyclin D2, D3, or CDK4, or loss of p16INK4a resulted in tumor formation (9).
Conversely, genetic ablation of D-cyclins, CDK4, or CDK6 decreased tumor sensitivity (9). For instance, Ccnd1– or Cdk4-null mice, or knock-in mice expressing kinase-inactive cyclin D1–CDK4/6, were resistant to develop human epidermal growth factor receptor 2 (HER2)–driven mammary carcinomas (2629). An acute, global shutdown of cyclin D1 in mice bearing HER2-driven tumors arrested tumor growth and triggered tumor-specific senescence while having no obvious impact on normal tissues (30). Likewise, an acute ablation of CDK4 arrested tumor cell proliferation and triggered tumor cell senescence in a KRAS-driven non–small-cell lung cancer (NSCLC) mouse model (31). These observations indicated that CDK4 and CDK6 might represent excellent therapeutic targets in cancer treatment.

CDK4/6 functions in cell proliferation and oncogenesis

The best-documented function of cyclin D–CDK4/6 in driving cell proliferation is phosphorylation of the retinoblastoma protein, RB1, and RB-like proteins, RBL1 and RBL2 (56) (Fig. 1). Unphosphorylated RB1 binds and inactivates or represses E2F transcription factors. According to the prevailing model, phosphorylation of RB1 by cyclin D–CDK4/6 partially inactivates RB1, leading to release of E2Fs and up-regulation of E2F-transcriptional targets, including cyclin E. Cyclin E forms a complex with its kinase partner, CDK2, and completes full RB1 phosphorylation, leading to activation of the E2F transcriptional program and facilitating S-phase entry (56). In normal, nontransformed cells, the activity of cyclin D–CDK4/6 is tightly regulated by the extracellular mitogenic milieu. This links inactivation of RB1 with mitogenic signals. In cancer cells carrying activating lesions in cyclin D–CDK4/6, the kinase is constitutively active, thereby decoupling cell division from proliferative and inhibitory signals (5).
This model has been questioned by the demonstration that RB1 exists in a monophosphorylated state throughout G1 phase and becomes inactivated in late G1 by cyclin E–CDK2, which “hyperphosphorylates” RB1 on multiple residues (32). However, recent single-cell analyses revealed that cyclin D–CDK4/6 activity is required for the hyperphosphorylation of RB1 throughout G1, whereas cyclin E/A–CDK maintains RB1 hyperphosphorylation in S phase (33). Moreover, phosphorylation of RB1 by cyclin D–CDK4/6 was shown to be required for normal cell cycle progression (34).
In addition to this kinase-dependent mechanism, up-regulation of D-cyclin expression and formation of cyclin D–CDK4/6 complexes lead to redistribution of KIP/CIP inhibitors from cyclin E–CDK2 complexes (which are inhibited by these proteins) to cyclin D–CDK4/6 (which use them as assembly factors), thereby activating the kinase activity of cyclin E–CDK2 (6). Cyclin E–CDK2 in turn phosphorylates RB1 and other cellular proteins and promotes cell cycle progression.
Cyclin D1–CDK4/6 directly phosphorylates, stabilizes, and activates the transcription factor FOXM1. This promotes cell cycle progression and protects cancer cells from entering senescence (35). Cyclin D–CDK4 also phosphorylates and inactivates SMAD3, which mediates transforming growth factor–β (TGF-β) antiproliferative response. CDK4/6-dependent phosphorylation of SMAD3 inhibits its transcriptional activity and disables the ability of TGF-β to induce cell cycle arrest (36). FZR1/CDH1, an adaptor protein of the APC complex, is another phosphorylation substrate of CDK4. Depletion of CDH1 in human cancer cells partially rescued the proliferative block upon CDK4/6 inhibition, and it cooperated with RB1 depletion in restoring full proliferation (37).
Cyclin D–CDK4/6 also phosphorylates and inactivates TSC2, a negative regulator of mTORC1, thereby resulting in mTORC1 activation. Conversely, inhibition of CDK4/6 led to decreased mTORC1 activity and reduced protein synthesis in cells representing different human tumor types. It was proposed that through TSC2 phosphorylation, activation of cyclin D–CDK4/6 couples cell growth with cell division (38). Consistent with this, the antiproliferative effect of CDK4/6 inhibition was reduced in cells lacking TSC2 (38).
MEP50, a co-regulatory factor of protein arginine-methyltransferase 5 (PRMT5), is phosphorylated by cyclin D1–CDK4. Through this mechanism, cyclin D1–CDK4/6 increases the catalytic activity of PRMT5/MEP50 (39). It was proposed that deregulation of cyclin D1–CDK4 kinase in tumor cells, by increasing PRMT5/MEP50 activity, reduces the expression of CUL4, a component of the E3 ubiquitin-ligase complex, and stabilizes CUL4 targets such as CDT1 (39). In addition, by stimulating PRMT5/MEP50-dependent arginine methylation of p53, cyclin D–CDK4/6 suppresses the expression of key antiproliferative and pro-apoptotic p53 target genes (40). Another study proposed that PRMT5 regulates splicing of the transcript encoding MDM4, a negative regulator of p53. CDK4/6 inhibition reduced PRMT5 activity and altered the pre-mRNA splicing of MDM4, leading to decreased levels of MDM4 protein and resulting in p53 activation. This, in turn, up-regulated the expression of a p53 target, p21CIP1, that blocks cell cycle progression (41).
During oncogenic transformation of hematopoietic cells, chromatin-bound CDK6 phosphorylates the transcription factors NFY and SP1 and induces the expression of p53 antagonists such as PRMT5, PPM1D, and MDM4 (42). Also, in acute myeloid leukemia cells expressing constitutively activated FLT3, CDK6 binds the promoter region of the FLT3 gene as well as the promoter of PIM1 pro-oncogenic kinase and stimulates their expression. Treatment of FLT3-mutant leukemic cells with a CDK4/6 inhibitor decreased FLT3 and PIM1 expression and triggered cell cycle arrest and apoptosis (43). The relevance of these various mechanisms in the context of human tumors is unclear and requires further study.

Mechanism of action of CDK4/6 inhibitors

Three small-molecule CDK4/6 inhibitors have been extensively characterized in preclinical studies: palbociclib and ribociclib, which are highly specific CDK4/6 inhibitors, and abemaciclib, which inhibits CDK4/6 and other kinases (Table 1). It has been assumed that these compounds act in vivo by directly inhibiting cyclin D–CDK4/6 (9). This simple model has been recently questioned by observations that palbociclib inhibits only cyclin D–CDK4/6 dimers, but not trimeric cyclin D–CDK4/6-p27KIP1 (44). However, it is unlikely that substantial amounts of cyclin D–CDK4 dimers ever exist in cells, because nearly all cyclin D–CDK4 in vivo is thought to be complexed with KIP/CIP proteins (111444). Palbociclib also binds monomeric CDK4 (44). Surprisingly, treatment of cancer cells with palbociclib for 48 hours failed to inhibit CDK4 kinase, despite cell cycle arrest, but it inhibited CDK2 (44). Hence, palbociclib might prevent the formation of active CDK4-containing complexes (through binding to CDK4) and indirectly inhibit CDK2 by liberating KIP/CIP inhibitors. This model needs to be reconciled with several observations. First, treatment of cells with CDK4/6 inhibitors results in a rapid decrease of RB1 phosphorylation on cyclin D–CDK4/6-dependent sites, indicating an acute inhibition of CDK4/6 (4547). Moreover, CDK4/6 immunoprecipitated from cells can be inhibited by palbociclib (48) and p21CIP-associated cyclin CDK4/6 kinase is also inhibited by treatment of cells with palbociclib (49). Lastly, CDK2 is dispensable for proliferation of several cancer cell lines (5051), hence the indirect inhibition of CDK2 alone is unlikely to be responsible for cell cycle arrest.
Name of compound IC50 Other known targets Stage of clinical development
Palbociclib (PD-0332991) D1-CDK4, 11 nM;
D2-CDK6, 15 nM;
D3-CDK4, 9 nM
FDA-approved for HR+/HER2 advanced
breast cancer in combination with
endocrine therapy; phase 2/3 trials
for several other tumor types
Ribociclib (LEE011) D1-CDK4, 10 nM;
D3-CDK6, 39 nM
FDA-approved for HR+/HER2 advanced
breast cancer in combination with
endocrine therapy; phase 2/3 trials
for several other tumor types
Abemaciclib (LY2835219) D1-CDK4, 0.6 to 2 nM;
D3-CDK6, 8 nM
Cyclin T1–CDK9, PIM1, HIPK2, CDKL5,
CAMK2A, CAMK2D, CAMK2G,
GSK3α/β, and (at higher doses)
cyclin E/A–CDK2 and cyclin B–CDK1
FDA-approved for early (adjuvant) and
advanced HR+/HER2 breast cancer in
combination with endocrine therapy;
FDA-approved as monotherapy in advanced
HR+/HER2 breast cancer; phase 2/3 trials
for several other tumor types
Trilaciclib (G1T28) D1-CDK4, 1 nM;
D3-CDK6, 4 nM
FDA-approved for small-cell lung cancer
to reduce chemotherapy-induced bone
marrow suppression; phase 2/3 trials
for other solid tumors
Lerociclib (G1T38) D1-CDK4, 1 nM;
D3-CDK6, 2 nM
Phase 1/2 trials for HR+/HER2 advanced
breast cancer and EGFR-mutant
non–small-cell lung cancer
SHR6390 CDK4, 12 nM;
CDK6, 10 nM
Phase 1/2/3 trials for HR+/HER2 advanced
breast cancer and other solid tumors
PF-06873600 CDK4, 0.13 nM (Ki),
CDK6, 0.16 nM (Ki)
CDK2, 0.09 nM (Ki) Phase 2 trials for HR+/HER2 advanced
breast cancer and other solid tumors
FCN-437 D1-CDK4, 3.3 nM;
D3-CDK6, 13.7 nM
Phase 1/2 trials for HR+/HER2 advanced
breast cancer and other solid tumors
Birociclib (XZP-3287) Not reported Phase 1/2 trials for HR+/HER2 advanced
breast cancer and other solid tumors
HS-10342 Not reported Phase 1/2 trials for HR+/HER2 advanced
breast cancer and other solid tumors
CS3002 Not reported Phase 1 trial for solid tumors

Expand for more

Table 1. Currently available CDK4/6 inhibitors.
This table lists major inhibitors of CDK4 and CDK6, half-maximal inhibitory concentration (IC50) for different cyclin D–CDK4/6 complexes (if known), other known targets, and the stage of clinical development. Ki, inhibitory constant.
OPEN IN VIEWER
Palbociclib, ribociclib, and abemaciclib were shown to block binding of CDK4 and CDK6 to CDC37, the kinase-targeting subunit of HSP90, thereby preventing access of CDK4/6 to the HSP90-chaperone system (52). Because the HSP90-CDC37 complex stabilizes several kinases (53), these observations suggest that CDK4/6 inhibitors, by disrupting the interaction between CDC37 and CDK4 or CDK6, might promote degradation of CDK4 and CDK6. However, depletion of CDK4/6 is typically not observed upon treatment with CDK4/6 inhibitors (54). More studies are needed to resolve these conflicting reports and to establish how CDK4/6 inhibitors affect the cell cycle machinery in cancer cells.

Validation of CDK4/6 inhibitors as anticancer agents

Consistent with the notion that RB1 represents the major rate-limiting substrate of cyclin D–CDK4/6 in cell cycle progression (5557), palbociclib, ribociclib, and abemaciclib were shown to block proliferation of several RB1-positive cancer cell lines, but not cell lines that have lost RB1 expression (465859). Breast cancer cell lines representing the luminal, estrogen receptor–positive (ER+) subtype were shown to be most susceptible to cell proliferation arrest upon palbociclib treatment (45). Palbociclib, ribociclib, abemaciclib, and another CDK4/6 inhibitor, lerociclib, were demonstrated to display potent antitumor activity in xenografts of several tumor types, including breast cancers (466062). Palbociclib and abemaciclib cross the blood-brain barrier and inhibit growth of intracranial glioblastoma (GBM) xenografts, with abemaciclib being more efficient in reaching the brain (6364). Recently, additional CDK4/6 inhibitors were shown to exert therapeutic effects in mouse xenograft models of various cancer types, including SHR6390 (65), FCN-437 (66), and compound 11 (67); the latter two were reported to cross the blood-brain barrier. In most in vivo studies, the therapeutic effect was dependent on expression of intact RB1 protein in tumor cells (4663). However, antitumor effects of palbociclib were also reported in bladder cancer xenografts independently of RB1 status; this was attributed to decreased phosphorylation of FOXM1 (68).

Tumor cell senescence upon CDK4/6 inhibition

In addition to blocking cell proliferation, inhibition of CDK4/6 can also trigger tumor cell senescence (63), which depends on RB1 and FOXM1 (3554). The role of RB1 in enforcing cellular senescence is well established (69). In addition, cyclin D–CDK4/6 phosphorylates and activates FOXM1, which has anti-senescence activity (3570). Senescence represents a preferred therapeutic outcome to cell cycle arrest, as it may lead to a durable inhibition of tumor growth.
It is not clear what determines the extent of senescence upon treatment of cancer cells with CDK4/6 inhibitors. A recent study showed that inhibition of CDK4/6 leads to an RB1-dependent increase in reactive oxygen species (ROS) levels, resulting in activation of autophagy, which mitigates the senescence of breast cancer cells in vitro and in vivo (71). Co-treatment with palbociclib plus autophagy inhibitors strongly augmented the ability of CDK4/6 inhibitors to induce tumor cell senescence and led to sustained inhibition of cancer cell proliferation in vitro and of xenograft growth in vivo (71). Decreased mTOR signaling after long-term CDK4/6 inhibition was shown to be essential for the induction of senescence in melanoma cells, and activation of mTORC1 overrode palbociclib-induced senescence (72). Others postulated that expression of the chromatin-remodeling enzyme ATRX and degradation of MDM2 determines the choice between quiescence and senescence upon CDK4/6 inhibition (73). Inhibition of CDK4 causes dissociation of the deubiquitinase HAUSP/USP7 from MDM2, thereby driving autoubiquitination and proteolytic degradation of MDM2, which in turn promotes senescence. This mechanism requires ATRX, which suggests that expression of ATRX can be used to predict the senescence response (73). Two additional proteins that play a role in this process are PDLIM7 and type II cadherin CDH18. Expression of CDH18 correlated with a sustained response to palbociclib in a phase 2 trial for patients with liposarcoma (74).

Markers predicting response to CDK4/6 inhibition

Only tumors with intact RB1 respond to CDK4/6 inhibitor treatment by undergoing cell cycle arrest or senescence (958). In addition, “D-cyclin activating features” (CCND1 translocation, CCND2 or CCND3 amplification, loss of the CCND1-3 3′-untranslated region, and deletion of FBXO31 encoding an F-box protein implicated in cyclin D1 degradation) were shown to confer a strong response to abemaciclib in cancer cell lines (58). Moreover, co-deletion of CDKN2A and CDKN2C (encoding p16INK4A/p19ARF and p18INK4C, respectively) confers palbociclib sensitivity in glioblastoma (75). Thr172 phosphorylation of CDK4 and Tyr88 phosphorylation of p27KIP1 (both associated with active cyclin D–CDK4) correlate with sensitivity of breast cancer cell lines or tumor explants to palbociclib (7677). Surprisingly, in PALOMA-1, PALOMA-2, and PALOMA-3 trials (7880), and in another independent large-scale study (81), CCND1 gene amplification or elevated levels of cyclin D1 mRNA or protein were not predictive of palbociclib efficacy. Conversely, overexpression of CDK4, CDK6, or cyclin E1 is associated with resistance of tumors to CDK4/6 inhibitors (see below).

Synergy of CDK4/6 inhibitors with other compounds

Several preclinical studies have documented the additive or synergistic effects of combining CDK4/6 inhibitors with inhibitors of the receptor tyrosine kinases as well as phosphoinositide 3-kinase (PI3K), RAF, or MEK (Table 2). This synergism might be because these pathways impinge on the cell cycle machinery through cyclin D–CDK4/6 (8286). In some cases, the effect was seen in the presence of specific genetic lesions, such as EGFRBRAFV600EKRAS, and PIK3CA mutations (598789) (Table 2). When comparing different dosing regimens, continuous treatment with a MEK inhibitor with intermittent palbociclib resulted in more complete tumor responses than other combination schedules (90). Treatment with CDK4/6 inhibitors sensitized cancer cells to ionizing radiation (63) or cisplatin (68). The synergism with platinum-based chemotherapy was attributed to the observation that upon this treatment, CDK6 phosphorylates and stabilizes the FOXO3 transcription factor, thereby promoting tumor cell survival. Consequently, inhibition of CDK6 increases platinum sensitivity by enhancing tumor cell death (91).
CDK4/6 inhibitor Synergistic target Inhibitor Disease
Palbociclib PI3K Taselisib, pictilisib PIK3CA mutant TNBC
AR Enzalutamide Androgen receptor–positive TNBC
EGFR Erlotinib TNBC, esophageal squamous cell carcinoma
RAF PLX4720 BRAF-V600E mutant melanoma
MEK Trametinib KRAS mutant colorectal cancer
MEK PD0325901 (mirdametinib) KRAS or BRAFV600E mutant colorectal cancer
MEK MEK162 (binimetinib) KRAS mutant colorectal cancer
MEK AZD6244 (selumetinib) Pancreatic ductal adenocarcinoma
PI3K/mTOR BEZ235 (dactolisib), AZD0855, GDC0980 (apitolisib) Pancreatic ductal adenocarcinoma
IGF1R/InsR BMS-754807 Pancreatic ductal adenocarcinoma
mTOR Temsirolimus Pancreatic ductal adenocarcinoma
mTOR AZD2014 (vistusertib) ER+ breast cancer
mTOR MLN0128 (sapanisertib) Intrahepatic cholangiocarcinoma
mTOR Everolimus Melanoma, glioblastoma
Ribociclib PI3K GDC-0941 (pictilisib), BYL719 (alpelisib) PIK3CA mutant breast cancer
PDK1 GSK2334470 ER+ breast cancer
EGFR Nazartinib EGFR-mutant lung cancer
RAF Encorafenib BRAF-V600E mutant melanoma
mTOR Everolimus T-ALL
Inflammation Glucocorticoid dexamethasone T-ALL
γ-Secretase Compound E T-ALL
Abemaciclib HER2 Trastuzumab HER2+ breast cancer
EGFR and HER2 Lapatinib HER2+ breast cancer
RAF LY3009120, vemurafenib KRAS mutant lung or colorectal cancer, NRAS or
BRAF-V600E mutant melanoma
Temozolomide (alkylating agent) Glioblastoma

Expand for more

Table 2. Combination treatments that demonstrated synergy with CDK4/6 inhibitors in preclinical studies.
TNBC, triple-negative breast cancer; AR, androgen receptor; ER+, estrogen receptor–positive; T-ALL, T cell acute lymphoblastic leukemia; HER2+, human epidermal growth factor receptor 2–positive; PI3K, phosphoinositide 3-kinase; EGFR, epidermal growth factor receptor; IGF1R, insulin-like growth factor 1 receptor, InsR, insulin receptor.
OPEN IN VIEWER
In several instances, co-treatment with CDK4/6 inhibitors prevented the development of resistance to other compounds or inhibited the proliferation of resistant tumor cells. Co-treatment of melanoma patient-derived xenografts (PDXs) with ribociclib plus the RAF inhibitor encorafenib delayed or prevented development of encorafenib resistance (92). PDXs that acquired encorafenib resistance remained sensitive to the combination of encorafenib plus ribociclib (59). Treatment of BRAFV600E-mutant melanoma xenografts with palbociclib plus the BRAFV600E inhibitor PLX4720 prevented development of resistance (89). BRAFV600E-mutant melanoma cell lines that acquired resistance to the BRAFV600E inhibitor vemurafenib remained sensitive to palbociclib or abemaciclib, and xenografts underwent senescence and tumor regression upon CDK4/6 inhibition (7293). Treatment of ALK-mutant, ALK kinase inhibitor–resistant neuroblastoma xenografts with palbociclib restored the sensitivity to these compounds (94). A combination of PI3K and CDK4/6 inhibitors overcame the intrinsic and acquired resistance of breast cancers to PI3K inhibitors and resulted in regression of PIK3CA-mutant xenografts (88).
Up-regulation of cyclin D1 expression was shown to mediate acquired resistance of HER2+ tumors to anti-HER2 therapies in a mouse breast cancer model (95). Treatment of mice bearing trastuzumab-resistant tumors or PDXs of resistant HER2+ mammary carcinomas with abemaciclib restored the sensitivity of tumors to HER2 inhibitors and inhibited tumor cell proliferation. Moreover, in the case of treatment-naïve tumors, co-administration of abemaciclib significantly delayed the development of resistance to anti-HER2 therapies (95).
Several anticancer treatments, such as chemotherapy, target dividing cells. Because CDK4/6 inhibitors block tumor cell proliferation, they might impede the effects of chemotherapy. Indeed, several reports have documented that co-administration of CDK4/6 inhibitors antagonized the antitumor effects of compounds that act during S phase (doxorubicin, gemcitabine, methotrexate, mercaptopurine) or mitosis (taxanes) (9697). However, some authors reported synergistic effects (9899), although the molecular underpinnings are unclear.
A recent report documented that administration of CDK4/6 inhibitors prior to taxanes inhibited tumor cell proliferation and impeded the effect of taxanes (100). By contrast, administration of taxanes first (or other chemotherapeutic compounds that act on mitotic cells or cells undergoing DNA synthesis), followed by CDK4/6 inhibitors, had a strong synergistic effect. The authors showed that by repressing the E2F-dependent transcriptional program, CDK4/6 inhibitors impaired the expression of genes required for DNA-damage repair via homologous recombination. Because treatment of cancer cells with chemotherapy triggers DNA damage, the impairment of DNA-damage repair induced cytotoxicity, thereby explaining the synergistic effect (100).
Cells with impaired homologous recombination rely on poly-(ADP-ribose) polymerase (PARP) for double-stranded DNA-damage repair, which renders them sensitive to PARP inhibition. Indeed, a strong synergistic effect has been demonstrated between CDK4/6 inhibitors and PARP inhibitors in PDX-derived cell lines (100). Such synergy was also reported for ovarian cancer cells (101). Another study found that inhibition of CDK4/6 resulted in down-regulation of PARP levels (102).

Protection against chemotherapy-induced toxicity

Administration of palbociclib to mice induced reversible quiescence in hematopoietic stem/progenitor cells (HSPCs). This effect protected mice from myelosuppression after total-body irradiation. Moreover, treatment of tumor-bearing mice with CDK4/6 inhibitors together with irradiation mitigated radiation-induced toxicity without compromising the therapeutic effect (103). Co-administration of a CDK4/6 inhibitor, trilaciclib, with cytotoxic chemotherapy (5-FU, etoposide) protected animals from chemotherapy-induced exhaustion of HSPCs, myelosuppression, and apoptosis of bone marrow (60104). These observations led to phase 2 clinical trial, which evaluated the effects of trilaciclib administered prior to etoposide and carboplatin for treatment of small-cell lung cancer. Trilaciclib improved myelopreservation while having no adverse effect on antitumor efficacy (105). A similar phase 2 clinical trial investigating trilaciclib in combination with gemcitabine and carboplatin chemotherapy in patients with metastatic triple-negative breast cancer (TNBC) did not observe a significant difference in myelosuppression. However, this study demonstrated an overall survival benefit of the combination therapy (106107).

Metabolic function of CDK4/6 in cancer cells

The role of CDK4/6 in tumor metabolism is only starting to be appreciated (Fig. 2A). Treatment of pancreatic cancer cells with CDK4/6 inhibitors was shown to induce tumor cell metabolic reprogramming (108). CDK4/6 inhibition increased the numbers of mitochondria and lysosomes, activated mTOR, and increased the rate of oxidative phosphorylation, likely through an RB1-dependent mechanism (108). Combined inhibition of CDK4/6 and mTOR strongly suppressed tumor cell proliferation (108). Moreover, CDK4/6 can phosphorylate and inactivate TFEB, the master regulator of lysosomogenesis, and through this mechanism reduce lysosomal numbers. Conversely, CDK4/6 inhibition activated TFEB and increased the number of lysosomes (109). Another mechanism linking CDK4/6 and lysosomes was provided by the observation that treatment of TNBC cells with CDK4/6 inhibitors decreased mTORC1 activity and impaired the recruitment of mTORC1 to lysosomes (110). Consistent with the idea that mTORC1 inhibits lysosomal biogenesis, CDK4/6 inhibition increased the number of lysosomes in tumor cells. Because an increased lysosomal biomass underlies some cases of CDK4/6 inhibitor resistance (see below) (111), stimulation of lysosomogenesis by CDK4/6 inhibitors might limit their clinical efficacy by inducing resistance.
Fig. 2. CDK4 and CDK6: More than cell cycle kinases.
Although the role of CDK4 and CDK6 in cell cycle progression has been well documented, both kinases regulate several other functions that are only now starting to be unraveled. (A) Inhibition of CDK4/6 (CDK4/6i) affects lysosome and mitochondrial numbers as well as oxidative phosphorylation. Cyclin D3–CDK6 phosphorylates glycolytic enzymes 6-phosphofructokinase (PFKP) and pyruvate kinase M2 (PKM2), thereby controlling ROS levels via the pentose phosphate (PPP) and serine synthesis pathways. (B) Inhibition of CDK4/6 affects antitumor immunity, acting both within cancer cells and on the immune system of the host. In tumor cells, inhibition of CDK4/6 impedes expression of an E2F target, DNA methyltransferase (DNMT). DNMT inhibition reduces methylation of endogenous retroviral genes (ERV) and increases intracellular levels of double-stranded RNA (dsRNA) (114). In effector T cells, inhibition of CDK4/6 stimulates NFAT transcriptional activity and enhances secretion of IFN-γ and interleukin 2 (IL-2) (115).
OPEN IN VIEWER
Lastly, CDK4/6 inhibition impaired lysosomal function and the autophagic flux in cancer cells. It was argued that this lysosomal dysfunction was responsible for the senescent phenotype in CDK4/6 inhibitor–treated cells (110). Because lysosomes are essential for autophagy, the authors co-treated TNBC xenografts with abemaciclib plus an AMPK activator, A769662 (which induces autophagy), and found that this led to cancer cell death and subsequent regression of tumors (110).
Cyclin D3–CDK6 phosphorylates and inhibits two rate-limiting glycolytic enzymes, 6-phosphofructokinase and pyruvate kinase M2. This redirects glycolytic intermediates into the pentose phosphate pathway (PPP) and serine synthesis pathway. Through this mechanism, cyclin D3–CDK6 promotes the production of reduced nicotinamide adenine dinucleotide phosphate (NADPH) and reduced glutathione (GSH) and helps to neutralize ROS (112). Treatment of tumors expressing high levels of cyclin D3–CDK6 (such as leukemias) with CDK4/6 inhibitors reduced the PPP- and serine-synthesis pathway flow, thereby depleting the antioxidants NADPH and GSH. This increased ROS levels and triggered tumor cell apoptosis (112).
Another link between cyclin D–CDK4/6 in metabolism and cancer was provided by the observation that livers of obese/diabetic mice up-regulate cyclin D1 expression (113). Treatment of these mice with an antidiabetic compound, metformin, reduced liver cyclin D1 levels and largely protected mice against development of hepatocellular carcinoma. Also, genetic ablation of cyclin D1 protected obese/diabetic mice from liver cancer, and administration of palbociclib inhibited liver cancer progression. These treatments had no effect on tumors in nonobese animals (113). These observations raise the possibility of using antidiabetic compounds with CDK4/6 inhibitors for treatment of liver cancers in obese patients.

CDK4/6 inhibitors and antitumor immune responses

Several recent reports have started to unravel how inhibition of CDK4/6 influences antitumor immune responses, acting both on tumor cells as well as on the tumor immune environment (Fig. 2B). Treatment of breast cancer–bearing mice or breast cancer cells with abemaciclib activated expression of endogenous retroviral elements in tumor cells, thereby increasing the levels of double-stranded RNA. This, in turn, stimulated production of type III interferons and increased presentation of tumor antigens. Hence, CDK4/6 inhibitors, by inducing viral gene expression, trigger antiviral immune responses that help to eliminate the tumor (114).
Inhibition of CDK4/6 also affects the immune system by impeding the proliferation of CD4+FOXP3+ regulatory T cells (Tregs), which normally inhibit the antitumor response. Because cytotoxic CD8+ T cells are less affected by CDK4/6 inhibition, abemaciclib treatment decreases the Treg/CD8+ ratio of intratumoral T cells and facilitates tumor cell killing by cytotoxic CD8+ T cells (114).
Inhibition of CDK4/6 also resulted in activation of T cells through derepression of NFAT signaling. NFAT4 (and possibly other NFATs) are phosphorylated by cyclin D3–CDK6 (115). Inhibition of CDK4/6 decreased phosphorylation of NFATs, resulting in their nuclear translocation and enhanced transcriptional activity. This caused up-regulation of NFAT targets, resulting in T cell activation, which enhanced the antitumor immune response. In addition, CDK4/6 inhibitors increased the infiltration of effector T cells into tumors, likely because of elevated levels of chemokines CXCL9 and CXCL10 after CDK4/6 inhibitor treatment (115). Abemaciclib treatment also induced inflammatory and activated T cell phenotypes in tumors and up-regulated the expression of immune checkpoint proteins CD137, PD-L1, and TIM-3 on CD4+ and CD8+ cells (116).
CDK4/6 inhibition also caused up-regulation of PD-L1 protein expression in tumor cells (117118). This effect was shown to be independent of RB1 status in the tumor. Mechanistically, CDK4/6 phosphorylates and stabilizes SPOP, which promotes PD-L1 polyubiquitination and degradation (118). Cyclin D–CDK4 also represses expression of PD-L1 through RB1. Specifically, cyclin D–CDK4/6-mediated phosphorylation of RB1 on S249/T252 promotes binding of RB1 to NF-κB protein p65, and this represses the expression of a subset NF-κB–regulated genes, including PD-L1 (119).
These observations prompted tests of the efficacy of combining CDK4/6 inhibitors with antibodies that elicit immune checkpoint blockade. Indeed, treatment of mice bearing autochthonous breast cancers, or cancer allografts, with CDK4/6 inhibitors together with anti-PD-1/PD-L1 antibodies enhanced the efficacy of immune checkpoint blockade and led to complete tumor regression in a high proportion of animals (114115118). Conversely, activation of the cyclin D–CDK4 pathway by genomic lesions in human melanomas correlated with resistance to anti–PD-1 therapy (117).
Some authors did not observe synergy when abemaciclib was administered concurrently with immune checkpoint inhibitors in allograft tumor models (116120). However, a strong synergistic antitumor effect was detected when abemaciclib was administered first (and continued) and anti–PD-L1 antibody was administered later. The combined treatment induced immunological memory, as mice that underwent tumor regression were resistant to rechallenge with the same tumor (116). Abemaciclib plus anti–PD-L1 treatment increased infiltration of CD4+ and CD8+ T cells into tumors, and increased the expression of major histocompatibility complex class I (MHC-I) and MHC-II on tumor cells and on macrophages and MHC-I on dendritic cells (116). In the case of anti–CTLA-4 plus anti–PD-1 treatment in melanoma allograft model, the synergistic effect was observed when immune checkpoint inhibitor treatment was started first, followed by abemaciclib (120).
The synergistic antitumor effect of PI3K and CDK4/6 inhibitors in TNBC is mediated, in part, by enhancement of tumor immunogenicity (121). Combined treatment of TNBC cells with ribociclib plus the PI3K inhibitor apelisib synergistically up-regulated the expression of immune-related pathways in tumor cells, including proteins involved in antigen presentation. Co-treatment of tumor-bearing mice also decreased proliferation of CD4+FOXP3+ Treg cells, increased activation of intratumoral CD4+ and CD8+ T cells, increased the frequency of tumor-infiltrating NKT cells, and decreased the numbers of intratumoral immunosuppressive myeloid-derived suppressor cells. Moreover, combined treatment strongly augmented the response to immune checkpoint therapy with PD-1 and CTLA-4 antibodies (121).
Single-cell RNA sequencing of human melanomas identified an immune resistance program expressed by tumor cells that correlates with T cell exclusion from the tumor mass and immune evasion by tumor cells. The program can predict the response of tumors to immune checkpoint inhibitors. Treatment of human melanoma cells with abemaciclib repressed this program in an RB1-dependent fashion (120).
Together, these findings indicate that CDK4/6 inhibitors may convert immunologically “cold” tumors into “hot” ones. The most pressing issue is to validate these findings in a clinical setting. The utility of combining CDK4/6 inhibitors with PD-1 or PD-L1 antibodies is currently being evaluated in several clinical trials. Note that the effects of CDK4/6 inhibition on the immune system of the host are independent of tumor cell RB1 status, raising the possibility of using CDK4/6 inhibitors to also boost the immune response against RB1-negative tumors.

CDK4/6 inhibitors in clinical trials

Table 3 summarizes major clinical trials with CDK4/6 inhibitors. Given early preclinical data indicating that breast cancers—in particular, the hormone receptor–positive ones—are very sensitive to CDK4/6 inhibition (as discussed above), many clinical trials have focused on this cancer type. Most studies have evaluated CDK4/6 inhibitors administered together with anti-estrogens (the aromatase inhibitors letrozole or anastrozole, or the estrogen receptor antagonist fulvestrant) for treatment of advanced/metastatic HR+/HER2 breast cancers in postmenopausal women. Addition of CDK4/6 inhibitors significantly extended median progression-free survival (78122130) and prolonged median overall survival (131134). Moreover, abemaciclib has shown clinical activity when administered as a single agent (135). Consequently, palbociclib, ribociclib, and abemaciclib have been approved by the US Food and Drug Administration (FDA) for treatment of patients with advanced/metastatic HR+/HER2 breast cancer (Box 1). A recent phase 3 clinical trial, MonarchE, evaluated abemaciclib plus standard endocrine therapy in treatment of patients with early-stage, high-risk, lymph node–positive HR+/HER2 breast cancer. Addition of abemaciclib reduced the risk of breast cancer recurrence (136). This is in contrast to the similar PALLAS study reported this year, which found no benefit of adding palbociclib to endocrine therapy for women with early-stage breast cancer (137). Analysis of patient populations in these two trials may help to explain the different outcomes. It is also possible that the favorable outcome of the MonarchE study reflects a broader spectrum of kinases inhibited by abemaciclib. The utility of CDK4/6 inhibitors in early-stage breast cancer remains unclear and is being addressed in ongoing clinical trials (PALLAS, PENELOPE-B, EarLEE-1, MonarchE) (138).
CDK4/6
inhibitor
Trial name Trial details Treatment Patients Outcome Ref. Other outcomes
Palbociclib PALOMA-1 Randomized
phase 2
Aromatase inhibitor
letrozole alone
(standard of care)
versus letrozole
plus palbociclib
Postmenopausal women
with advanced ER+/HER2
breast cancer who had
not received any systemic
treatment for their
advanced disease
Addition of palbociclib markedly
increased median PFS from
10.2 months in the
letrozole group to
20.2 months in the
palbociclib plus
letrozole group
(78) On the basis of this result, palbociclib
received a “Breakthrough Therapy”
designation status from FDA and was
granted accelerated approval, in
combination with letrozole, for the
treatment of ER+/HER2 metastatic
breast cancer
Palbociclib PALOMA-2 Double-blind
phase 3
Palbociclib plus
letrozole as first-
line therapy
Postmenopausal women
with ER+/HER2
breast cancer
Addition of palbociclib strongly
increased median PFS:
14.5 months in the placebo-
letrozole group versus
24.8 months in the
palbociclib-letrozole group
(123) Palbociclib was equally efficacious in
patients with luminal A and B breast
cancers, and there was no single
biomarker associated with the lack of
clinical benefit, except for RB1 loss;
CDK4 amplification was associated
with endocrine resistance, but this
was mitigated by addition of
palbociclib; tumors with high levels
of FGFR2 and ERBB3 mRNA
displayed greater PFS gain
after addition of palbociclib (79)
Palbociclib PALOMA-3 Randomized
phase 3
Estrogen receptor
antagonist
fulvestrant plus
placebo versus
fulvestrant plus
palbociclib
Women with HR+/HER2
metastatic breast cancer
that had progressed on
previous endocrine therapy
The study demonstrated a
substantial prolongation
of median PFS in the palbociclib-
treated group: 4.6 months in the
placebo plus fulvestrant group
versus 9.5 months in the
palbociclib plus fulvestrant
group; addition of palbociclib
also extended median overall
survival from 28.0 months
(placebo-fulvestrant) to
34.9 months (palbociclib-
fulvestrant); estimated rate
of survival at 3 years was
41% versus 50%, respectively
(124125135)
Palbociclib NeoPalAna Palbociclib
in an
neoadjuvant
setting (i.e.,
prior to
surgery)
Compared the effects
of an aromatase
inhibitor anastrozole
versus palbociclib
plus anastrozole
on tumor cell
proliferation
Women with newly
diagnosed clinical
stage II/III ER+/HER2
breast cancer
Addition of palbociclib enhanced
the antiproliferative effect
of anastrozole
(161)
Palbociclib PALLAS Randomized
phase 3
Palbociclib plus
standard endocrine
therapy versus
endocrine therapy
alone
Patients with early
(stage 2 or 3),
HR+/HER2
breast cancer
Preliminary results indicate that
the trial is unlikely to show
a statistically significant
improvement of invasive
disease-free survival
(138)
Palbociclib PENELOPE-B Palbociclib in
patients with
early breast
cancer at high
risk of recurrence
Ongoing
Ribociclib MONA
LEESA-2
Randomized
phase 3
Ribociclib plus
letrozole versus
placebo plus
letrozole
First-line treatment for
postmenopausal women
with HR+/HER2 recurrent
or metastatic breast
cancer who had not
received previous
systemic therapy for
advanced disease
At 18 months, PFS
was 42.2% in the
placebo-letrozole
group and 63.0%
in the ribociclib-
letrozole group
(126)
Ribociclib MONA
LEESA-3
Phase 3 Ribociclib plus
fulvestrant
Patients with advanced
(metastatic or recurrent)
HR+/HER2 breast cancer
who have either received no
treatment for the advanced
disease or previously
received a single line of
endocrine therapy for the
advanced disease
Addition of ribociclib significantly
extended median PFS, from
12.8 months (placebo-fulvestrant)
to 20.5 months (ribociclib-
fulvestrant); overall survival at
42 months was also extended
from 45.9% (placebo-fulvestrant)
to 57.8% (ribociclib-fulvestrant)
(127133)
Ribociclib MONA
LEESA-7
Phase 3
randomized,
double-blind
Ribociclib versus
placebo together
with an anti-
estrogen tamoxifen
or an aromatase
inhibitor (letrozole
or anastrozole)
Premenopausal and
perimenopausal women
with HR+/HER2 advanced
breast cancer who had not
received previous treatment
with CDK4/6 inhibitors
Ribociclib significantly increased
median PFS from 13.0 months in
the placebo-endocrine therapy
group to 23.8 months in the
ribociclib-endocrine therapy
group; overall survival was also
strongly prolonged in the ribociclib
group (estimated overall survival
at 42 months was 46.0% for the
placebo group and 70.2% in the
ribociclib group)
(128132)
Ribociclib EarLEE-1 Phase 3 trial Ribociclib in the
treatment of early-
stage, high-risk
HR+/HER2
breast cancers
Ongoing
Abemaciclib MONARCH 1 Phase 2 trial Abemaciclib as a
single agent
Women with HR+/HER2
metastatic breast cancer
who had progressed on or
after prior endocrine therapy
and had 1 or 2 chemotherapy
regimens in the metastatic
setting
Abemaciclib exhibited promising activity
in these heavily pretreated patients
with poor prognosis; median
PFS was 6.0 months and overall
survival 17.7 months
(136) The most common adverse events
were diarrhea, fatigue, and
nausea (136)
Abemaciclib MONARCH 2 Double-blind
phase 3
Abemaciclib in
combination
with fulvestrant
Women with HR+/HER2 breast
cancer who had progressed
while receiving endocrine
therapy, or while receiving
first-line endocrine therapy for
metastatic disease
Addition of abemaciclib significantly
increased PFS from 9.3 months in
the placebo-fulvestrant to 16.4 in
the abemaciclib-fulvestrant group;
median overall survival was also
extended from 37.3 months
to 46.7 months
(129134)
Abemaciclib MONARCH 3 Randomized
phase 3
double-blind
Abemaciclib plus
an aromatase
inhibitor
(anastrozole
or letrozole)
Postmenopausal women
with advanced HR+/HER2
breast cancer who had
no prior systemic therapy
in the advanced setting
Addition of abemaciclib prolonged
PFS from 14.8 months (in
the placebo-aromatase
inhibitor group) to 28.2 months
(abemaciclib-aromatase
inhibitor group)
(130131)
Abemaciclib MonarchE Phase 3 study Endocrine with
or without
abemaciclib
Patients with HR+/HER2
lymph node–positive,
high-risk early
breast cancer
Preliminary analysis indicates that
addition of abemaciclib resulted
in a significant improvement of
invasive disease-free survival
and of distant relapse-
free survival
(137)
Trilaciclib Randomized
phase 2 study
Chemotherapy alone
(gemcitabine and
carboplatin),
versus concurrent
administration of
trilaciclib plus
chemotherapy,
versus
administration of
trilaciclib prior to
chemotherapy
(to mitigate the
cytotoxic effect of
chemotherapy on
bone marrow)
Patients with recurrent or
metastatic triple-negative
breast cancer who had no
more than two previous
lines of chemotherapy
Addition of trilaciclib did not offer
detectable myeloprotection, but
resulted in increased overall
survival (from 12.8 months in the
chemotherapy-only group to
20.1 months in the concurrent
trilaciclib and chemotherapy
group and 17.8 months in trilaciclib
before chemotherapy group)
(162) The most common adverse events were
neutropenia, thrombocytopenia,
and anemia (162)

Expand for more

Table 3. Major past clinical trials with CDK4/6 inhibitors.
ER+, estrogen receptor–positive; HER2, human epidermal growth factor receptor 2–negative; HR+, hormone receptor–positive; PFS, progression-free survival. FGFR2, fibroblast growth factor receptor 2; ERBB3, receptor tyrosine-protein kinase erbB-3.
OPEN IN VIEWER

Palbociclib

Approved by FDA in 2016, in combination with fulvestrant for the treatment of hormone receptor–positive, HER2-negative (HR+/HER2) advanced or metastatic breast cancer in women with disease progression following endocrine therapy. Approved in 2017 for the treatment of HR+/HER2 advanced or metastatic breast cancer in combination with an aromatase inhibitor as initial endocrine-based therapy in postmenopausal women.
Palbociclib is administered at a dose of 125 mg (given orally) daily for 3 weeks followed by 1 week off, or 200 mg daily for 2 weeks followed by 1 week off. The rate-limiting toxicities are neutropenia, thrombocytopenia, and anemia.

Ribociclib

Approved by FDA in 2017, in combination with an aromatase inhibitor as initial endocrine-based therapy for the treatment of postmenopausal women with HR+/HER2 advanced or metastatic breast cancer. In 2018, the FDA expanded the indication for ribociclib in combination with an aromatase inhibitor for pre/perimenopausal women with HR+/HER2 advanced or metastatic breast cancer, as initial endocrine-based therapy. FDA also approved ribociclib in combination with fulvestrant for postmenopausal women with HR+/HER2 advanced or metastatic breast cancer, as initial endocrine-based therapy or following disease progression on endocrine therapy.
Ribociclib is administered at a dose of 600 mg (given orally) daily for 3 weeks followed by 1 week off. The main toxicities are neutropenia and thrombocytopenia.

Abemaciclib

Approved by FDA in 2017, in combination with fulvestrant for women with HR+/HER2 advanced or metastatic breast cancer with disease progression following endocrine therapy. In addition, abemaciclib was approved as monotherapy for women and men with HR+/HER2 advanced or metastatic breast cancer with disease progression following endocrine therapy and prior chemotherapy in the metastatic setting. Approved by FDA in 2018 in combination with an aromatase inhibitor as initial endocrine-based therapy for postmenopausal women with HR+/HER2 advanced or metastatic breast cancer. Approved by FDA in 2021 for adjuvant treatment of early-stage HR+/HER2 breast cancer in combination with endocrine therapy.
Abemaciclib is administered at a dose of 200 mg (given orally) every 12 hours. The dose-limiting toxicity is fatigue. Neutropenia is also observed but is not rate-limiting. Other severe side effects include diarrhea and nausea.
Currently, palbociclib is being used in 164 active or recruiting clinical trials, ribociclib in 69 trials, and abemaciclib in 98 trials for more than 50 tumor types (139). These trials evaluate combinations of CDK4/6 inhibitors with a wide range of compounds (Table 4). Trials with trilaciclib test the benefit of this compound in preserving bone marrow and the immune system.
Additional target Inhibitor Immune
checkpoint
inhibitor
Tumor
type
Trial identifier
Palbociclib
Aromatase Letrozole, anastrozole,
exemestane
HR+ breast cancer, HR+ ovarian
cancer, metastatic breast cancer,
metastatic endometrial cancer
NCT04130152,
NCT03054363,
NCT03936270,
NCT04047758,
NCT02692755,
NCT02806050,
NCT03870919,
NCT02040857,
NCT04176354,
NCT02028507,
NCT03220178,
NCT02592083,
NCT02603679,
NCT04256941,
NCT03425838,
NCT02894398,
NCT02297438,
NCT02730429,
NCT02142868,
NCT02942355
LHRH LHRH agonists: goserelin,
leuprolide
HR+ breast cancer NCT03969121,
NCT03423199,
NCT01723774,
NCT02917005,
NCT02592746,
NCT03628066
ER ER antagonists: fulvestrant,
tamoxifen
HR+ breast cancer, metastatic
breast cancer
NCT02668666,
NCT02738866,
NCT03184090,
NCT04526028,
NCT02513394,
NCT03560856,
NCT02760030,
NCT03079011,
NCT03227328,
NCT03809988,
NCT02764541,
NCT03007979,
NCT03633331
ER Selective estrogen receptor
degraders (SERDs): G1T48,
ZN-c5, SAR439859,
AZD9833, GDC-9545
HR+ breast cancer NCT03455270,
NCT04546009,
NCT04436744,
NCT04478266,
NCT03560531,
NCT03616587,
NCT03284957,
NCT03332797
ER Selective estrogen receptor
modulator (SERM):
bazedoxifene
HR+ breast cancer NCT03820830,
NCT02448771
Aromatase + PD-1 Letrozole, anastrozole Pembrolizumab,
nivolumab
Stage IV ER+
breast cancer
NCT02778685,
NCT04075604
PD-1 Nivolumab,
pembrolizumab,
MGA012
Liposarcoma NCT04438824
PD-L1 Avelumab AR+ breast cancer, TNBC,
ER+/HER2 metastatic
breast cancer
NCT04360941,
NCT03147287
EGFR + PD-L1 Cetuximab Avelumab Squamous cell carcinoma
of the head and neck
NCT03498378
HER2 Tucatinib, trastuzumab,
pertuzumab,
T-DM1, ZW25
HER2+ breast cancer NCT03530696,
NCT03054363,
NCT02448420,
NCT03709082,
NCT03304080,
NCT02947685
EGFR/HER2 Neratinib Advanced solid tumors with
EGFR mutation/amplification,
HER2 mutation/amplification,
HER3/4 mutation, or
KRAS mutation
NCT03065387
EGFR Cetuximab Metastatic colorectal cancer,
squamous cell carcinoma
of the head and neck
NCT03446157,
NCT02499120
FGFR Erdafitinib ER+/HER2/FGFR-amplified
metastatic breast cancer
NCT03238196
FGFR1-3 Rogaratinib FGFR1-3+/HR+ breast cancer NCT04483505
IGF-1R Ganitumab Ewing sarcoma NCT04129151
VEGF1-3 receptors
+ PD-L1
Axitinib Avelumab NSCLC NCT03386929
RAF Sorafenib Leukemia NCT03132454
MEK PD-0325901,
binimetinib
KRAS mutant NSCLC, TNBC,
KRAS and NRAS mutant
metastatic or unresectable
colorectal cancer
NCT02022982,
NCT03170206,
NCT04494958,
NCT03981614
ERK Ulixertinib Advanced pancreatic cancer
and other solid tumors
NCT03454035
PI3K Copanlisib HR+ breast cancer NCT03128619
PI3K Taselisib, pictilisib,
GDC-0077
PIK3CA mutant advanced solid
tumors, PIK3CA mutant and
HR+ breast cancer
NCT02389842,
NCT04191499,
NCT03006172
PI3K/mTOR Gedatolisib Metastatic breast cancer,
advanced squamous cell lung,
pancreatic, head and neck
cancer and other solid tumors
NCT02684032,
NCT03065062,
NCT02626507
mTOR Everolimus, vistusertib HR+ breast cancer NCT02871791
AKT Ipatasertib HR+ breast cancer, metastatic
breast cancer, metastatic
gastrointestinal tumors,
NSCLC
NCT03959891,
NCT04060862,
NCT04591431
BTK Ibrutinib Mantle cell lymphoma NCT03478514
BCL-2 Venetoclax ER+/BCL-2+ advanced
or metastatic breast
cancer
NCT03900884
AR AR antagonists: bicalutamide AR+ metastatic breast cancer NCT02605486
Lysosome +
aromatase
Hydroxychloroquine + letrozole ER+ breast cancer NCT03774472
Proliferating cells Standard chemotherapy Stage IV ER+ breast cancer NCT03355157
Proliferating cells Radiation Stage IV ER+ breast cancer NCT03870919,
NCT03691493,
NCT04605562
BCR-ABL Bosutinib HR+ breast cancer NCT03854903
Ribociclib
Aromatase Letrozole, anastrozole,
exemestane
HR+ breast cancer,
metastatic breast
cancer, ovarian
cancer
NCT04256941,
NCT03425838,
NCT03822468,
NCT02712723,
NCT03673124,
NCT02941926,
NCT03248427,
NCT03671330,
NCT02333370,
NCT01958021,
NCT03425838
LHRH LHRH agonists:
goserelin, leuprolide
HR+ breast cancer NCT03944434
ER ER antagonists: fulvestrant HR+ breast cancer,
advanced
breast cancer
NCT03227328,
NCT02632045,
NCT02632045,
NCT03555877
PD-1 Spartalizumab Breast cancer and ovarian
cancer, recurrent and/or
metastatic head and neck
squamous cell carcinoma,
melanoma
NCT03294694,
NCT04213404,
NCT03484923
HER2 Trastuzumab, pertuzumab,
T-DM1
HER2+ breast cancer NCT03913234,
NCT02657343
EGFR Nazartinib (EGF816) EGFR mutant NSCLC NCT03333343
RAF Encorafenib, LXH254 NSCLC, BRAF
mutant melanoma
NCT02974725,
NCT03333343,
NCT04417621,
NCT02159066
MEK Binimetinib BRAF V600-dependent
advanced solid tumors,
melanoma
NCT01543698,
NCT02159066
PI3K Alpelisib Breast cancer with
PIK3CA mutation
NCT03439046
mTOR Everolimus Advanced dedifferentiated
liposarcoma, leiomyosarcoma,
glioma, astrocytoma,
glioblastoma,
endometrial carcinoma,
pancreatic cancer,
neuroendocrine tumors
NCT03114527,
NCT03355794,
NCT03834740,
NCT03008408,
NCT02985125,
NCT03070301
mTOR + inflammation Everolimus + dexamethasone ALL NCT03740334
SHP2 TNO155 Advanced solid tumors NCT04000529
AR AR antagonists:
bicalutamide,
enzalutamide
TNBC, metastatic
prostate carcinoma
NCT03090165,
NCT02555189
HDAC Belinostat TNBC, ovarian cancer NCT04315233
proliferating cells Standard chemotherapy Ovarian cancer, metastatic
solid tumors, soft tissue
sarcoma, hepatocellular
carcinoma
NCT03056833,
NCT03237390,
NCT03009201,
NCT02524119
Abemaciclib
Aromatase Letrozole, anastrozole,
exemestane
HR+ breast cancer,
metastatic breast
cancer, endometrial
cancer
NCT04256941,
NCT03425838,
NCT04227327,
NCT04393285,
NCT04305236,
NCT03643510,
NCT03675893,
NCT04352777,
NCT04293393,
NCT02057133
ER ER antagonists: fulvestrant Advanced breast cancer,
low-grade serous
ovarian cancer
NCT03227328,
NCT03531645,
NCT04158362,
NCT01394016
PD-1 Nivolumab,
pembrolizumab
Head and neck cancer, g
astroesophageal
cancer, NSCLC,
HR+ breast cancer
NCT04169074,
NCT03655444,
NCT03997448,
NCT02779751
ER + PD-L1 ER antagonists: fulvestrant Atezolizumab HR+ breast cancer, metastatic
breast cancer
NCT03280563
AKT + ER + PD-L1 Ipatasertib + ER
antagonists: fulvestrant
Atezolizumab HR+ breast cancer NCT03280563
PD-L1 LY3300054 Advanced solid tumors NCT02791334
HER2 Trastuzumab HER2+ metastatic
breast cancer
NCT04351230
Receptor tyrosine
kinases
Sunitinib Metastatic renal
cell carcinoma
NCT03905889
IGF-1/IGF-2 Xentuzumab HR+ breast cancer NCT03099174
VEGF-A Bevacizumab Glioblastoma NCT04074785
PI3K Copanlisib HR+ breast cancer, metastatic
breast cancer
NCT03939897
PI3K/mTOR LY3023414 Metastatic cancer NCT01655225
ERK1/2 LY3214996 tumors with ERK1/2
mutations, glioblastoma,
metastatic cancer
NCT04534283,
NCT04391595,
NCT02857270
Trilaciclib
Proliferating cells Chemotherapy SCLC: This trial evaluates the
potential clinical benefit of
trilaciclib in preventing
chemotherapy-induced
myelosuppression in patients
receiving chemotherapy
NCT04504513
Proliferating cells +
PD-L1
Carboplatin + etoposide Atezolizumab SCLC: This trial investigates the
potential clinical benefit of trilaciclib
in preserving the bone marrow and
the immune system, and enhancing
antitumor efficacy when
administered with chemotherapy
NCT03041311
Proliferating cells Topotecan SCLC: This trial investigates the
potential clinical benefit of
trilaciclib in preserving the
bone marrow and the immune
system, and enhancing the
antitumor efficacy of chemotherapy
when administered prior
to chemotherapy
NCT02514447
Proliferating cells Carboplatin + gemcitabine Metastatic TNBC: This study
investigates the potential
clinical benefit of trilaciclib in
preserving the bone marrow
and the immune system, and
enhancing the antitumor efficacy
of chemotherapy when administered
prior to chemotherapy
NCT02978716
Lerociclib
ER ER antagonist: fulvestrant HR+/HER2 metastatic
breast cancer
NCT02983071
EGFR Osimertinib EGFR mutant NSCLC NCT03455829
SHR6390
ER ER antagonist: fulvestrant HR+/HER2 recurrent/
metastatic breast cancer
NCT03481998
Aromatase Letrozole, anastrozole HR+/HER2 recurrent/
metastatic breast cancer
NCT03966898,
NCT03772353
EGFR/HER2 Pyrotinib HER2+ gastric cancer, HER2+
metastatic breast cancer
NCT04095390,
NCT03993964
AR AR antagonists: SHR3680 metastatic TNBC NCT03805399
PF-06873600
Endocrine therapy Single agent and then
in combination with
endocrine therapy
HR+/HER2 metastatic breast
cancer, ovarian and fallopian tube
cancer, TNBC and other tumors
NCT03519178
FCN-473c
Aromatase Letrozole ER+/HER2 advanced
breast cancer
NCT04488107

Expand for more

Table 4. Ongoing clinical trials testing new combinations with CDK4/6 inhibitors.
HR+, hormone receptor–positive; LHRH, luteinizing hormone–releasing hormone; ER+, estrogen receptor–positive; PD-1, programmed cell death protein 1; PD-L1, programmed cell death 1 ligand 1; AR+, androgen receptor–positive; TNBC, triple-negative breast cancer; EGFR, epidermal growth factor receptor; HER2+, human epidermal growth factor receptor 2–positive; FGFR, fibroblast growth factor receptor; IGFR, insulin-like growth factor receptor; VEGF, vascular endothelial growth factor receptor; PI3K, phosphoinositide 3-kinase; NSCLC, non–small-cell lung cancer; ALL, acute lymphoblastic leukemia; SCLC, small-cell lung cancer.
OPEN IN VIEWER

Resistance to CDK4/6 inhibitors

Although CDK4/6 inhibitors represent very effective agents in cancer treatment, nearly all patients eventually develop resistance and succumb to the disease. Moreover, a substantial fraction of tumors show intrinsic resistance to treatment with CDK4/6 inhibitors (Fig. 3).
Fig. 3. Mechanisms of cancer cell resistance to CDK4/6 inhibition.
Known mechanisms include loss of RB1, activation of pathways impinging on CycD-CDK4/6, amplification of the CDK4/6 genes and overexpression of CDK6 protein, activation of CycE-CDK2, and lysosomal sequestration of CDK4/6 inhibitors. Blank pieces of the puzzle denote additional mechanisms that remain to be discovered.
OPEN IN VIEWER
The best-documented mechanism of preexisting and acquired resistance is the loss of RB1 (7181140). Acquired RB1 loss has been detected in PDXs (141), in circulating tumor DNA (ctDNA) (142143), and in tumors from patients treated with CDK4/6 inhibitors (144145). However, RB1 mutations are likely subclonal and are seen in only 5 to 10% of patients (143145).
Increased expression of CDK6 was shown to underlie acquired resistance to CDK4/6 inhibitors. Amplification of the CDK6 gene and the resulting overexpression of CDK6 protein were found in abemaciclib-resistant ER+ breast cancer cells (146) and in ctDNA of patients with ER+ breast cancers that progressed during treatment with palbociclib plus endocrine therapy (147). Also, CDK4 gene amplification conferred insensitivity to CDK4/6 inhibition in GBM and sarcomas (148150), whereas overexpression of CDK4 protein was associated with resistance to endocrine therapy in HR+ breast cancers (79).
Resistant breast cancer cells can also up-regulate the expression of CDK6 through suppression of the TGF-β/SMAD4 pathway by the microRNA miR-432-5p. In this mechanism, exosomal expression of miR-432-5p mediates the transfer of the resistance phenotype between neighboring cell populations (151). Another mechanism of CDK6 up-regulation in ER+ breast cancers is the loss of FAT1, which represses CDK6 expression via the Hippo pathway. Loss of FAT1 triggers up-regulation of CDK6 expression by the Hippo pathway effectors TAZ and YAP. Moreover, genomic alterations in other components of the Hippo pathway, although rare, are also associated with reduced sensitivity to CDK4/6 inhibitors (81).
Genetic lesions that activate pathways converging on D-type cyclins can cause resistance to CDK4/6 inhibitors. These include (i) FGFR1/2 gene amplification or mutational activation, detected in ctDNA from patients with ER+ breast cancers that progressed upon treatment with palbociclib plus endocrine therapy (147); (ii) hyperactivation of the MAPK pathway in resistant prostate adenocarcinoma cells, possibly due to increased production of EGF by cancer cells (152); and (iii) increased secretion of FGF in palbociclib-resistant KRAS-mutant NSCLC cells, which stimulates FGFR1 signaling in an autocrine or paracrine fashion, resulting in activation of ERK1/2 and mTOR as well as up-regulation of D-cyclin, CDK6, and cyclin E expression (153). Analyses of longitudinal tumor biopsies from a melanoma patient revealed an activating mutation in the PIK3CA gene that conferred resistance to ribociclib plus MEK inhibitor treatment (154). It is possible that these lesions elevate the cellular levels of active cyclin D–CDK4/6 complexes, thereby increasing the threshold for CDK4/6 inhibition.
Formation of a noncanonical cyclin D1–CDK2 complex was shown to represent another mechanism of acquired CDK4/6 inhibitor resistance. Such a complex was observed in palbociclib-treated ER+ breast cancer cells and was implicated in overcoming palbociclib-induced cell cycle arrest (141). Also, depletion of AMBRA1 promoted the interaction of D-cyclins with CDK2, resulting in resistance to CDK4/6 inhibitors (2022); it remains to be seen whether this represents an intrinsic or acquired resistance mechanism in human tumors.
Genetic analyses revealed that activation of cyclin E can bypass the requirement for cyclin D–CDK4/6 in development and tumorigenesis (155156). Hence, it comes as no surprise that increased activity of cyclin E–CDK2 is responsible for a large proportion of intrinsic and acquired resistance to CDK4/6 inhibitors. Several different mechanisms can activate cyclin E–CDK2 kinase in resistant tumor cells: (i) Down-regulation of KIP/CIP inhibitors results in increased activity of cyclin E–CDK (54157). (ii) Loss of PTEN expression, which activates AKT signaling, leads to nuclear exclusion of p27KIP1. This in turn prevents access of p27KIP1 to CDK2, resulting in increased CDK2 kinase activity (144). (iii) Activation of the PI3K/AKT pathway causes decreased levels of p21CIP1. Co-treatment of melanoma PDXs with MDM2 inhibitors (which up-regulate p21CIP1 via p53) sensitized intrinsically resistant tumor cells to CDK4/6 inhibitors (158). (iv) Up-regulation of cyclin D1 levels triggers sequestration of KIP/CIP inhibitors from cyclin E–CDK2 to cyclin D–CDK4/6, thereby activating the former (158). (v) Amplification of the CCNE1 gene and increased levels of cyclin E1 protein result in elevated activity of E-CDK2 kinase (141). (vi) mTOR signaling has been shown to up-regulate cyclin E1 (and D1) in KRAS-mutated pancreatic cancer cells; CDK2 activity was essential for CDK4/6 inhibitor resistance in this setting (159). (vii) Up-regulation of PDK1 results in activation of the AKT pathway, which increases the expression of cyclins E and A and activates CDK2 (160). (viii) In CDK4/6 inhibitor–resistant melanoma cells, high levels of RNA-binding protein FXR1 increase translation of the amino acid transporter SLC36A1. Up-regulation of SLC36A1 expression activates mTORC1, which in turn increases CDK2 expression (161). All these lesions are expected to allow cell proliferation, despite CDK4/6 inhibition, as a consequence of the activation of the downstream cell cycle kinase CDK2.
The role for cyclin E–CDK2 in CDK4/6 inhibitor resistance has been confirmed in clinical trials. In patients with advanced ER+ breast cancer treated with palbociclib and letrozole or fulvestrant, the presence of proteolytically cleaved cytoplasmic cyclin E in tumor tissue conferred strongly shortened progression-free survival (71). Moreover, analyses of PALOMA-3 trial for patients with ER+ breast cancers revealed lower efficacy of palbociclib plus fulvestrant in patients displaying high cyclin E mRNA levels in metastatic biopsies (80). Amplification of the CCNE1 gene was detected in ctDNA of patients with ER+ breast cancers that progressed on palbociclib plus endocrine therapy (147). Also, amplification of the CCNE2 gene (encoding cyclin E2) was seen in a fraction of CDK4/6 inhibitor–resistant HR+ mammary carcinomas (145162).
Collectively, these analyses indicate that resistant cells may become dependent on CDK2 for cell cycle progression. Indeed, depletion of CDK2 or inhibition of CDK2 kinase activity in combination with CDK4/6 inhibitors blocked proliferation of CDK4/6 inhibitor–resistant cancer cells (111141158161). Recently, two CDK2-specific inhibitors, PF-07104091 (163) and BLU0298 (164), have been reported. PF-07104091 is now being tested in a phase 2 clinical trial in combination with palbociclib plus antiestrogens. Another recent study identified a novel compound, PF-3600, that inhibits CDK4/6 and CDK2 (165). PF3600 had potent antitumor effects against xenograft models of intrinsic and acquired resistance to CDK4/6 inhibition (165). A phase 2 clinical trial is currently evaluating this compound as a single agent and in combination with endocrine therapy in patients with HR+/HER2 breast cancer and other cancer types.
Whole-exome sequencing of 59 HR+/HER2 metastatic breast tumors from patients treated with CDK4/6 inhibitors and anti-estrogens revealed eight alterations that likely conferred resistance: RB1 loss; amplification of CCNE2 or AURKA; activating mutations or amplification of AKT1FGFR2, or ERBB2; activating mutations in RAS genes; and loss of ER expression. The frequent activation of AURKA (in 27% of resistant tumors) raises the possibility of combining CDK4/6 inhibitors with inhibitors of Aurora A kinase to overcome resistance (145).
In contrast to ER+ mammary carcinomas, TNBCs are overall resistant to CDK4/6 inhibition (45). A subset of TNBCs display high numbers of lysosomes, which causes sequestration of CDK4/6 inhibitors into the expanded lysosomal compartment, thereby preventing their action on nuclear CDK4/6. Preclinical studies revealed that lysosomotropic agents that reverse the lysosomal sequestration (such as chloroquine, azithromycin, or siramesine) render TNBC cells fully sensitive to CDK4/6 inhibition (71111). These observations now need to be tested in clinical trials for TNBC patients.

Outlook

Although D-cyclins and CDK4/6 were discovered 30 years ago, several aspects of cyclin D–CDK4/6 biology, such as their role in antitumor immunity, are only now starting to be appreciated. The full range of cyclin D–CDK4/6 functions in tumor cells remains unknown. It is likely that these kinases play a much broader role in cancer cells than is currently appreciated. Hence, the impact of CDK4/6 inhibition on various aspects of tumorigenesis requires further study. Also, treatment of patients with CDK4/6 inhibitors likely affects several aspects of host physiology, which may be relevant to cancer progression.
In the next years, we will undoubtedly witness the development and testing of new CDK4/6 inhibitors. Because activation of CDK2 represents a frequent CDK4/6 inhibitor resistance mechanism, compounds that inhibit CDK4/6 and CDK2 may prevent or delay the development of resistance. Conversely, selective compounds that inhibit CDK4 but not CDK6 may allow more aggressive dosing, as they are expected not to result in bone marrow toxicity caused by CDK6 inhibition. New, less basic CDK4/6 inhibitor compounds (111) may escape lysosomal sequestration and may be efficacious against resistant cancer types such as TNBC. Degrader compounds, which induce proteolysis of cyclin D rather than inhibit cyclin D–CDK4/6 kinase, may have superior properties, as they would extinguish both CDK4/6-dependent and -independent functions of D-cyclins in tumorigenesis. Moreover, dissolution of cyclin D–CDK4/6 complexes is expected to liberate KIP/CIP inhibitors, which would then inhibit CDK2. D-cyclins likely play CDK-independent functions in tumorigenesis—for example, by regulating gene expression (166). However, their role in tumor biology and the utility of targeting these functions for cancer treatment remain largely unexplored.
An important challenge will be to test and identify combinatorial treatments involving CDK4/6 inhibitors for the treatment of different tumor types. CDK4/6 inhibitors trigger cell cycle arrest of tumor cells and, in some cases, senescence. It will be essential to identify combination treatments that convert CDK4/6 inhibitors from cytostatic compounds to cytotoxic ones, which would unleash the killing of tumor cells. Genome-wide high-throughput screens along with analyses of mouse cancer models and PDXs will help to address this point. Another largely unexplored area of cyclin D–CDK4/6 biology is the possible involvement of these proteins in other pathologies, such as metabolic disorders. Research in this area may extend the use of CDK4/6 inhibitors to treatment of other diseases. All these unresolved questions ensure that CDK4/6 biology will remain an active area of basic, translational, and clinical research for several years to come.

CDK inhibitors and Breast Cancer

The U.S. Food and Drug Administration today granted accelerated approval to Ibrance (palbociclib) to treat advanced (metastatic) breast cancer inr postmenopausal women with estrogen receptor (ER)-positive, human epidermal growth factor receptor 2 (HER2)-negative metastatic breast cancer who have not yet received an endocrine-based therapy. It is to be used in combination with letrozole, another FDA-approved product used to treat certain kinds of breast cancer in postmenopausal women.

See Dr. Melvin Crasto’s blog posts on the announcement of approval of Ibrance (palbociclib) at

http://newdrugapprovals.org/2015/02/05/fda-approves-ibrance-for-postmenopausal-women-with-advanced-breast-cancer/

and about the structure and mechanism of action of palbociclib

http://newdrugapprovals.org/2014/01/05/palbociclib/

 

From the CancerNetwork at http://www.cancernetwork.com/aacr-2014/cdk-inhibitors-show-impressive-activity-advanced-breast-cancer

CDK Inhibitors Show Impressive Activity in Advanced Breast Cancer

News | April 08, 2014 | AACR 2014, Breast Cancer

By Anna Azvolinsky, PhD

Ibrance structure

 

Chemical structure of palbociclib

 

 

Palbociclib and LY2835219 are both cyclin-dependent kinase (CDK) 4/6 inhibitors. CDK4 and CDK6 are kinases that, together with cyclin D1, facilitate the transition of dividing cells from the G1 to the S (synthesis) phase of the cell cycle. Preclinical studies have shown that breast cancer cells rely on CDK4 and CDK6 for division and growth, and that selective CDK4/6 inhibitors can arrest the cells at this G1/S phase checkpoint.

The results of the phase II trial of palbociclib and phase I trial of LY2835219 both indicated that hormone receptor (HR)-positive disease appears to be the best marker to predict patient response.

LY2835219 Phase I Trial Demonstrates Early Activity

The CDK4/6 inhibitor LY2835219 has demonstrated early activity in heavily pretreated women with metastatic breast cancer. Nineteen percent of these women (9 out of 47) had a partial response and 51% (24 out of 47) had stable disease following monotherapy with the oral CDK4/6 inhibitor. Patients had received a median of seven prior therapies, and 75% had metastatic disease in the lung, liver, or brain. The median age of patients was 55 years.

All of the partial responses were in patients with HR-positive disease. The overall response rate for this patient subset was 25% (9 of 36 patients). Twenty of the patients with stable disease had HR-positive disease, with 13 patients having stable disease lasting 24 weeks or more.

Despite treatment, disease progression occurred in 23% of the patients.

These results were presented at a press briefing by Amita Patnaik, MD, associate director of clinical research at South Texas Accelerated Research Therapeutics in San Antonio, Texas, at the 2014 American Association for Cancer Research (AACR) Annual Meeting, held April 5–9, in San Diego.

The phase I trial of LY2835219 enrolled 132 patients with five different tumor types, including metastatic breast cancer. Patients received 150-mg to 200-mg doses of the oral drug every 12 hours.

The overall disease control rate was 70% for all patients and 81% among the 36 HR-positive patients.

The median progression-free survival (PFS) was 5.8 months for all patients and 9.1 months for HR-positive patients. Patnaik noted that the median PFS is still a moving target, as 18 patients, all with HR-positive disease, remain on therapy.

“The data are rather encouraging for a very heavily pretreated patient population,” said Patnaik during the press briefing.

Even though the trial was not designed to compare efficacy based on breast cancer subpopulations, the results in HR-positive tumors are particularly encouraging, according to Patnaik.

Common adverse events thought to be treatment-related were diarrhea, nausea, fatigue, vomiting, and neutropenia. These adverse events occurred in 5% or less of patients at grade 3 or 4 toxicity, except neutropenia, which occurred as a grade 3 or 4 toxicity in 11% of patients. Patnaik noted during the press briefing that the neutropenia was uncomplicated and did not result in discontinuation of therapy by any of the patients.

Palbociclib Phase II Data “Impressive”

The addition of the oral CDK4/6 inhibitor palbociclib resulted in an almost doubling of PFS in first-line treatment of postmenopausal metastatic breast cancer patients with HR-positive disease compared with a control population. The patients in this trial were not previously treated for their metastatic breast cancer, unlike the patient population in the phase I LY2835219 trial.

Patients receiving the combination of palbociclib at 125 mg once daily plus letrozole at 2.5 mg once daily had a median PFS of 20.2 months compared with a median of 10.2 months for patients treated with letrozole alone (hazard ratio = 0.488; P = .0004).

Richard S. Finn, MD, assistant professor of medicine at the University of California, Los Angeles, presented the data from the phase II PALOMA-1 trial at a press briefing at the AACR Annual Meeting.

A total of 165 patients were randomized 1:1 to either the experimental arm or control arm.

Forty-three percent of patients in the combination arm had an objective response compared with 33% of patients in the control arm.

Overall survival (OS), a secondary endpoint in this trial, was encouraging but the results are still preliminary, said Finn during the press briefing. The median OS was 37.5 months in the palbociclib arm compared with 33.3 months in the letrozole alone arm (P = .21). Finn noted that long-term follow-up is necessary to establish the median OS. “This first look of the survival data is encouraging. This is a front-line study, and it is encouraging that there is early [separation] of the curves,” he said.

No new toxicities were reported since the interim trial results. Common adverse events included leukopenia, neutropenia, and fatigue. The neutropenia could be quickly resolved and was uncomplicated and not accompanied by fever, said Finn.

Palbociclib is currently being tested in two phase III clinical trials: The PALOMA-3 trial is testing the combination of palbociclib with letrozole and fulvestrant in late-stage metastatic breast cancer patients who have failed endocrine therapy. The PENELOPE-B trial is testing palbociclib in combination with standard endocrine therapy in HR-positive breast cancer patients with residual disease after neoadjuvant chemotherapy and surgery.

References

  1. Patnaik A, Rosen LS, Tolaney SM, et al. Clinical activity of LY2835219, a novel cell cycle inhibitor selective for CDK4 and CDK6, in patients with metastatic breast cancer. American Association for Cancer Research Annual Meeting 2014; April 5–9, 2014; San Diego. Abstr CT232.
  2. Finn RS, Crown JP, Lang I, et al. Final results of a randomized phase II study of PD 0332991, a cyclin-dependent kinase (CDK)-4/6 inhibitor, in combination with letrozole vs letrozole alone for first-line treatment of ER+/HER2-advanced breast cancer (PALOMA-1; TRIO-18). American Association for Cancer Research Annual Meeting 2014; April 5–9, 2014; San Diego. Abstr CT101.

– See more at: http://www.cancernetwork.com/aacr-2014/cdk-inhibitors-show-impressive-activity-advanced-breast-cancer#sthash.f29smjxi.dpuf

 

The Cell Cycle and Anti-Cancer Targets

 

graph_cell_cycle

 

From Cell Cycle in Cancer: Cyclacel Pharmaceuticals™ (note dotted arrows show inhibition of steps e.g. p21, p53)

For a nice video slideshow explaining a bit more on cyclins and the cell cycle please see video below:

 

Cell Cycle. 2012 Nov 1; 11(21): 3913.

doi:  10.4161/cc.22390

PMCID: PMC3507481

Cyclin-dependent kinase 4/6 inhibition in cancer therapy

Neil Johnson and Geoffrey I. Shapiro*

See the article “Therapeutic response to CDK4/6 inhibition in breast cancer defined by ex vivo analyses of human tumors” in volume 11 on page 2756.

See the article “CDK4/6 inhibition antagonizes the cytotoxic response to anthracycline therapy” in volume 11 on page 2747.

This article has been cited by other articles in PMC.

Cyclin-dependent kinases (CDKs) drive cell cycle progression and control transcriptional processes. The dysregulation of multiple CDK family members occurs commonly in human cancer; in particular, the cyclin D-CDK4/6-retinoblastoma protein (RB)-INK4 axis is universally disrupted, facilitating cancer cell proliferation and prompting long-standing interest in targeting CDK4/6 as an anticancer strategy. Most agents that have been tested inhibit multiple cell cycle and transcriptional CDKs and have carried toxicity. However, several selective and potent inhibitors of CDK4/6 have recently entered clinical trial. PD0332991, the first to be developed, resulted from the introduction of a 2-aminopyridyl substituent at the C2-position of a pyrido(2,3-d)pyrimidin-7-one backbone, affording exquisite selectivity toward CDK4/6.1 PD0332991 arrests cells in G1 phase by blocking RB phosphorylation at CDK4/6-specfic sites and does not inhibit the growth of RB-deficient cells.2 Phase I studies conducted in patients with advanced RB-expressing cancers demonstrated mild side effects and dose-limiting toxicities of neutropenia and thrombocytopenia, with prolonged stable disease in 25% of patients.3,4 In cyclin D1-translocated mantle cell lymphoma, PD0332991 extinguished CDK4/6 activity in patients’ tumors, resulting in markedly reduced proliferation, and translating to more than 1 year of stability or response in 5 of 17 cases.5

Two recent papers from the Knudsen laboratory make several important observations that will help guide the continued clinical development of CDK4/6 inhibitors. In the study by Dean et al., surgically resected patient breast tumors were grown on a tissue culture matrix in the presence or absence of PD0332991. Crucially, these cultures retained associated stromal components known to play important roles in cancer pathogenesis and therapeutic sensitivities, as well as key histological and molecular features of the primary tumor, including expression of ER, HER2 and Ki-67. Similar to results in breast cancer cell lines,6 the authors demonstrate that only RB-positive tumors have growth inhibition in response to PD0332991, irrespective of ER or HER2 status, while tumors lacking RB were completely resistant. This result underscores RB as the predominant target of CDK4/6 in breast cancer cells and the primary marker of drug response in primary patient-derived tumors. As expected, RB-negative tumors routinely demonstrated robust expression of p16INK4A; however, p16INK4A expression was not always a surrogate marker for RB loss, supporting the importance of direct screening of tumors for RB expression to select patients appropriate for CDK4/6 inhibitor clinical trials.

In the second study, McClendon et al. investigated the efficacy of PD0332991 in combination with doxorubicin in triple-negative breast cancer cell lines. Again, RB functionality was paramount in determining response to either PD0332991 monotherapy or combination treatment. In RB-deficient cancer cells, CDK4/6 inhibition had no effect in either instance. However, in RB-expressing cancer cells, CDK4/6 inhibition and doxorubicin provided a cooperative cytostatic effect, although doxorubicin-induced cytotoxicity was substantially reduced, assessed by markers for mitotic catastrophe and apoptosis. Additionally, despite cytostatic cooperativity, CDK4/6 inhibition maintained the viability of RB-proficient cells in the presence of doxorubicin, which repopulated the culture after removal of drug. These results reflect previous data demonstrating that ectopic expression of p16INK4A can protect cells from the lethal effects of DNA damaging and anti-mitotic chemotherapies.7 Similar results have been reported in MMTV-c-neu mice bearing RB-proficient HER2-driven tumors, where PD0332991 compromised carboplatin-induced regressions,8 suggesting that DNA-damaging treatments should not be combined concomitantly with CDK4/6 inhibition in RB-proficient tumors.

To combine CDK4/6 inhibition with cytotoxics, sequential treatment may be considered, in which CDK4/6 inhibition is followed by DNA damaging chemotherapy; cells relieved of G1 arrest may synchronously enter S phase, where they may be most susceptible to agents disrupting DNA synthesis. Release of myeloma cells from a prolonged PD0332991-mediated G1 block leads to S phase synchronization; interestingly, all scheduled gene expression is not completely restored (including factors critical to myeloma survival such as IRF4), further favoring apoptotic responses to cytotoxic agents.9 Furthermore, in RB-deficient tumors, CDK4/6 inhibitors may be used to maximize the therapeutic window between transformed and non-transformed cells treated with chemotherapy. In contrast to RB-deficient cancer cells, RB-proficient non-transformed cells arrested in G1 in response to PD0332991 are afforded protection from DNA damaging agents, thereby reducing associated toxicities, including bone marrow suppression.8

In summary, the current work provides evidence for RB expression as a determinant of response to CDK4/6 inhibition in primary tumors and highlights the complexity of combining agents targeting the cell cycle machinery with DNA damaging treatments.

Go to:

Notes

Dean JL, McClendon AK, Hickey TE, Butler LM, Tilley WD, Witkiewicz AK, Knudsen ES. Therapeutic response to CDK4/6 inhibition in breast cancer defined by ex vivo analyses of human tumors Cell Cycle 2012 11 2756 61 doi: 10.4161/cc.21195.

McClendon AK, Dean JL, Rivadeneira DB, Yu JE, Reed CA, Gao E, Farber JL, Force T, Koch WJ, Knudsen ES. CDK4/6 inhibition antagonizes the cytotoxic response to anthracycline therapy Cell Cycle 2012 11 2747 55 doi: 10.4161/cc.21127.

Go to:

Footnotes

Previously published online: www.landesbioscience.com/journals/cc/article/22390

Go to:

References

  1. Toogood PL, et al. J Med Chem. 2005;48:2388–406. doi: 10.1021/jm049354h. [PubMed] [Cross Ref]
  2. Fry DW, et al. Mol Cancer Ther. 2004;3:1427–38. [PubMed]
  3. Flaherty KT, et al. Clin Cancer Res. 2012;18:568–76. doi: 10.1158/1078-0432.CCR-11-0509. [PubMed] [Cross Ref]
  4. Schwartz GK, et al. Br J Cancer. 2011;104:1862–8. doi: 10.1038/bjc.2011.177. [PMC free article] [PubMed] [Cross Ref]
  5. Leonard JP, et al. Blood. 2012;119:4597–607. doi: 10.1182/blood-2011-10-388298. [PubMed] [Cross Ref]
  6. Dean JL, et al. Oncogene. 2010;29:4018–32. doi: 10.1038/onc.2010.154. [PubMed] [Cross Ref]
  7. Stone S, et al. Cancer Res. 1996;56:3199–202. [PubMed]
  8. Roberts PJ, et al. J Natl Cancer Inst. 2012;104:476–87. doi: 10.1093/jnci/djs002. [PMC free article] [PubMed] [Cross Ref]
  9. Huang X, et al. Blood. 2012;120:1095–106. doi: 10.1182/blood-2012-03-415984. [PMC free article] [PubMed] [Cross Ref]

Cell Cycle. 2012 Jul 15; 11(14): 2756–2761.

doi:  10.4161/cc.21195

PMCID: PMC3409015

Therapeutic response to CDK4/6 inhibition in breast cancer defined by ex vivo analyses of human tumors

Jeffry L. Dean, 1 , 2 A. Kathleen McClendon, 1 , 2 Theresa E. Hickey, 3 Lisa M. Butler, 3 Wayne D. Tilley, 3 Agnieszka K. Witkiewicz, 4 , 2 ,* and Erik S. Knudsen 1 , 2 ,*

Author information ► Copyright and License information ►

See commentary “Cyclin-dependent kinase 4/6 inhibition in cancer therapy” in volume 11 on page 3913.

This article has been cited by other articles in PMC.

Go to:

Abstract

To model the heterogeneity of breast cancer as observed in the clinic, we employed an ex vivo model of breast tumor tissue. This methodology maintained the histological integrity of the tumor tissue in unselected breast cancers, and importantly, the explants retained key molecular markers that are currently used to guide breast cancer treatment (e.g., ER and Her2 status). The primary tumors displayed the expected wide range of positivity for the proliferation marker Ki67, and a strong positive correlation between the Ki67 indices of the primary and corresponding explanted tumor tissues was observed. Collectively, these findings indicate that multiple facets of tumor pathophysiology are recapitulated in this ex vivo model. To interrogate the potential of this preclinical model to inform determinants of therapeutic response, we investigated the cytostatic response to the CDK4/6 inhibitor, PD-0332991. This inhibitor was highly effective at suppressing proliferation in approximately 85% of cases, irrespective of ER or HER2 status. However, 15% of cases were completely resistant to PD-0332991. Marker analyses in both the primary tumor tissue and the corresponding explant revealed that cases resistant to CDK4/6 inhibition lacked the RB-tumor suppressor. These studies provide important insights into the spectrum of breast tumors that could be treated with CDK4/6 inhibitors, and defines functional determinants of response analogous to those identified through neoadjuvant studies.

Keywords: ER, PD0332991, breast cancer, cell cycle, ex vivo

Go to:

Introduction

Breast cancer is a highly heterogeneous disease.14 Such heterogeneity is known to influence patient response to both standard of care and experimental therapeutics. In regards to biomarker-driven treatment of breast cancers, it was initially recognized that the presence of the estrogen receptor α (ER) in a fraction of breast cancer cells was associated with the response to tamoxifen and similar anti-estrogenic therapies.5,6 Since this discovery, subsequent marker analyses and gene expression profiling studies have further divided breast cancer into a series of distinct subtypes that harbor differing and often divergent therapeutic sensitivities.13 While clearly important in considering the use of several current standard of care therapies, these markers, or molecular sub-types, do not necessarily predict the response to new therapeutic approaches that are currently undergoing clinical development. Thus, there is the continued need for functional analyses of drug response and the definition of new markers that can be used to direct treatment strategies.

Currently, all preclinical cancer models are associated with specific limitations. It is well known that cell culture models lack the tumor microenvironment known to have a significant impact on tumor biology and therapeutic response.79 Xenograft models are dependent on the host response for the engraftment of tumor cells in non-native tissues, which do not necessarily recapitulate the nuances of complex tumor milieu.10 In addition, genetically engineered mouse models, while enabling the tumor to develop in the context of the host, can develop tumors that do not mirror aspects of human disease.10 Furthermore, it remains unclear whether any preclinical model truly represents the panoply of breast cancer subtypes that are observed in the clinic. Herein, we utilized a primary human tumor explant culture approach to interrogate drug response, as well as specific determinants of therapeutic response, in an unselected series of breast cancer cases.

Cell Cycle. 2012 Jul 15; 11(14): 2747–2755.

doi:  10.4161/cc.21127

PMCID: PMC3409014

CDK4/6 inhibition antagonizes the cytotoxic response to anthracycline therapy

  1. Kathleen McClendon, 1 , † Jeffry L. Dean, 1 , † Dayana B. Rivadeneira, 1 Justine E. Yu, 1 Christopher A. Reed, 1 Erhe Gao, 2 John L. Farber, 3 Thomas Force, 2 Walter J. Koch, 2 and Erik S. Knudsen 1 ,*

Author information ► Copyright and License information ►

See commentary “Cyclin-dependent kinase 4/6 inhibition in cancer therapy” in volume 11 on page 3913.

This article has been cited by other articles in PMC.

Go to:

Abstract

Triple-negative breast cancer (TNBC) is an aggressive disease that lacks established markers to direct therapeutic intervention. Thus, these tumors are routinely treated with cytotoxic chemotherapies (e.g., anthracyclines), which can cause severe side effects that impact quality of life. Recent studies indicate that the retinoblastoma tumor suppressor (RB) pathway is an important determinant in TNBC disease progression and therapeutic outcome. Furthermore, new therapeutic agents have been developed that specifically target the RB pathway, potentially positioning RB as a novel molecular marker for directing treatment. The current study evaluates the efficacy of pharmacological CDK4/6 inhibition in combination with the widely used genotoxic agent doxorubicin in the treatment of TNBC. Results demonstrate that in RB-proficient TNBC models, pharmacological CDK4/6 inhibition yields a cooperative cytostatic effect with doxorubicin but ultimately protects RB-proficient cells from doxorubicin-mediated cytotoxicity. In contrast, CDK4/6 inhibition does not alter the therapeutic response of RB-deficient TNBC cells to doxorubicin-mediated cytotoxicity, indicating that the effects of doxorubicin are indeed dependent on RB-mediated cell cycle control. Finally, the ability of CDK4/6 inhibition to protect TNBC cells from doxorubicin-mediated cytotoxicity resulted in recurrent populations of cells specifically in RB-proficient cell models, indicating that CDK4/6 inhibition can preserve cell viability in the presence of genotoxic agents. Combined, these studies suggest that while targeting the RB pathway represents a novel means of treatment in aggressive diseases such as TNBC, there should be a certain degree of caution when considering combination regimens of CDK4/6 inhibitors with genotoxic compounds that rely heavily on cell proliferation for their cytotoxic effects.

 

 

Click on Video Link for Dr. Tolaney slidepresentation of recent data with CDK4/6 inhibitor trial results https://youtu.be/NzJ_fvSxwGk

Audio and slides for this presentation are available on YouTube: http://youtu.be/NzJ_fvSxwGk

Sara Tolaney, MD, MPH, a breast oncologist with the Susan F. Smith Center for Women’s Cancers at Dana-Farber Cancer Institute, gives an overview of phase I clinical trials and some of the new drugs being tested to treat breast cancer. This talk was originally given at the Metastatic Breast Cancer Forum at Dana-Farber on Oct. 5, 2013.

A great article on current clinical trials and explanation of cdk inhibitors by Sneha Phadke, DO; Alexandra Thomas, MD at the site OncoLive

 

http://www.onclive.com/publications/contemporary-oncology/2014/november-2014/targeting-cell-cycle-progression-cdk46-inhibition-in-breast-cancer/1

 

cdk4/6 inhibitor Ibrance Has Favorable Toxicity and Adverse Event Profile

 

As discussed in earlier posts and the Introduction to this chapter on Cytotoxic Chemotherapeutics, most anti-cancer drugs developed either to target DNA, DNA replication, or the cell cycle usually have similar toxicity profile which can limit their therapeutic use. These toxicities and adverse events usually involve cell types which normally exhibit turnover in the body, such as myeloid and lymphoid and granulocytic series of blood cells, epithelial cells lining the mucosa of the GI tract, as well as follicular cells found at hair follicles. This understandably manifests itself as common toxicities seen with these types of agents such as the various cytopenias in the blood, nausea vomiting diarrhea (although there are effects on the chemoreceptor trigger zone), and alopecia.

It was felt that the cdk4/6 inhibitors would show serious side effects similar to other cytotoxic agents and this definitely may be the case as outlined below:

(Side effects of palbociclib) From navigatingcancer.com

Palbociclib may cause side effects. Tell your doctor if any of these symptoms are severe or do not go away:

  • nausea
  • diarrhea
  • vomiting
  • decreased appetite
  • tiredness
  • numbness or tingling in your arms, hands, legs, and feet
  • sore mouth or throat
  • unusual hair thinning or hair loss

Some side effects can be serious. If you experience any of these symptoms, call your doctor immediately or get emergency medical treatment:

  • fever, chills, or signs of infection
  • shortness of breath
  • sudden, sharp chest pain that may become worse with deep breathing
  • fast, irregular, or pounding heartbeat
  • rapid breathing
  • weakness
  • unusual bleeding or bruising
  • nosebleeds

The following is from FDA Drug Trials Snapshot of Ibrance™:

 

See PDF on original submission and CDER review

original FDA Ibrance submission

original FDA Ibrance submission

CDER Review Ibrance

CDER Review Ibrance

 

4.3 Preclinical Pharmacology/Toxicology

 

For full details, please see Pharmacology/Toxicology review by Dr. Wei Chen The nonclinical studies adequately support the safety of oral administration of palbociclib for the proposed indication and the recommendation from the team is for approval. Non-clinical studies of palbociclib included safety pharmacology studies, genotoxicity

studies, reproductive toxicity studies, pharmacokinetic studies, toxicokinetic studies and repeat-dose general toxicity studies which were conducted in rats and dogs. The pivotal toxicology studies were conducted in compliance with Good Laboratory Practice regulation.

 

Pharmacology:

As described above, palbociclib is an inhibitor of CDK4 and CDK6. Palbociclib modulates downstream targets of CDK4 and CDK6 in vitro and induces G1 phase cell cycle arrest and therefore acts to inhibit DNA synthesis and cell proliferation. Combination of palbociclib with anti-estrogen agents demonstrated synergistic inhibition

of cell proliferation in ER+ breast cancer cells. Palbociclib showed anti-tumor efficacy in animal tumor model studies. Safety pharmacology studies with palbociclib demonstrated adverse effects on both the respiratory and cardiovascular function of dogs at a dose of 125mg/day (four times and 50-times the human clinical exposure

respectively) based on mean unbound Cmax.

 

General toxicology:

Palbociclib was studied in single dose toxicity studies and repeated dose studies in rats and dogs. Adverse effects in the bone marrow, lymphoid tissues, and male reproductive organs were observed at clinically relevant exposures. Partial to complete reversibility of toxicities to the hematolymphopoietic and male reproductive systems was demonstrated following a recovery period (4-12 weeks), with the exception of the male reproductive organ findings in dogs. Gastrointestinal, liver, kidney, endocrine/metabolic (altered glucose metabolism), respiratory, ocular, and adrenal effects were also seen.

 

Genetic toxicology:

Palbociclib was evaluated for potential genetic toxicity in in vitro and in vivo studies. The Ames bacterial mutagenicity assay in the presence or absence of metabolic activation demonstrated non-mutagenicity. In addition, palbociclib did not induce chromosomal aberrations in cultured human peripheral blood lymphocytes in the presence or absence of metabolic activation. Palbociclib was identified as aneugenic based on kinetochore analysis of micronuclei formation in an In vitro assay in CHO-WBL cells. In addition, palbociclib was shown to induce micronucleus formation in male rats at doses 100

mg/kg/day (10x human exposure at the therapeutic dose) in an in vivo rat micronucleus assay.

 

Reproductive toxicology: No effects on estrous cycle and no reproductive toxicities were noticed in standard assays.

 

Pharmacovigilance (note please see PDF for more information)

Deaths Associated With Trials: Although a few deaths occurred during some trials no deaths were attributed to the drug.

Non-Serious Adverse Events:

(note a reviewers comment below concerning incidence of pulmonary embolism is a combination trial with letrazole)

 

fda ibrance reviewers SAE comment

 

Other article in this Open Access Journal on Cell Cycle and Cancer Include:

 

Tumor Suppressor Pathway, Hippo pathway, is responsible for Sensing Abnormal Chromosome Numbers in Cells and Triggering Cell Cycle Arrest, thus preventing Progression into Cancer

Nonhematologic Cancer Stem Cells [11.2.3]

New methods for Study of Cellular Replication, Growth, and Regulation

Multiple Lung Cancer Genomic Projects Suggest New Targets, Research Directions for Non-Small Cell Lung Cancer

Proteomics, Metabolomics, Signaling Pathways, and Cell Regulation: a Compilation of Articles in the Journal http://pharmaceuticalintelligence.com

In Focus: Targeting of Cancer Stem Cells

 

 

 

 

 

 

 

 

Read Full Post »

Can IntraTumoral Heterogeneity Be Thought of as a Mechanism of Resistance?

Curator/Reporter: Stephen J. Williams, Ph.D.

Therapeutic resistance remains one of the most challenging problems for the oncologist, despite the increase of new therapeutics in the oncologist’s toolkit. As new targeted therapies are developed, and new novel targets are investigated as potential therapies, especially cytostatic therapies which it has become evident our understanding of chemoresistance is expanding beyond mechanisms to circumvent a drug’s pharmacologic mechanism of action (i.e. increased DNA repair and cisplatin) or pharmacokinetic changes (i.e. increased efflux by acquisition of a MDR phenotype).

In a talk at the 2015 AACR National Meeting, Dr. Charles Swanton discusses the development of tumor heterogeneity in the light of developing, or acquired, drug resistance. Chemoresistance is either categorized as acquired resistance (where resistance develops upon continued exposure to drug) or inherent resistance (related to a tumor being refractory or unresponsive to drug). Dr Swanton discusses findings where development of this heterogeneity (discussed here in a posting on Issues in Personalized Medicine in Cancer: Intratumor Heterogeneity and Branched Evolution Revealed by Multiregion Sequencing) and here (Notes On Tumor Heterogeneity: Targets and Mechanisms, from the 2015 AACR Meeting in Philadelphia PA) on recent findings on Branched Chain Heterogeneity) is resulting in clones resistant to the initial drug treatment.

To recount a bit of background I list the overall points of the one of previous posts on tumor heterogeneity (and an interview with Dr. Charles Swanton) are as follows:

Multiple biopsies of primary tumor and metastases are required to determine the full mutational landscape of a patient’s tumor

The intratumor heterogeneity will have an impact on the personalized therapy strategy for the clinician

Metastases arising from primary tumor clones will have a greater genomic instability and mutational spectrum than the tumor from which it originates

Tumors and their metastases do NOT evolve in a linear path but have a branched evolution and would complicate biomarker development and the prognostic and resistance outlook for the patient

 

The following is a curation of various talks and abstracts from the 2015 AACR National Meeting in Philadelphia on effects of clonal evolution and intratumoral heterogeneity of a tumor with respect to development of chemoresistance. As this theory of heterogeneity and clonal evolution is particularly new I attempted to present all works (although apologize for the length upfront) to forgo bias and so the reader may extract any information pertinent to their clinical efforts and research. However I will give a brief highlight summary below:

 

From the 2015 AACR National Meeting in Philadelphia

 

 

 

 

PresentationNumber:NGO2

Presentation Title: Polyclonal and heterogeneous resistance to targeted therapy in leukemia
Presentation Time: Monday, Apr 20, 2015, 10:40 AM -10:55 AM
Location: Room 201, Pennsylvania Convention Center
Author Block: Catherine C. Smith, Amy Paguirigan, Chen-Shan Chin, Michael Brown, Wendy Parker, Mark J. Levis, Alexander E. Perl, Kevin Travers, Corynn Kasap, Jerald P. Radich, Susan Branford, Neil P. Shah. University of California, San Francisco, CA, Fred Hutchinson Cancer Research Center, Seattle, WA, Pacific Biosciences, Menlo Park, CA, Royal Adelaide Hospital, Adelaide, Australia, Sidney Kimmel Comprehensive Cancer Center at Johns Hopkins, Baltimore, MD, Abramson Cancer Center of the University of Pennsylvania, Philadelphia, PA, University of California, San Francisco, CA
Abstract Body: Genomic studies in solid tumors have revealed significant branching intratumoral clonal genetic heterogeneity. Such complexity is not surprising in solid tumors, where sequencing studies have revealed thousands of mutations per tumor genome. However, in leukemia, the genetic landscape is considerably less complex. Chronic myeloid leukemia (CML) is the human malignancy most definitively linked to a single genetic lesion, the BCR-ABL gene fusion. Genome wide sequencing of acute myeloid leukemia (AML) has revealed that AML is the most genetically straightforward of all extensively sequenced adult cancers to date, with an average of 13 coding mutations and 3 or less clones identified per tumor.
In CML, tyrosine kinase inhibitors (TKIs) of BCR-ABL have resulted in high rates of remission. However, despite excellent initial response rates with TKI monotherapy, patients still relapse, including virtually all patients with Philadelphia-positive acute lymphoblastic leukemia and blast crisis CML. Studies of clinical resistance highlight BCR-ABL as the sole genetic driver in CML as secondary kinase domain (KD) mutations that prevent drug binding are the predominant mechanism of relapse on BCR-ABL TKIs.
In AML, a more diverse panel of disease-defining genetic mutations has been uncovered. However, in individual patients, a single oncogene can still drive disease. This is the case in FLT3 mutant AML, in which the investigational FLT3 TKI quizartinib achieved an initial response rate of ~50% in relapsed/refractory AML patients with activating FLT3 internal tandem duplication (ITD) mutations, though most patients eventually relapsed. Confirming the importance of FLT3 in disease maintenance, we showed that 8 of 8 patients who relapsed on quizartinib did so due to acquired drug-resistant FLT3 KD mutations.
Studies in CML have revealed that sequential TKI therapy is associated with additional complexity where multiple mutations can coexist separately in an individual patient (“polyclonality”) or in tandem on a single allele (“compound mutations”). In AML, we observed polyclonal FLT3-ITD KD mutations in 2 of 8 patients examined in our initial study of quizartinib resistance.
In light of the polyclonal KD mutations observed in CML and AML at the time of TKI relapse, we undertook next generation sequencing studies to determine the true genetic complexity in CML and AML patients at the time of relapse on targeted therapy. We used Pacific Biosciences RS Single Molecule Real Time (SMRT) third generation sequencing technology to sequence the entire ABL KD or the entire FLT3 juxtamembrane and KD on a single strand of DNA. Using this method, we assessed a total of 103 samples from 79 CML patients on ABL TKI therapy and 36 paired pre-treatment and relapse samples from 18 FLT3-ITD+ AML patients who responded to investigational FLT3 TKI therapy.
In CML, using SMRT sequencing, we detected all mutations previously detected by direct sequencing. Of samples in which multiple mutations were detectable by direct sequencing, 85% had compound mutant alleles detectable in a variety of combinations. Compound mutant alleles were comprised of both dominant and minor mutations, some which were not detectable by direct sequencing. In the most complex case, 12 individual mutant alleles comprised of 7 different mutations were identified in a single sample.
For 12 CML patients, we interrogated longitudinal samples (2-4 time points per patient) and observed complex clonal relationships with highly dynamic shifts in mutant allele populations over time. We detected compound mutations arising from ancestral single mutant clones as well as parallel evolution of de novo polyclonal and compound mutations largely in keeping with what would be expected to cause resistance to the second generation TKI therapy received by that patient.
We used a phospho-flow cytometric technique to assesses the phosphorylation status of the BCR-ABL substrate CRKL in as a method to test the ex vivo biochemical responsiveness of individual mutant cell populations to TKI therapy and assess functional cellular heterogeneity in a given patient at a given timepoint. Using this technique, we observed co-existing cell populations with differential ex vivo response to TKI in 2 cases with detectable polyclonal mutations. In a third case, we identified co-existence of an MLL-AF9 containing cell population that retained the ability to modulate p-CRKL in response to BCR-ABL TKIs along with a BCR-ABL containing only population that showed biochemical resistance to all TKIs, suggesting the co-existence of BCR-ABL independent and dependent resistance in a single patient.
In AML, using SMRT sequencing, we identified acquired quizartinib resistant KD mutations on the FLT3-ITD (ITD+) allele of 9 of 9 patients who relapsed after response to quizartinib and 4 of 9 patients who relapsed after response to the investigational FLT3 inhibitor, PLX3397. In 4 cases of quizartinib resistance and 3 cases of PLX3397 resistance, polyclonal mutations were observed, including 7 different KD mutations in one patient with PLX3397 resistance. In 7 quizartinib-resistant cases and 3 PLX3397-resistant cases, mutations occurred at the activation loop residue D835. When we examined non-ITD containing (ITD-) alleles, we surprisingly uncovered concurrent drug-resistant FLT3 KD mutations on ITD- alleles in 7 patients who developed quizartinib resistance and 4 patients with PLX3397 resistance. One additional PLX3397-resistant patient developed a D835Y mutation only in ITD- alleles at the time of resistance, suggesting selection for a non-ITD containing clone. All of the individual substitutions found on ITD- alleles were the same substitutions identified on ITD+ alleles for each individual patient.
Given that the same individual mutations found on ITD- alleles were also found on ITD+ alleles, we sought to determine whether these mutations were found in the same cell or were indicative of polyclonal blast populations in each patient. To answer this question, we performed single cell sorting of viably frozen blasts from 3 quizartinib-resistant patients with D835 mutations identified at the time of relapse and genotyped single cells for the presence or absence of ITD and D835 mutations. This analysis revealed striking genetic heterogeneity. In 2/3 cases, polyclonal D835 mutations were found in both ITD+ and ITD- cells. In all cases, FLT3-ITD and D835 mutations were found in both heterozygous and homozygous combinations. Most surprisingly, in all 3 patients, approximately 30-40% of FLT3-ITD+ cells had no identified quizartinib resistance-causing FLT3 KD mutation to account for resistance, suggesting the presence of non-FLT3 dependent resistance in all patients.
To determine that ITD+ cells lacking FLT3 KD mutations observed in patients relapsed on quizartinib are indeed consistent with leukemic blasts functionally resistant to quizartinib and do not instead represent a population of differentiated or non-proliferating cells, we utilized relapse blasts from another patient who initially achieved clearance of bone marrow blasts on quizartinib and developed a D835Y mutation at relapse. We performed a colony assay in the presence of 20nM quizartinib. As expected, this dose of quizartinib was unable to suppress the colony-forming ability of blasts from this relapsed patient when compared to DMSO treatment. Genotyping of individual colonies grown from this relapse sample in the presence of 20nM quizartinib again showed remarkable genetic heterogeneity, including ITD+ and ITD- colonies with D835Y mutations in homozygous and heterozygous combinations as well as ITD+ colonies without D835Y mutations, again suggesting the presence of blasts with non-FLT3 dependent resistance. Additionally, 4 colonies with no FLT3 mutations at all were identified in this sample, suggesting the presence of a quizartinib-resistant non-FLT3 mutant blast population. To see if we could identify specific mechanisms of off-target resistance, we performed targeted exome sequencing 33-AML relevant genes from relapse and pre-treatment DNA from all four patients and detected no new mutations in any genes other than FLT3 acquired at the time of disease relapse. Clonal genetic heterogeneity is not surprising in solid tumors, where multiple driver mutations frequently occur, but in CML and FLT3-ITD+ AML, where disease has been shown to be exquisitely dependent on oncogenic driver mutations, our studies suggest a surprising amount of clonal diversity. Our findings show that clinical TKI resistance in these diseases is amazingly intricate on the single allele level and frequently consists of both polyclonal and compound mutations that give rise to an complicated pool of TKI-resistant alleles that can change dynamically over time. In addition, we demonstrate that cell populations with off-target resistance can co-exist with other TKI-resistant populations, underscoring the emerging complexity of clinical TKI resistance. Such complexity argues strongly that monotherapy strategies in advanced CML and AML may be ultimately doomed to fail due to heterogeneous cell intrinsic resistance mechanisms. Ultimately, combination strategies that can address both on and off target resistance will be required to effect durable therapeutic responses.
Session Title: Tumor Heterogeneity and Evolution
Session Type: Educational Session
Session Start/End Time: Saturday, Apr 18, 2015, 1:00 PM – 3:00 PM
Location: Terrace Ballroom II-III (400 Level), Pennsylvania Convention Center
CME: CME-Designated
CME/CE Hours: 2
Session Description: One of the major challenges for both the measurement and management of cancer is its heterogeneity. Recent studies have revealed both extensive inter- and intra-tumor heterogeneity at the genotypic and phenotypic levels. Leaders in the field will discuss this challenge, its origins, dynamics and clinical importance. They will also review how we can best measure and deal with tumor heterogeneity, particularly intra-tumor heterogeneity.
Presentations:
Chairperson
Saturday, Apr 18, 2015, 1:00 PM – 3:00 PM
Carlo C. Maley. UCSF Helen Diller Family Comp. Cancer Center, San Francisco, CA
Universal biomarkers: How to handle tumor heterogeneity
Saturday, Apr 18, 2015, 1:00 PM – 1:25 PM
Carlo C. Maley. UCSF Helen Diller Family Comp. Cancer Center, San Francisco, CA
Discussion
Saturday, Apr 18, 2015, 1:25 PM – 1:30 PM
Heterogeneity of resistance to cancer therapy
Saturday, Apr 18, 2015, 1:30 PM – 1:55 PM
Ivana Bozic. HARVARD UNIV., Cambridge, MA
Discussion
Saturday, Apr 18, 2015, 1:55 PM – 2:00 PM
Determinants of phenotypic intra-tumor heterogeneity: integrative approach
Saturday, Apr 18, 2015, 2:00 PM – 2:25 PM
Andriy Marusyk, Michalina Janiszewska, Doris Tabassum. Dana-Farber Cancer Institute, Boston, MA, Dana-Farber Cancer Institute, Boston, MA
Discussion
Saturday, Apr 18, 2015, 2:25 PM – 2:30 PM
Cancer clonal complexity and evolution at the macro- and microheterogeneity scale
Saturday, Apr 18, 2015, 2:30 PM – 2:55 PM
Marco Gerlinger. Institute of Cancer Research, London, United Kingdom
Discussion
Saturday, Apr 18, 2015, 2:55 PM – 3:00 PM

From Ivana Bozic:

A spatial model predicts that dispersal and cell turnover limit intratumour heterogeneity.

Waclaw B, Bozic I, Pittman ME, Hruban RH, Vogelstein B, Nowak MA.

Nature. 2015 Sep 10;525(7568):261-4. doi: 10.1038/nature14971. Epub 2015 Aug 26.

PMID:

26308893

Similar articles

Select item 253494242.

Timing and heterogeneity of mutations associated with drug resistance in metastatic cancers.

Bozic I, Nowak MA.

Proc Natl Acad Sci U S A. 2014 Nov 11;111(45):15964-8. doi: 10.1073/pnas.1412075111. Epub 2014 Oct 27.

PMID:

25349424

Free PMC Article

Similar articles

Select item 238053823.

Evolutionary dynamics of cancer in response to targeted combination therapy.

Bozic I, Reiter JG, Allen B, Antal T, Chatterjee K, Shah P, Moon YS, Yaqubie A, Kelly N, Le DT, Lipson EJ, Chapman PB, Diaz LA Jr, Vogelstein B, Nowak MA.

Elife. 2013 Jun 25;2:e00747. doi: 10.7554/eLife.00747.

PMID:

23805382

Free PMC Article

Similar articles

Select item 227228434.

The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers.

Diaz LA Jr, Williams RT, Wu J, Kinde I, Hecht JR, Berlin J, Allen B, Bozic I, Reiter JG, Nowak MA, Kinzler KW, Oliner KS, Vogelstein B.

Nature. 2012 Jun 28;486(7404):537-40. doi: 10.1038/nature11219.

PMID:

22722843

Free PMC Article

Similar articles

 

Session Title: Mechanisms of Cancer Therapy Resistance
Session Type: Educational Session
Session Start/End Time: Saturday, Apr 18, 2015, 1:00 PM – 3:00 PM
Location: Room 204, Pennsylvania Convention Center
CME: CME-Designated
CME/CE Hours: 2
Session Description: Despite dramatic advances in the treatment of cancer, therapy resistance remains the most significant hurdle in improving the outcome of cancer patients. In this session, we will discuss many different aspects of therapy resistance, including a summary of our current understanding of therapy resistant tumor cell populations as well as analyses of the challenges associated with intratumoral heterogeneity and adaptive responses to targeted therapies.
Presentations:
Chairperson
Saturday, Apr 18, 2015, 1:00 PM – 3:00 PM
Charles Swanton. Cancer Research UK London Research Institute, London, United Kingdom
Tumor heterogeneity and drug resistance
Saturday, Apr 18, 2015, 1:00 PM – 1:30 PM
Charles Swanton. Cancer Research UK London Research Institute, London, United Kingdom
Discussion

Saturday, Apr 18, 2015, 1:30 PM – 1:40 PM
Discussion Discussion, Discussion

Principles of resistance to targeted therapy
Saturday, Apr 18, 2015, 1:40 PM – 2:10 PM
Levi A. Garraway. Dana-Farber Cancer Institute, Boston, MA
Discussion

Saturday, Apr 18, 2015, 2:10 PM – 2:20 PM
Discussion Discussion, Discussion

Adaptive re-wiring of signaling pathways driving drug resistance to targeted therapies
Saturday, Apr 18, 2015, 2:20 PM – 2:50 PM
Taru E. Muranen. Harvard Medical School, Boston, MA
Discussion

Saturday, Apr 18, 2015, 2:50 PM – 3:00 PM
Discussion Discussion, Discussion

Presentation Abstract  

 

 

 

Abstract Number: 737
Presentation Title: Clonal evolution of the HER2 L755S mutation as a mechanism of acquired HER-targeted therapy resistance
Presentation Time: Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Location: Section 30
Poster Board Number: 29
Author Block: Xiaowei Xu1, Agostina Nardone1, Huizhong Hu1, Lanfang Qin1, Sarmistha Nanda1, Laura Heiser2, Nicholas Wang2, Kyle Covington1, Edward Chen1, Alexander Renwick1, Tamika Mitchell1, Marty Shea1, Tao Wang1, Carmine De Angelis1, Alejandro Contreras1, Carolina Gutierrez1, Suzanne Fuqua1, Gary Chamness1, Chad Shaw1, Marilyn Li1, David Wheeler1, Susan Hilsenbeck1, Mothaffar Fahed Rimawi1, Joe Gray2, C.Kent Osborne1, Rachel Schiff1. 1Baylor College of Medicine, Houston, TX; 2Oregon Health & Science University, Portland, OR
Abstract Body: Background: Targeting HER2 with lapatinib (L), trastuzumab (T), or the LT combination, is effective in HER2+ breast cancer (BC), but acquired resistance commonly occurs. In our 12-week neoadjuvant
trial (TBCRC006) of LT without chemotherapy in HER2+ BC, the overall pathologic complete response (pCR) rate was 27%. To investigate resistance mechanisms, we developed 10 HER2+ BC cell line
models resistant (R) to one or both drugs (LR/TR/LTR). To discover potential predictive markers/therapeutic targets to circumvent resistance, we completed genomic profiling of the cell lines and a
subset of pre-treatment specimens from TBCRC006.
Methods: Parental (P) and LR/TR/LTR lines of 10 cell line models were profiled with whole exome/RNA sequencing. Mutations detected in R lines but not in P lines of the same model were identified. Mutation-specific Q-PCR was designed for sensitive quantification. Resistant cell and xenograft tumor growth were measured in response to drugs. Whole exome sequencing (>100X) and Ampliseq of 17 baseline tumor/normal pairs from TBCRC006 were performed.
Results: We found and validated the HER2 L755S mutation in the BT474/ATCC-LTR line and BT474/AZ-LR line (in ~30% of DNA/RNA), in which the HER pathway was reactivated for resistance. Overexpression of this mutation was previously shown to induce LR in HER2-negative BC cell lines, and resistant growth of BT474/AZ-LR line is significantly inhibited by HER2-L755S-specific siRNA knock-down, suggesting its role as an acquired L/LT resistance driver in HER2+ BC. Sequencing of BT474/AZ-LR single cell clones found the mutation in ~30% of HER2 copies in every cell. Using mutation-specific Q-PCR, we found statistically higher HER2 L755S levels in two BT474 parentals compared to P lines of SKBR3, AU565, and UACC812. These data suggest that HER2 L755S resistant subclones preexist in the BT474 parentals and were selected by L treatment to become the major clone in the two R lines. The HER1/2 irreversible tyrosine kinase inhibitor (TKI) afatinib (Afa) robustly inhibited growth of BT474/AZ-LR and BT474/ATCC-LTR cells (IC50: Afa 0.02µM vs. L 3 µM) and BT474/AZ-LR xenografts. Whole exome sequencing/Ampliseq of TBCRC006 found the HER2 L755S mutation in 1/17 primaries. This patient did not achieve pCR. The variant was present in 2% of DNA on both platforms, indicating a subclonal event of the resistance mutation.
Conclusion: Acquired L/LT resistance in the two BT474 R lines is due to selection of HER2 L755S subclones present in parental cells. The higher HER2 L755S
levels in BT474 parentals compared with other parentals, and detection of its subclonal presence in a pre-treatment HER2+ BC patient, suggest that sensitive mutation detection methods will be needed to identify patients with potentially actionable HER family mutations in primary tumor. Treating this patient group
with an irreversible TKI like Afa may prevent resistance and improve clinical outcome of this subset of HER2+ BC.
Presentation Number: SY07-04
Presentation Title: The evolutionary landscape of CLL: Therapeutic implications
Presentation Time: Sunday, Apr 19, 2015, 2:25 PM – 2:45 PM
Location: Grand Ballroom (300 Level), Pennsylvania Convention Center
Author Block: Catherine J. Wu. Dana-Farber Cancer Institute, Boston, MA
Abstract Body: Clonal evolution is a key feature of cancer progression and relapse. Recent studies across cancers have demonstrated the extensive degree of intratumoral heterogeneity present within individual cancers. We hypothesized that evolutionary dynamics contribute to the variations in disease tempo and response to therapy that are highly characteristic of chronic lymphocytic leukemia (CLL). We have recently investigated this phenomenon by developing a pipeline that estimates the fraction of cancer cells harboring each somatic mutation within a tumor through integration of whole-exome sequence (WES) and local copy number data (Landau et al., Cell 2013). By applying this analysis approach to 149 CLL cases, we discovered earlier and later cancer drivers, uncovered patterns of clonal evolution in CLL and linked the presence of subclones harboring driver mutations with adverse clinical outcome. Thus, our study, generated from a heterogeneous sample cohort, strongly supports the concept that CLL clonal evolution arises from mass extinction and therapeutic bottlenecks which lead to the emergence of highly fit (and treatment resistant) subclones. We further hypothesized that epigenetic heterogeneity also shapes CLL clonal evolution through interrelation with genetic heterogeneity. Indeed, in recent work, we have uncovered stochastic methylation disorder as the primary cause of methylation changes in CLL and cancer in general, and that this phenomena impacts gene transcription, genetic evolution and clinical outcome. Thus, integrated studies of genetic and epigenetic heterogeneity in CLL have revealed the complex and diverse evolutionary trajectories of these cancer cells.
Immunotherapy is exquisitely suited for specifically and simultaneously targeting multiple lesions. We have developed an approach that leverages whole-exome sequencing to systematically identify personal tumor mutations with immunogenic potential, which can be incorporated as antigen targets in multi-epitope personalized therapeutic vaccines. We are pioneering this approach in an ongoing trial in melanoma and will now expand this concept to address diverse malignancies. Our expectation is that the choice of tumor neoantigens for a vaccine bypasses thymic tolerance and thus generates highly specific and potent high-affinity T cell responses to eliminate tumors in any cancer, including both ‘trunk’ and ‘branch’ lesions.

 

Abstract Number: LB-056
Presentation Title: TP53 and RB1 alterations promote reprogramming and antiandrogen resistance in advanced prostate cancer
Presentation Time: Sunday, Apr 19, 2015, 4:50 PM – 5:05 PM
Location: Room 122, Pennsylvania Convention Center
Author Block: Ping Mu, Zhen Cao, Elizabeth Hoover, John Wongvipat, Chun-Hao Huang, Wouter Karthaus, Wassim Abida, Elisa De Stanchina, Charles Sawyers. Memorial Sloan Kettering Cancer Center, New York, NY
Abstract Body: Castration-resistant prostate cancer (CRPC) is one of the most difficult cancers to treat with conventional methods and is responsible for nearly all prostate cancer deaths in the US. The Sawyers laboratory first showed that the primary mechanism of resistance to antiandrogen therapy is elevated androgen receptor (AR) expression. Research based on this finding has led to the development of next-generation antiandrogen: enzalutamide. Despite the exciting clinical success of enzalutamide, about 60% of patients exhibit various degrees of resistance to this agent. Highly variable responses to enzalutamide limit the clinical benefit of this novel antiandrogen, underscoring the importance of understanding the mechanisms of enzalutamide resistance. Most recently, an unbiased SU2C-Prostate Cancer Dream Team metastatic CRPC sequencing project led by Dr. Sawyers and Dr. Chinnaiyan revealed that mutations in the TP53 locus are the most significantly enriched alteration in CRPC tumors when compared to primary prostate cancers. Moreover, deletions and decreased expressions of the TP53 and RB1 loci (co-occurrence and individual occurrence) are more commonly associated with CRPC than with primary tumors. These results established that alteration of the TP53 and RB1 pathways are associated with the development of antiandrogen resistance.
By knockdowning TP53 or/and RB1 in the castration resistant LNCaP/AR model, we demonstrate that the disruption of either TP53 or RB1 alone confers significant resistance to enzalutamide both in vitro and in vivo. Strikingly, the co-inactivation of these pathways confers the most dramatic resistance. Since up-regulation of either AR or AR target genes is not observed in the resistant tumors, loss of TP53 and RB1 function confers enzalutamide resistance likely through an AR independent mechanism. In the clinic, resistance to enzalutamide is increasingly being associated with a transition to a poorly differentiated or neuroendocrine-like histology. Interestingly, we observed significant up-regulations of the basal cell marker Ck5 and the neuroendocrine-like cell marker Synaptophysin in the TP53 and RB1 inactivated cells, as well as down-regulation of the luminal cell marker Ck8. The differences between these markers became even greater after enzalutamide treatment. By using the p53-stabilizing drug Nutlin, level of p53 is rescued and consequently the the decrease of AR protein caused by RB1 and TP53 knockdown is reversed. These results strongly suggest that interference of TP53 and RB1 pathways confers antiandrogen resistance by “priming” prostate cancer cells to reprogramming or transdifferentiation, likely neuroendocrine-like differentiation, in response to treatment. Futher experiments will be performed to assess the molecular mechanism of TP53/RB1 alterations in mediating cell programming and conferring antiandrogen resistance.

 

Abstract Number: LB-146
Presentation Title: TGF-β-induced tumor heterogeneity and drug resistance of cancer stem cells
Presentation Time: Monday, Apr 20, 2015, 1:00 PM – 5:00 PM
Location: Section 41
Author Block: Naoki Oshimori1, Daniel Oristian1, Elaine Fuchs2. 1Rockefeller University, New York, NY; 2HHMI/Rockefeller University, New York, NY
Abstract Body: Among the most common and life-threatening cancers world-wide, squamous cell carcinoma (SCC) exhibit high rates of tumor recurrence following anti-cancer therapy. Subsets of cancer stem cells (CSCs) often escape anti-cancer therapeutics and promote recurrence. However, its sources and mechanisms that generate tumor heterogeneity and therapy-resistant cell population are largely unknown. Tumor microenvironment may drive intratumor heterogeneity by transmitting signaling factors, oxygen and metabolites to tumor cells depending on their proximity to the local sources. While the hypothesis is attractive, experimental evidence is lacking, and non-genetic mechanisms that drive functional heterogeneity remain largely unknown. As a potential non-genetic factor, we focused on TGF-β because of its multiple roles in tumor progression.
Here we devise a functional reporter system to monitor, track and modify TGF-β signaling in mouse skin SCC in vivo. Using this approach, we found that perivascular TGF-β in the tumor microenvironment generates heterogeneity in TGF-β signaling in neighboring CSCs. This heterogeneity is functionally important: small subsets of TGF-β-responding CSCs proliferate more slowly than their non-responding counterparts. They also exhibit invasive morphology and a malignant differentiation program compared to their non-responding neighbors. By lineage tracing, we show that although TGF-β-responding CSCs clonally expand more slowly they gain a growth advantage in a remarkable ability to escape cisplatin-induced apoptosis. We show that indeed it is their progenies that make a substantial contribution in tumor recurrence. Surprisingly, the slower proliferating state of this subset of CSCs within the cancer correlated with but did not confer the survival advantage to anti-cancer drugs. Using transcriptomic, biochemical and genetic analyses, we unravel a novel mechanism by which heterogeneity in the tumor microenvironment allows a subset of CSCs to respond to TGF-β, and evade anti-cancer drugs.
Our findings also show that TGF-β established ability to suppress proliferation and promote invasion and metastasis do not happen sequentially, but rather simultaneously. This new work build upon the roles of this factor in tumor progression, and sets an important paradigm for a non-genetic factor that produces tumor heterogeneity.
Abstract Number: LB-129
Presentation Title: Identifying tumor subpopulations and the functional consequences of intratumor heterogeneity using single-cell profiling of breast cancer patient-derived xenografts
Presentation Time: Monday, Apr 20, 2015, 1:00 PM – 5:00 PM
Location: Section 41
Author Block: Paul Savage1, Sadiq M. Saleh1, Ernesto Iacucci1, Timothe Revil1, Yu-Chang Wang1, Nicholas Bertos1, Anie Monast1, Hong Zhao1, Margarita Souleimanova1, Keith Szulwach2, Chandana Batchu2, Atilla Omeroglu1, Morag Park1, Ioannis Ragoussis1. 1McGill University, Montreal, QC, Canada; 2Fluidigm Corporation, South San Francisco, CA
Abstract Body: Human breast tumors have been shown to exhibit extensive inter- and intra-tumor heterogeneity. While recent advances in genomic technologies have allowed us to deconvolute this heterogeneity, few studies have addressed the functional consequences of diversity within tumor populations. Here, we identified an index case for which we have derived a patient-derived xenograft (PDX) as a renewable tissue source to identify subpopulations and perform functional assays. On pathology, the tumor was an invasive ductal carcinoma which was hormone receptor-negative, HER2-positive (IHC 2+, FISH average HER2/CEP17 2.4), though the FISH signal was noted to be heterogeneous. On gene expression profiling of bulk samples, the primary tumor and PDX were classified as basal-like. We performed single cell RNA and exome sequencing of the PDX to identify population structure. Using a single sample predictor of breast cancer subtype, we have identified single basal-like, HER2-enriched and normal-like cells co-existing within the PDX tumor. Genes differentially expressed between these subpopulations are involved in proliferation and differentiation. Functional studies distinguishing these subpopulations are ongoing. Microfluidic whole genome amplification followed by whole exome capture of 81 single cells showed high and homogeneous target enrichment with >75% of reads mapping uniquely on target. Variant calling using GATK and Samtools revealed founder mutations in key genes as BRCA1 and TP53, as well as subclonal mutations that are being investigated further. Loss of heterozygocity was observed in 16 TCGA cancer driver genes and novel mutations in 7 cancer driver genes. These findings may be important in understanding the functional consequences of intra-tumor heterogeneity with respect to clinically important phenotypes such as invasion, metastasis and drug-resistance.
Abstract Number: 2847
Presentation Title: High complexity barcoding to study clonal dynamics in response to cancer therapy
Presentation Time: Monday, Apr 20, 2015, 4:35 PM – 4:50 PM
Location: Room 118, Pennsylvania Convention Center
Author Block: Hyo-eun C. Bhang1, David A. Ruddy1, Viveksagar Krishnamurthy Radhakrishna1, Rui Zhao2, Iris Kao1, Daniel Rakiec1, Pamela Shaw1, Marissa Balak1, Justina X. Caushi1, Elizabeth Ackley1, Nicholas Keen1, Michael R. Schlabach1, Michael Palmer1, William R. Sellers1, Franziska Michor2, Vesselina G. Cooke1, Joshua M. Korn1, Frank Stegmeier1. 1Novartis Institutes for BioMedical Research, Cambridge, MA; 2Dana-Farber Cancer Institute, Boston, MA
Abstract Body: Targeted therapies, such as erlotinib and imatinib, lead to dramatic clinical responses, but the emergence of resistance presents a significant challenge. Recent studies have revealed intratumoral heterogeneity as a potential source for the emergence of therapeutic resistance. However, it is still unclear if relapse/resistance is driven predominantly by pre-existing or de novo acquired alterations. To address this question, we developed a high-complexity barcode library, ClonTracer, which contains over 27 million unique DNA barcodes and thus enables the high resolution tracking of cancer cells under drug treatment. Using this library in two clinically relevant resistance models, we demonstrate that the majority of resistant clones pre-exist as rare subpopulations that become selected in response to therapeutic challenge. Furthermore, our data provide direct evidence that both genetic and non-genetic resistance mechanisms pre-exist in cancer cell populations. The ClonTracer barcoding strategy, together with mathematical modeling, enabled us to quantitatively dissect the frequency of drug-resistant subpopulations and evaluate the impact of combination treatments on the clonal complexity of these cancer models. Hence, monitoring of clonal diversity in drug-resistant cell populations by the ClonTracer barcoding strategy described here may provide a valuable tool to optimize therapeutic regimens towards the goal of curative cancer therapies.
Abstract Number: 3590
Presentation Title: Resistance mechanisms to ALK inhibitors
Presentation Time: Tuesday, Apr 21, 2015, 8:00 AM -12:00 PM
Location: Section 31
Poster Board Number: 13
Author Block: Ryohei Katayama1, Noriko Yanagitani1, Sumie Koike1, Takuya Sakashita1, Satoru Kitazono1, Makoto Nishio1, Yasushi Okuno2, Jeffrey A. Engelman3, Alice T. Shaw3, Naoya Fujita1. 1Japanese Foundation for Cancer Research, Tokyo, Japan; 2Graduate School of Medicine, Kyoto University, Kyoto, Japan; 3Massachusetts General Hospital Cancer Center, Boston, MA
Abstract Body: Purpose: ALK-rearranged non-small cell lung cancer (NSCLC) was first reported in 2007. Approximately 3-5% of NSCLCs harbor an ALK gene rearrangement. The first-generation ALK tyrosine kinase inhibitor (TKI) crizotinib is a standard therapy for patients with advanced ALK-rearranged NSCLC. Several next-generation ALK-TKIs have entered the clinic and have shown promising antitumor activity in crizotinib-resistant patients. As patients still relapse even on these next-generation ALK-TKIs, we examined mechanisms of resistance to one next-generation ALK-TKI – alectinib – and potential strategies to overcome this resistance.
Experimental Procedure: We established a cell line model of alectinib resistance, and analyzed resistant tumor specimens from patients who had relapsed on alectinib. Cell lines were also established under an IRB-approved protocol when there was sufficient fresh tumor tissue. We established Ba/F3 cells expressing EML4-ALK and performed ENU mutagenesis to compare potential crizotinib or alectinib-resistance mutations. In addition, we developed Ba/F3 models harboring ALK resistance mutations and evaluated the potency of multiple next-generation ALK-TKIs including 3rd generation ALK inhibitor in these models and in vivo. To elucidate structure-activity-relationships of ALK resistance mutations, we performed computational thermodynamic simulation with MP-CAFEE.
Results: We identified multiple resistance mutations, including ALK I1171N, I1171S, and V1180L, from the ENU mutagenesis screen and the cell line model. In addition we found secondary mutations at the I1171 residue from the Japanese patients who developed resistance to alectinib or crizotinib. Both ALK mutations (V1180L and I1171 mutations) conferred resistance to alectinib as well as to crizotinib, but were sensitive to ceritinib and other next-generation ALK-TKIs. Based on thermodynamics simulation, each resistance mutation is predicted to lead to distinct structural alterations that decrease the binding affinity of ALK-TKIs for ALK.
Conclusions: We have identified multiple alectinib-resistance mutations from the cell line model, patient derived cell lines, and tumor tissues, and ENU mutagenesis. ALK secondary mutations arising after alectinib exposure are sensitive to other next generation ALK-TKIs. These findings suggest a potential role for sequential therapy with multiple next-generation ALK-TKIs in patients with advanced, ALK-rearranged cancers.
Session Title: Mechanisms of Resistance: From Signaling Pathways to Stem Cells
Session Type: Major Symposium
Session Start/End Time: Tuesday, Apr 21, 2015, 10:30 AM -12:30 PM
Location: Terrace Ballroom II-III (400 Level), Pennsylvania Convention Center
CME: CME-Designated
CME/CE Hours: 2
Session Description: Even the most effective cancer therapies are limited due to the development of one or more resistance mechanisms. Acquired resistance to targeted therapies can, in some cases, be attributed to the selective propagation of a small population of intrinsically resistant cells. However, there is also evidence that cancer drugs themselves can drive resistance by triggering the biochemical- or genetic-reprogramming of cells within the tumor or its microenvironment. Therefore, understanding drug resistance at the molecular and biological levels may enable the selection of specific drug combinations to counteract these adaptive responses. This symposium will explore some of the recent advances addressing the molecular basis of cancer cell drug resistance. We will address how tumor cell signaling pathways become rewired to facilitate tumor cell survival in the face of some of our most promising cancer drugs. Another topic to be discussed involves how drugs select for or induce the reprogramming of tumor cells toward a stem-like, drug resistant fate. By targeting the molecular driver(s) of rewired signaling pathways and/or cancer stemness it may be possible to select drug combinations that prevent the reprogramming of tumors and thereby delay or eliminate the onset of drug resistance.
Presentations:
Chairperson
Tuesday, Apr 21, 2015, 10:30 AM -12:30 PM
David A. Cheresh. UCSD Moores Cancer Center, La Jolla, CA
Introduction
Tuesday, Apr 21, 2015, 10:30 AM -10:40 AM
Resistance to tyrosine kinase inhibitors: Heterogeneity and therapeutic strategies.
Tuesday, Apr 21, 2015, 10:40 AM -10:55 AM
Jeffrey A. Engelman. Massachusetts General Hospital, Boston, MA
Discussion
Tuesday, Apr 21, 2015, 10:55 AM -11:00 AM
NG04: Clinical acquired resistance to RAF inhibitor combinations in BRAF mutant colorectal cancer through MAPK pathway alterations
Tuesday, Apr 21, 2015, 11:00 AM -11:15 AM
Ryan B. Corcoran, Leanne G. Ahronian, Eliezer Van Allen, Erin M. Coffee, Nikhil Wagle, Eunice L. Kwak, Jason E. Faris, A. John Iafrate, Levi A. Garraway, Jeffrey A. Engelman. Massachusetts General Hospital Cancer Center, Boston, MA, Dana-Farber Cancer Institute, Boston, MA
Discussion
Tuesday, Apr 21, 2015, 11:15 AM -11:20 AM
SY27-02: Tumour heterogeneity and therapy resistance in melanoma
Tuesday, Apr 21, 2015, 11:20 AM -11:35 AM
Claudia Wellbrock. Univ. of Manchester, Manchester, United Kingdom

Presentation Number: SY27-02
Presentation Title: Tumour heterogeneity and therapy resistance in melanoma
Presentation Time: Tuesday, Apr 21, 2015, 11:20 AM -11:35 AM
Location: Terrace Ballroom II-III (400 Level), Pennsylvania Convention Center
Author Block: Claudia Wellbrock. Univ. of Manchester, Manchester, United Kingdom
Abstract Body: Solid tumors are structurally very complex; they consist of heterogeneous cancer cell populations, other non-cancerous cell types and a distinct extracellular matrix. Interactions of cancer cells with non-cancerous cells is well investigated, and our recent work in melanoma has demonstrated that the cellular environment that surrounds cancer cells has a major impact on the way a patient responds to MAP-kinase pathway targeting therapy.
We have shown that intra-tumor signaling within a heterogeneous tumor can have a major impact on the efficacy of BRAF and MEK inhibitors. With the increasing evidence of genetic and phenotypic heterogeneity within tumors, intra-tumor signaling between individual cancer-cell subpopulations is therefore a crucial factor that needs to be considered in future therapy approaches. Our work has identified the ‘melanocyte-lineage survival oncogene’ MITF as an important player in phenotypic heterogeneity (MITFhigh and MITFlow cells) in melanoma, and MITF expression levels are crucial for the response to MAP-kinase pathway targeted therapy. We found that ‘MITF heterogeneity’ can be caused by cell-autonomous mechanisms or by the microenvironment, including the immune-microenvironment.
We have identified various mechanisms underlying MITF action in resistance to BRAF and MEK inhibitors in melanoma. In MITFhigh expressing cells, MITF confers cell-autonomous resistance to MAP-kinase pathway targeted therapy. Moreover, it appears that in melanomas heterogeneous for MITF expression (MITFhigh and MITFlow cells), individual subpopulations of resistant and sensitive cells communicate and MITF can contribute to overall tumor-resistance through intra-tumor signaling. Finally, we have identified a novel approach of interfering with MITF action, which profoundly sensitizes melanoma to MAP-kinase pathway targeted therapy.
Discussion
Tuesday, Apr 21, 2015, 11:35 AM -11:40 AM
SY27-03: Breast cancer stem cell state transitions mediate therapeutic resistance
Tuesday, Apr 21, 2015, 11:40 AM -11:55 AM
Max S. Wicha. University of Michigan, Comprehensive Cancer Center, Ann Arbor, MI
Discussion
Tuesday, Apr 21, 2015, 11:55 AM -12:00 PM
SY27-04: Induction of cancer stemness and drug resistance by EGFR blockade
Tuesday, Apr 21, 2015, 12:00 PM -12:15 PM
David A. Cheresh. UCSD Moores Cancer Center, La Jolla, CA

 

Cellular Reprogramming in Carcinogenesis: Implications for Tumor Heterogeneity, Prognosis, and Therapy
Session Type: Major Symposium
Session Start/End Time: Tuesday, Apr 21, 2015, 10:30 AM -12:30 PM
Location: Room 103, Pennsylvania Convention Center
CME: CME-Designated
CME/CE Hours: 2
Session Description: Cancers, both solid and liquid, consist of phenotypically heterogeneous cell types that make up the full cellular complement of disease. Deep sequencing of bulk cancers also frequently reveals a genetic intratumoral heterogeneity that reflects clonal evolution in space and in time and under the influence of treatment. How the distinct phenotypic and genotypic cells contribute to individual cancer growth and progression is incompletely understood. In this symposium, we will discuss issues of cancer heterogeneity and effects on growth and treatment resistance, with emphasis on cancer cell functional properties and influences of the microenvironment, interclonal genomic heterogeneity, and lineage relationships between cancer cells with stem cell and differentiated properties. Understanding these complex cellular relationships within cancers will have critical implications for devising more effective treatments.
Presentations:
Chairperson
Tuesday, Apr 21, 2015, 10:30 AM -12:30 PM
Peter B. Dirks. Univ. of Toronto Hospital for Sick Children, Toronto, ON, Canada
Introduction

Tuesday, Apr 21, 2015, 10:30 AM -10:40 AM

Origins, evolution and selection in childhood leukaemia
Tuesday, Apr 21, 2015, 10:40 AM -11:00 AM
Tariq Enver. Cancer Research UK, London, United Kingdom
Discussion

Tuesday, Apr 21, 2015, 11:00 AM -11:05 AM

Cytokine-controlled stem cell plasticity inintestinal tumorigenesis
Tuesday, Apr 21, 2015, 11:05 AM -11:25 AM
Florian Greten. Georg-Speyer-Haus, Frankfurt, Germany
Discussion

Tuesday, Apr 21, 2015, 11:25 AM -11:30 AM

SY23-03: Intratumoural heterogeneity in human serous ovarian carcinoma
Tuesday, Apr 21, 2015, 11:30 AM -11:50 AM
John P. Stingl. Cancer Research UK Cambridge Research Inst., Cambridge, United Kingdom
Discussion

Tuesday, Apr 21, 2015, 11:50 AM -11:55 AM

Functional and genomic heterogeneity in brain tumors
Tuesday, Apr 21, 2015, 11:55 AM -12:15 PM

 

Proc Natl Acad Sci U S A. 2015 Jan 20;112(3):851-6. doi: 10.1073/pnas.1320611111. Epub 2015 Jan 5.

Single cell-derived clonal analysis of human glioblastoma links functional and genomic heterogeneity.

Meyer M1, Reimand J2, Lan X3, Head R1, Zhu X1, Kushida M1, Bayani J4, Pressey JC5, Lionel AC6, Clarke ID7, Cusimano M8, Squire JA9, Scherer SW6, Bernstein M10, Woodin MA5, Bader GD11, Dirks PB12.

Author information

Abstract

Glioblastoma (GBM) is a cancer comprised of morphologically, genetically, and phenotypically diverse cells. However, an understanding of the functional significance of intratumoral heterogeneity is lacking. We devised a method to isolate and functionally profile tumorigenic clones from patient glioblastoma samples. Individual clones demonstrated unique proliferation and differentiation abilities. Importantly, naïve patient tumors included clones that were temozolomide resistant, indicating that resistance to conventional GBM therapy can preexist in untreated tumors at a clonal level. Further, candidate therapies for resistant clones were detected with clone-specific drug screening. Genomic analyses revealed genes and pathways that associate with specific functional behavior of single clones. Our results suggest that functional clonal profiling used to identify tumorigenic and drug-resistant tumor clones will lead to the discovery of new GBM clone-specific treatment strategies.

—————————————————————————————————

 

739: Tumor cell plasticity with transition to a mesenchymal phenotype is a mechanism of chemoresistance that is reversed by Notch pathway inhibition in lung adenocarcinoma
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Khaled A. Hassan. University Of Michigan, Ann Arbor, MI

745: Oncostatin M receptor activation leads to molecular targeted therapy resistance in non-small cell lung cancer
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Kazuhiko Shien1, Vassiliki A. Papadimitrakopoulou1, Dennis Ruder1, Nana E. Hanson1, Neda Kalhor1, J. Jack Lee1, Waun Ki Hong1, Ximing Tang1, Roy S. Herbst2, Luc Girard3, John D. Minna3, Jonathan M. Kurie1, Ignacio I. Wistuba1, Julie G. Izzo1. 1University of Texas MD Anderson Cancer Center, Houston, TX; 2Yale Cancer Center, Yale School of Medicine, New Haven, CT; 3Hamon Center for Therapeutic Oncology Research, University of Texas Southwestern Medical Center, Dallas, TX

746: Activation of EGFR bypass signaling through TGFα overexpression induces acquired resistance to alectinib in ALK-translocated lung cancer cells
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Tetsuo Tani, Hiroyuki Yasuda, Junko Hamamoto, Aoi Kuroda, Daisuke Arai, Kota Ishioka, Keiko Ohgino, Ichiro Kawada, Katsuhiko Naoki, Hayashi Yuichiro, Tomoko Betsuyaku, Kenzo Soejima. Keio University, Tokyo, Japan

752: Elucidating the mechanisms of acquired resistance in lung adenocarcinomas
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Sandra Ortiz-Cuarán1, Lynnette Fernandez-Cuesta1, Christine M. Lovly2, Marc Bos1, Matthias Scheffler3, Sebastian Michels3, Kerstin Albus4, Lydia Meyer4, Katharina König4, Ilona Dahmen1, Christian Mueller1, Luca Ozretić4, Lars Tharun4, Philipp Schaub1, Alexandra Florin4, Berit Pinther1, Nike Bahlmann1, Sascha Ansén3, Martin Peifer1, Lukas C. Heukamp4, Reinhard Buettner4, Martin L. Sos1, Jürgen Wolf3, William Pao2, Roman K. Thomas1. 1University of Cologne, Cologne, Germany; 2Department of Medicine, Vanderbilt University, Nashville, TN; 3Department of Internal Medicine, Center for Integrated Oncology Köln-Bonn, University Hospital Cologne, Cologne, Germany; 4Institute of Pathology, University Hospital Cologne, Cologne, Germany

760: On the evolution of erlotinib-resistant NSCLC subpopulations
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Michael E. Ramirez1, Robert J. Steininger, III1, Lani F. Wu2, Steven J. Altschuler2. 1UT Southwestern, Dallas, TX; 2UCSF, San Francisco, CA
763: Implications of resistance patterns with NSCLC targeted agents
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
David J. Stewart, Paul Wheatley-Price, Rob MacRae, Jason Pantarotto. University of Ottawa, Ottawa, ON, Canada

 

768: A kinome-wide siRNA screen identifies modifiers of sensitivity to the EGFR T790M-targeted tyrosine kinase inhibitor (TKI), AZD9291, in EGFR mutant lung adenocarcinoma
Sunday, Apr 19, 2015, 1:00 PM – 5:00 PM
Eiki Ichihara1, Joshua A. Bauer2, Pengcheng Lu3, Fei Ye3, Darren Cross4, William Pao1, Christine M. Lovly1. 1Vanderbilt University School of Medicine, Nashville, TN; 2Vanderbilt Institute of Chemical Biology High-Throughput Screening Facility, Nashville, TN; 3Vanderbilt University Medical Center, Nashville, TN; 4AstraZeneca Oncology Innovative Medicines, United Kingdom

LB-055: Clinical acquired resistance to RAF inhibitor combinations in BRAF-mutant colorectal cancer through MAPK pathway alterations
Sunday, Apr 19, 2015, 4:35 PM – 4:50 PM
Leanne G. Ahronian1, Erin M. Sennott1, Eliezer M. Van Allen2, Nikhil Wagle2, Eunice L. Kwak1, Jason E. Faris1, Jason T. Godfrey1, Koki Nishimura1, Kerry D. Lynch3, Craig H. Mermel1, Elizabeth L. Lockerman1, Anuj Kalsy1, Joseph M. Gurski, Jr.1, Samira Bahl4, Kristin Anderka4, Lisa M. Green4, Niall J. Lennon4, Tiffany G. Huynh3, Mari Mino-Kenudson3, Gad Getz1, Dora Dias-Santagata3, A. John Iafrate3, Jeffrey A. Engelman1, Levi A. Garraway2, Ryan B. Corcoran1. 1Massachusetts General Hospital Cancer Center, Boston, MA; 2Dana Farber Cancer Institute, Boston, MA; 3Massachusetts General Hospital Department of Pathology, Boston, MA; 4Broad Institute of Massachusetts Institute of Technology and Harvard, Cambridge, MA

 

Other Articles on this Site Related to Tumor Heterogeneity Include

Notes On Tumor Heterogeneity: Targets and Mechanisms, from the 2015 AACR Meeting in Philadelphia PA

Issues in Personalized Medicine: Discussions of Intratumor Heterogeneity from the Oncology Pharma forum on LinkedIn

Issues in Personalized Medicine in Cancer: Intratumor Heterogeneity and Branched Evolution Revealed by Multiregion Sequencing

CANCER COMPLEXITY: Heterogeneity in Tumor Progression and Drug Response – 2015 Annual Symposium @Koch Institute for Integrative Cancer Research at MIT – W34, 6/12/2015 9:00 AM EDT – 4:30 PM EDT

In vitro Models of Tumor Microenvironment for New Cancer Target and Drug Discovery, 11/17 – 11/19/2014, Hyatt Boston Harbor

What can we expect of tumor therapeutic response?

 

Read Full Post »

Humanized Mice May Revolutionize Cancer Drug Discovery

Curator: Stephen J. Williams, Ph.D.

Humanized Mice May Revolutionize Cancer Drug Discovery

Word Cloud by Zach Day

Decades ago cancer research and the process of oncology drug discovery was revolutionized by the development of mice deficient in their immune system, allowing for the successful implantation of human-derived tumors. The ability to implant human tumors without rejection allowed researchers to study how the kinetics of human tumor growth in its three-dimensional environment, evaluate potential human oncogenes and drivers of oncogenesis, and evaluate potential chemotherapeutic therapies. Indeed, the standard preclinical test for antitumor activity has involved the subcutaneous xenograft model in immunocompromised (SCID or nude athymic) mice. More detail is given in the follow posts in which I describe some early pioneers in this work as well as the development of large animal SCID models:

Heroes in Medical Research: Developing Models for Cancer Research

The SCID Pig: How Pigs are becoming a Great Alternate Model for Cancer Research

The SCID Pig II: Researchers Develop Another SCID Pig, And Another Great Model For Cancer Research

This strategy (putting human tumor cells into immunocompromised mice and testing therapeutic genes and /or compounds) has worked extremely well for most cytotoxic chemotherapeutics (those chemotherapeutic drugs with mechanisms of action related to cell kill, vital cell functions, and cell cycle). For example the NCI 60 panel of human tumor cell lines has proved predictive for the chemosensitivity of a wide range of compounds.

Even though the immunocompromised model has contributed greatly to the chemotherapeutic drug discovery process. using these models to develop the new line of immuno-oncology products has been met with challenges three which I highlight below with curated database of references and examples.

From a practical standpoint development of a mouse which can act as a recipient for human tumors yet have a humanized immune system allows for the preclinical evaluation of antitumoral effect of therapeutic antibodies without the need to use neutralizing antibodies to the comparable mouse epitope,   thereby reducing the complexity of the study and preventing complications related to pharmacokinetics.

Champions Oncology Files Patents for Use of PDX Platform in Immune-Oncology

Hackensack, NJ – August 17, 2015 – Champions Oncology, Inc. (OTC: CSBR), engaged in the development of advanced technology solutions and services to personalize the development and use of oncology drugs, today announced that it has filed two patent applications with the United States Patent and Trademark Office (USPTO) relating to the development and use of mice with humanized immune systems to test immune-oncology drugs and therapeutic cancer vaccines.

Dr. David Sidransky, the founder and Chairman of Champions Oncology commented, “Drug development ‎in the immune-oncology space is fundamentally changing our approach to cancer treatment. These patents represent potentially invaluable tools for developing and personalizing immune therapy based on cutting edge sequence analysis, bioinformatics and our unique in vivo models.”

Joel Ackerman, Chief Executive Officer of Champions Oncology stated, “Developing intellectual property related to our Champions TumorGraft® platform has been an important component of strategy. The filing of these patents is an important milestone in leveraging our research and development investment to expand our platform and create proprietary tools for use by our pharmaceutical partners. We continue to look for additional revenue streams to supplement our fee-for-service business and we believe these patents will help us capture more of the value we create for our customers in the future.”

The first patent filing covers the methodology used by the Company to create a mouse model, containing a humanized immune system and a human tumor xenograft, which is capable of testing the efficacy of immune-oncology agents, both as single agents and in combination with anti-neoplastic drugs. The second patent filing relates to the detection of neoantigens and their role in the development of anti-cancer vaccines.

Keren Pez, Chief Scientific Officer, explained, “In the last few years, there has been a significant increase in cancer research that focuses on exploring the power of the human immune system to attack tumors. However, it’s challenging to test immune-oncology agents in traditional animal models due to the major differences between human and murine immune systems. The Champions ImmunoGraft™ platform has the unique ability of mimicking a human adaptive immune response in the mice, which allows us to specifically evaluate a variety of cancer therapeutics that modulate human immunity.

“Therapeutic vaccines that trigger the immune system to mount a response against a growing tumor are another area of intense interest. The development of an effective vaccine remains challenging but has an outstanding curative potential. Tumors harbor mutations in DNA that result in the translation of aberrant proteins. While these proteins have the potential to provoke an immune response that destructs early-stage cancer development, often the immune response becomes insufficient. Vaccines can trigger it by proactively challenging the system with these specific mutated peptides. Nevertheless, developing anti-cancer vaccines that effectively inhibit tumor growth has been complicated, partially due to challenges in finding the critical mutations, among others difficulties. With the more recent advances in genome sequencing, it’s now possible to identify tumor-specific antigens, or neoantigens, that naturally develop as an individual’s tumor grows and mutates,” she continued.

Traumatic spinal cord injury in mice with human immune systems.

Carpenter RS, Kigerl KA, Marbourg JM, Gaudet AD, Huey D, Niewiesk S, Popovich PG.

Exp Neurol. 2015 Jul 17;271:432-444. doi: 10.1016/j.expneurol.2015.07.011. [Epub ahead of print]

Inflamm Bowel Dis. 2015 Jul;21(7):1652-73. doi: 10.1097/MIB.0000000000000446.

Use of Humanized Mice to Study the Pathogenesis of Autoimmune and Inflammatory Diseases.

Koboziev I1, Jones-Hall Y, Valentine JF, Webb CR, Furr KL, Grisham MB.

Author information

Abstract

Animal models of disease have been used extensively by the research community for the past several decades to better understand the pathogenesis of different diseases and assess the efficacy and toxicity of different therapeutic agents. Retrospective analyses of numerous preclinical intervention studies using mouse models of acute and chronic inflammatory diseases reveal a generalized failure to translate promising interventions or therapeutics into clinically effective treatments in patients. Although several possible reasons have been suggested to account for this generalized failure to translate therapeutic efficacy from the laboratory bench to the patient’s bedside, it is becoming increasingly apparent that the mouse immune system is substantially different from the human. Indeed, it is well known that >80 major differences exist between mouse and human immunology; all of which contribute to significant differences in immune system development, activation, and responses to challenges in innate and adaptive immunity. This inconvenient reality has prompted investigators to attempt to humanize the mouse immune system to address important human-specific questions that are impossible to study in patients. The successful long-term engraftment of human hematolymphoid cells in mice would provide investigators with a relatively inexpensive small animal model to study clinically relevant mechanisms and facilitate the evaluation of human-specific therapies in vivo. The discovery that targeted mutation of the IL-2 receptor common gamma chain in lymphopenic mice allows for the long-term engraftment of functional human immune cells has advanced greatly our ability to humanize the mouse immune system. The objective of this review is to present a brief overview of the recent advances that have been made in the development and use of humanized mice with special emphasis on autoimmune and chronic inflammatory diseases. In addition, we discuss the use of these unique mouse models to define the human-specific immunopathological mechanisms responsible for the induction and perpetuation of chronic gut inflammation.

J Immunother Cancer. 2015 Apr 21;3:12. doi: 10.1186/s40425-015-0056-2. eCollection 2015.

Human tumor infiltrating lymphocytes cooperatively regulate prostate tumor growth in a humanized mouse model.

Roth MD1, Harui A1.

Author information

Abstract

BACKGROUND:

The complex interactions that occur between human tumors, tumor infiltrating lymphocytes (TIL) and the systemic immune system are likely to define critical factors in the host response to cancer. While conventional animal models have identified an array of potential anti-tumor therapies, mouse models often fail to translate into effective human treatments. Our goal is to establish a humanized tumor model as a more effective pre-clinical platform for understanding and manipulating TIL.

METHODS:

The immune system in NOD/SCID/IL-2Rγnull (NSG) mice was reconstituted by the co-administration of human peripheral blood lymphocytes (PBL) or subsets (CD4+ or CD8+) and autologous human dendritic cells (DC), and animals simultaneously challenged by implanting human prostate cancer cells (PC3 line). Tumor growth was evaluated over time and the phenotype of recovered splenocytes and TIL characterized by flow cytometry and immunohistochemistry (IHC). Serum levels of circulating cytokines and chemokines were also assessed.

RESULTS:

A tumor-bearing huPBL-NSG model was established in which human leukocytes reconstituted secondary lymphoid organs and promoted the accumulation of TIL. These TIL exhibited a unique phenotype when compared to splenocytes with a predominance of CD8+ T cells that exhibited increased expression of CD69, CD56, and an effector memory phenotype. TIL from huPBL-NSG animals closely matched the features of TIL recovered from primary human prostate cancers. Human cytokines were readily detectible in the serum and exhibited a different profile in animals implanted with PBL alone, tumor alone, and those reconstituted with both. Immune reconstitution slowed but could not eliminate tumor growth and this effect required the presence of CD4+ T cell help.

CONCLUSIONS:

Simultaneous implantation of human PBL, DC and tumor results in a huPBL-NSG model that recapitulates the development of human TIL and allows an assessment of tumor and immune system interaction that cannot be carried out in humans. Furthermore, the capacity to manipulate individual features and cell populations provides an opportunity for hypothesis testing and outcome monitoring in a humanized system that may be more relevant than conventional mouse models.

Methods Mol Biol. 2014;1213:379-88. doi: 10.1007/978-1-4939-1453-1_31.

A chimeric mouse model to study immunopathogenesis of HCV infection.

Bility MT1, Curtis A, Su L.

Author information

Abstract

Several human hepatotropic pathogens including chronic hepatitis C virus (HCV) have narrow species restriction, thus hindering research and therapeutics development against these pathogens. Developing a rodent model that accurately recapitulates hepatotropic pathogens infection, human immune response, chronic hepatitis, and associated immunopathogenesis is essential for research and therapeutics development. Here, we describe the recently developed AFC8 humanized liver- and immune system-mouse model for studying chronic hepatitis C virus and associated human immune response, chronic hepatitis, and liver fibrosis.

PMID:

25173399

[PubMed – indexed for MEDLINE]

PMCID:

PMC4329723

Free PMC Article

Immune humanization of immunodeficient mice using diagnostic bone marrow aspirates from carcinoma patients.

Werner-Klein M, Proske J, Werno C, Schneider K, Hofmann HS, Rack B, Buchholz S, Ganzer R, Blana A, Seelbach-Göbel B, Nitsche U, Männel DN, Klein CA.

PLoS One. 2014 May 15;9(5):e97860. doi: 10.1371/journal.pone.0097860. eCollection 2014.

From 2015 AACR National Meeting in Philadelphia

LB-050: Patient-derived tumor xenografts in humanized NSG mice: a model to study immune responses in cancer therapy
Sunday, Apr 19, 2015, 3:20 PM – 3:35 PM
Minan Wang1, James G. Keck1, Mingshan Cheng1, Danying Cai1, Leonard Shultz2, Karolina Palucka2, Jacques Banchereau2, Carol Bult2, Rick Huntress2. 1The Jackson Laboratory, Sacramento, CA; 2The Jackson Laboratory, Bar Harbor, ME

References

  1. Paull KD, Shoemaker RH, Hodes L, Monks A, Scudiero DA, Rubinstein L, Plowman J, Boyd MR. J Natl Cancer Inst. 1989;81:1088–1092. [PubMed]
  2. Shi LM, Fan Y, Lee JK, Waltham M, Andrews DT, Scherf U, Paull KD, Weinstein JN. J Chem Inf Comput Sci. 2000;40:367–379. [PubMed]
  3. Monks A, Scudiero D, Skehan P, Shoemaker R, Paull K, Vistica D, Hose C, Langley J, Cronise P, Vaigro-Wolff A, et al. J Natl Cancer Inst. 1991;83:757–766. [PubMed]
  4. Potti A, Dressman HK, Bild A, et al. Genomic signatures to guide the use of chemotherapeutics. Nat Med. 2006;12:1294–1300. [PubMed]
  5. Baggerly KA, Coombes KR. Deriving chemosensitivity from cell lines: forensic bioinformatics and reproducible research in high-throughput biology. Ann Appl Stat. 2009;3:1309–1334.
  6. Carlson, B. Putting Oncology Patients at Risk Biotechnol Healthc. 2012 Fall; 9(3): 17–21.
  7. Salter KH, Acharya CR, Walters KS, et al. An Integrated Approach to the Prediction of Chemotherapeutic Response in Patients with Breast Cancer. Ouchi T, ed. PLoS ONE. 2008;3(4):e1908. NOTE RETRACTED PAPER

Other posts on this site on Animal Models, Disease and Cancer Include:

Heroes in Medical Research: Developing Models for Cancer Research

Guidelines for the welfare and use of animals in cancer research

Model mimicking clinical profile of patients with ovarian cancer @ Yale School of Medicine

Vaccines, Small Peptides, aptamers and Immunotherapy [9]

Immunotherapy in Cancer: A Series of Twelve Articles in the Frontier of Oncology by Larry H Bernstein, MD, FCAP

Mouse With ‘Humanized Version’ Of Human Language Gene Provides Clues To Language Development

The SCID Pig: How Pigs are becoming a Great Alternate Model for Cancer Research

The SCID Pig II: Researchers Develop Another SCID Pig, And Another Great Model For Cancer Research

Read Full Post »

New NIH breast cancer research to focus on prevention

Reporter: Stephen J. Williams, PhD

Print this page

A new phase of the Breast Cancer and the Environment Research Program (BCERP), focused on prevention, is being launched at the National Institutes of Health. Grant-funded researchers will now work across scientific disciplines, involve new racially and ethnically diverse communities, and expand the study of risk factors that precede breast cancer, such as breast density.

These new directions reflect recommendations made by the Interagency Breast Cancer and Environmental Research Coordinating Committee (IBCERCC) in 2013. IBCERCC was congressionally mandated to review the state of the science around breast cancer and environmental influences by the Breast Cancer and Environmental Research Act. Recommendations included prioritizing prevention, involving transdisciplinary research teams, engaging public stakeholders, collaborating across federal agencies, and communicating the science to the public.

This broadened research focus will add to the growing knowledge of environmental and genetic factors that may influence breast cancer risk across the lifespan. The six new BCERP projects, plus a new coordinating center promoting cross-project collaboration, are jointly funded by the National Institute of Environmental Health Sciences (NIEHS) and the National Cancer Institute. All projects involve strong partnerships between researchers and organizations focused on breast cancer prevention or environmental health.

The new research will be conducted at the following institutions

  • Brigham and Women’s Hospital, Boston
  • City of Hope/Beckman Research Institute, Duarte, California
  • Columbia University, New York City
  • Georgetown Lombardi Comprehensive Cancer Center, Washington, D.C.
  • Michigan State University, Lansing
  • University of Massachusetts, Amherst
  • University of Wisconsin – Madison (Coordinating Center)

“The beauty of this research is that scientific discoveries and community observations inform each other, in order to dive deeper into the complex causes of breast cancer,” said Gwen Collman, Ph.D., director of NIEHS Division of Extramural Research and Training.

The focus on minority and socio-economically disadvantaged women is an important step in addressing disparities in breast cancer outcomes. Although African-American women are diagnosed with breast cancer less often than white women, more aggressive cancers and breast cancer deaths are more common among African-American women.

Another new direction for BCERP is research on the role of breast density as a possible intermediate risk factor for breast cancer. Dense breast tissue is one of the most common risk factors for breast cancer. Identifying links between environmental exposures and high breast density may provide new insights into prevention.

“These priorities reflect our continued commitment to breast cancer prevention,” noted Caroline Dilworth, Ph.D., BCERP program lead at NIEHS. “Our goal is to build on the high quality science we’ve been funding for more than a decade, while also being responsive to the expert recommendations of the IBCERCC report.”

Grant Numbers: U01ES026130, U01ES026137, U01ES026122, U01ES026132, U01ES026119, U01ES026140, U01ES026127

NIEHS supports research to understand the effects of the environment on human health and is part of NIH. For more information on environmental health topics, visit www.niehs.nih.gov. Subscribe to one or more of the NIEHS news lists to stay current on NIEHS news, press releases, grant opportunities, training, events, and publications.

The National Cancer Institute leads the National Cancer Program and the NIH’s efforts to dramatically reduce the prevalence of cancer and improve the lives of cancer patients and their families, through research into prevention and cancer biology, the development of new interventions, and the training and mentoring of new researchers. For more information about cancer, please visit the NCI website at http://www.cancer.gov or call NCI’s Cancer Information Service at 1-800-4-CANCER.

About the National Institutes of Health (NIH): NIH, the nation’s medical research agency, includes 27 Institutes and Centers and is a component of the U.S. Department of Health and Human Services. NIH is the primary federal agency conducting and supporting basic, clinical, and translational medical research, and is investigating the causes, treatments, and cures for both common and rare diseases. For more information about NIH and its programs, visit www.nih.gov.

Other posts on this site on  Cancer and Early Detection  include

Early Detection of Prostate Cancer: American Urological Association (AUA) Guideline

Mechanism involved in Breast Cancer Cell Growth: Function in Early Detection & Treatment

Warning signs may lead to better early detection of ovarian cancer

‘Mosaicism’ is Associated with Aging and Chronic Diseases like Cancer: detection of genetic mosaicism could be an early marker for detecting cancer.

CDC Findings: Due to Aging Population, Actual Number of Cancer Deaths is Rising while Risk of Dying From Cancer is Falling in the US

Read Full Post »

Curation of Recently Halted Oncology Trials Due to Serious Adverse Events – 2015

Curator: Stephen J. Williams, Ph.D.

The following is reports of oncology clinical trials in 2015 which have been halted for Serious Adverse Events (SAE), in most instances of an idiopathic nature. For comparison I have listed (as of this writing) the oncology drug approvals (8) for 2015. (from CenterWatch.com)

Oncology Drugs Approved in 2015

Farydak (panobinostat); Novartis; For the treatment of multiple myeloma, Approved February 2015

Ibrance (palbociclib); Pfizer; For the treatment of ER-positive, HER2-negative breast cancer, Approved February 2015

Lenvima (lenvatinib); Eisai; For the treatment of thyroid cancer, Approved February 2015

Lonsurf (trifluridine and tipiracil); Taiho Oncology; For the treatment of metastatic colorectal cancer , Approved September 2015

Odomzo (sonidegib); Novartis; For the treatment of locally advanced basal cell carcinoma, July 2015

Opdivo (nivolumab); Bristol-Myers Squibb; For the treatment of metastatic squamous non-small cell lung cancer, Approved March 2015

Unituxin (dinutuximab); United Therapeutics; For the treatment of pediatrics with high-risk neuroblastoma, Approved March 2015

Varubi (rolapitant); Tesaro; For the prevention of delayed nausea and vomiting associated with chemotherapy, Approved September 2015


Death Forces FDA to Place Clinical Hold on Advaxis (ADXS) Cancer Drug

from Biospace News

October 7, 2015
By Alex Keown, BioSpace.com Breaking News Staff

PRINCETON, N.J. – Following the death of a patient, the U.S. Food and Drug Administration (FDA) placed a hold on Advaxis (ADXS)’s experimental cancer treatment axalimogene filolisbac, which is currently in mid-stage trials.

In a statement issued this morning, Advaxis maintains the patient’s death was a result of the severity of her cancer and not due to the company’s experimental cancer treatment. It is seeking proof from the FDA that the drug was not a factor in the death. Still, the hold on the experimental cancer drug will cause the company to halt four clinical trials, Advaxis said. Other clinical trials, including those with the experimental ADXS-PSA and ADXS-HER2, are not affected by this hold. The company said it will continue to actively enroll and dose patients.

The FDA placed a hold on the drug on Oct. 2 after the company submitted a safety report to the regulatory agency that week. The drug is being developed to treat patients with persistent or recurrent metastatic (squamous or non-squamous cell) carcinoma of the cervix (PRmCC) who have progressed on at least one prior line of systemic therapy. Phase I trials released at the end of September showed treatment with axalimogene filolisbac resulted in a 38.5 percent 12-month overall survival rate in 26 patients. Patients typically fighting PRmCC who have failed at least one line of therapy have a typical survival rate of four to seven months.

Read full story here


FDA Halts Trial of Halozyme’s PEGPH20 for Pancreatic Cancer

Apr 9, 2014 Alex Philippidis

Halozyme Therapeutics acknowledged today that the FDA placed a formal clinical hold on its troubled Study 202 assessing its experimental drug PEGPH20 in patients with pancreatic cancer—less than a week after the company temporarily halted enrolling and dosing patients in the ongoing Phase II trial.

The agency told Halozyme it placed the clinical hold following the company’s pause in study activity. The trial’s independent data monitoring committee is evaluating data from the trial to learn why patients treated with PEGPH20 as well as nab-paclitaxel and gemcitabine saw a higher rate of blood clots and other thromboembolic events compared with patients treated with nab-paclitaxel and gemcitabine alone.

“We will be providing this information to the data monitoring committee and the FDA in parallel so they can complete their respective assessments,” Helen Torley, M.B. Ch.B., M.R.C.P., Halozyme’s president and CEO, said in a statement.

“Pancreatic cancer has one of the lowest survival rates of any cancer. We remain committed to evaluating PEGPH20 as a possible therapy to address this devastating disease,” Dr. Torley added.

As with Halozyme’s statement last week, the company’s latest remarks did not indicate when Halozyme expects to resume enrolling and dosing patients in Study 202, or how many patients had been enrolled and dosed when the temporary halt occurred.

The trial was envisioned as having 124 subjects, divided evenly between a treatment arm of PEGPH20 and nab-paclitaxel, and a gemcitabine arm, preceded by eight subject “run-in” phase assessing safety and tolerability, according to Study 202’s page on ClinicalTrials.gov (NCT01839487), last updated on January 27.

The study is one of two Phase II trials for PEGPH20; the other, SWOG, also aims to assess the drug for pancreatic cancer.

PEGPH20 is an investigational PEGylated form of Halozyme’s FDA-approved recombinant human hyaluronidase rHuPH20 (marketed as Hylenex®), designed to dramatically increases the half-life of the compound in the blood and allow for intravenous administration.

The temporary halt for Study 202 came two months after Halozyme publicly cited “potential acceleration of the PEGPH20 program” among several R&D programs for which it raised funds through a public offering of common stock that closed in February and generated approximately $107.8 million in net proceeds.

Read more at GenNEWS


FDA orders CytRx to halt patient enrollment after death of a cancer patient

CytRx ($CYTR) has run into an unexpected roadblock with its cancer drug conjugate aldoxorubicin, slamming the brakes on new patient recruitment in all their clinical trials after the FDA dropped a partial clinical hold on the program. According to the biotech the hold was forced by the death of a patient who was given the drug through a compassionate use program.

LA-based CytRx execs say that patients already enrolled in the studies will continue to receive the therapy as investigators added new safety measures, retooling trial protocols to include an “appropriate inclusion/exclusion criteria, an additional patient screening assessment and an evaluation of serum electrolytes prior to aldoxorubicin administration.” The patient who died, they added, had not qualified for any of its studies.

As it stands now, the biotech doesn’t know exactly how long the partial hold will last, but their announcement sought to calm jumpy investors, saying they expected to resolve the FDA’s demands “expeditiously” and can stick to their current timelines. CytRx says it expects to report preliminary results from their mid-stage study of Kaposi’s sarcoma in the second quarter of 2015 and preliminary results from the ongoing Phase II clinical trial of aldoxorubicin in glioblastoma multiforme in the first half of 2015. The company added that it is committed to completing enrollment in their Phase III trial by the end of next year.

hat reassurance appears to have helped with investors, who seemed to count this as more of a temporary setback than a catastrophe. Shares for CytRx were down about 9% in mid-morning trading.

Aldoxorubicin uses a linker molecule to attach to albumin in the blood and concentrate in tumors, where the acidic environment releases the chemotherapy doxorubicin in doses up to four times higher than what’s used now. Late last year their stock soared after their drug scored promising results for progression-free survival in a Phase IIb trial.

This case illustrates one reason why biotechs often quietly squirm under the pressure of compassionate use programs. They can be expensive to operate, time-consuming and raise fresh concerns when a patient dies or experiences a setback. On the other hand, if regulators take action like this following the death of an advanced stage cancer patient, there may have been something about the case that triggered broader concerns for the entire patient population


Clot risk in Lilly lung-cancer drug raises FDA concerns

July 7, 2015

Eli Lilly and Co.’s experimental lung cancer drug has raised concerns with U.S. regulators that it may increase patients’ risk of suffering potentially deadly blood clots.

The drug, known as necitumumab, improved patients’ overall chances of survival, yet people taking the medicine also experienced more risk, Food and Drug Administration staff said in a report Tuesday. Indianapolis-based Lilly is seeking to sell the medicine to treat a subset of the most common type of lung cancer.

FDA advisers will meet Thursday to discuss the risks and benefits of necitumumab for patients with advanced squamous non-small cell lung cancer, in combination with chemotherapy. The FDA is expected to decide if Lilly can sell the drug by the end of the year.

While the safety of necitumumab reflects that of similar drugs, the increased danger of clotting “in this already high risk population is of concern,” FDA staff wrote.

One study showed that out of 538 patients taking necitumumab and chemotherapy, 9 percent experienced a serious clot, compared with 5 percent of 541 patients given only chemotherapy, according to the staff report.

Squamous lung cancer accounts for 25 percent to 30 percent of all lung cancer, according to the American Cancer Society.

Patients in a clinical trial who took necitumumab lived a median of 11.5 months, 1.6 months longer than those who got only chemotherapy, the FDA staff report said.

Opdivo Side Effects Center (as seen on Rxlist.com) (NOTE:TRIAL NOT HALTED)

Last reviewed on RxList 10/05/2015

Opdivo (nivolumab) is a human monoclonal antibody used to treat patients with unresectable or metastatic melanoma and disease progression following ipilimumab and, if BRAF V600 mutation positive, a BRAF inhibitor; and to treat metastatic squamous non-small cell lung cancer with progression on or after platinum-based chemotherapy. Common side effects of Opdivo include fatigue, rash, itching, cough, upper respiratory tract infection, swelling of the extremities, shortness of breath, muscle pain, decreased appetite, nausea, vomiting, constipation, diarrhea, weakness, swelling, fever, abdominal pain, chest pain, joint pain, and weight loss.


Opdivo FDA Prescribing Information: Side Effects
(Adverse Reactions)

Clinical Trials Experience

Because clinical trials are conducted under widely varying conditions, adverse reaction rates observed in the clinical trials of a drug cannot be directly compared to rates in the clinical trials of another drug and may not reflect the rates observed in clinical practice.

The data described in the WARNINGS AND PRECAUTIONS section and below reflect exposure to OPDIVO in Trial 1, a randomized trial in patients with unresectable or metastatic melanoma and in Trial 3, a single-arm trial in patients with metastatic squamous non-small cell lung cancer (NSCLC).

Clinically significant adverse reactions were evaluated in a total of 691 patients enrolled in Trials 1, 3, or an additional dose finding study (n=306) administering OPDIVO at doses of 0.1 to 10 mg/kg every 2 weeks [see WARNINGS AND PRECAUTIONS].

Unresectable or Metastatic Melanoma

The safety of OPDIVO was evaluated in Trial 1, a randomized, open-label trial in which 370 patients with unresectable or metastatic melanoma received OPDIVO 3 mg/kg every 2 weeks (n=268) or investigator’s choice of chemotherapy (n=102), either dacarbazine 1000 mg/m² every 3 weeks or the combination of carboplatin AUC 6 every 3 weeks plus paclitaxel 175 mg/m² every 3 weeks [see Clinical Studies]. The median duration of exposure was 5.3 months (range: 1 day to 13.8+ months) with a median of eight doses (range: 1 to 31) in OPDIVO-treated patients and was 2 months (range: 1 day to 9.6+ months) in chemotherapy treated patients. In this ongoing trial, 24% of patients received OPDIVO for greater than 6 months and 3% of patients received OPDIVO for greater than 1 year.

In Trial 1, patients had documented disease progression following treatment with ipilimumab and, if BRAF V600 mutation positive, a BRAF inhibitor. The trial excluded patients with autoimmune disease, prior ipilimumab-related Grade 4 adverse reactions (except for endocrinopathies) or Grade 3 ipilimumab-related adverse reactions that had not resolved or were inadequately controlled within 12 weeks of the initiating event, patients with a condition requiring chronic systemic treatment with corticosteroids ( > 10 mg daily prednisone equivalent) or other immunosuppressive medications, a positive test for hepatitis B or C, and a history of HIV.

The study population characteristics in the OPDIVO group and the chemotherapy group were similar: 66% male, median age 59.5 years, 98% white, baseline ECOG performance status 0 (59%) or 1 (41%), 74% with M1c stage disease, 73% with cutaneous melanoma, 11% with mucosal melanoma, 73% received two or more prior therapies for advanced or metastatic disease, and 18% had brain metastasis. There were more patients in the OPDIVO group with elevated LDH at baseline (51% vs. 38%).

OPDIVO was discontinued for adverse reactions in 9% of patients. Twenty-six percent of patients receiving OPDIVO had a drug delay for an adverse reaction. Serious adverse reactions occurred in 41% of patients receiving OPDIVO. Grade 3 and 4 adverse reactions occurred in 42% of patients receiving OPDIVO. The most frequent Grade 3 and 4 adverse reactions reported in 2% to less than 5% of patients receiving OPDIVO were abdominal pain, hyponatremia, increased aspartate aminotransferase, and increased lipase.


FDA Approves Eisai’s LENVIMA™ (lenvatinib) for the Treatment of Patients with Locally Recurrent or Metastatic, Progressive, Radioactive Iodine-Refractory Differentiated Thyroid Cancer

– Press release from Eisai (NOTE: TRIAL NOT HALTED)

Feb 13, 2015

WOODCLIFF LAKE, N.J., Feb. 13, 2015 /PRNewswire/ — Eisai Inc. announced today that the U.S. Food and Drug Administration (FDA) approved the company’s receptor tyrosine kinase inhibitor LENVIMA™ (lenvatinib) for the treatment of locally recurrent or metastatic, progressive, radioactive iodine-refractory differentiated thyroid cancer (RAI-R DTC). LENVIMA was approved following a priority review by the FDA, which is designated for drugs the FDA believes have the potential to provide a significant improvement in the treatment of a serious condition. LENVIMA demonstrated a statistically significant progression-free survival (PFS) prolongation and response rate in patients with progressive, differentiated thyroid cancer who had become refractory to radioactive iodine (RAI) therapy.

In the clinical trial, adverse events led to dose reductions in 68% of patients who received LENVIMA and 5% of patients who received placebo. Some patients will need to discontinue treatment for serious adverse reactions. In the trial, 18% of patients treated with LENVIMA and 5% who received placebo discontinued treatment. The most common adverse reactions (at least 10%) that resulted in dose reductions of LENVIMA were hypertension (13%), proteinuria (11%), decreased appetite (10%), and diarrhea (10%).

AstraZeneca halts a pair of lung cancer trials over a safety scare

From October 9, 2015 | By of FierceBiotech

“AstraZeneca ($AZN) is pressing pause on trials combining two of its most important pipeline cancer treatments after tracking reports of lung disease, halting enrollment as it gathers more information.

The company is testing a combination of AZD9291 and durvalumab, formerly MEDI4736, in two studies involving patients with non-small cell lung cancer. Late last month, AstraZeneca hit the brakes on enrollment in both trials due to an increase in reports of interstitial lung disease, which can lead to dangerous scarring and impaired pulmonary function. The pauses are temporary, the company stressed in an emailed statement, and patients already enrolled in the study will be given new consent forms to ensure they understand the risks before choosing whether keep getting treatment.”

Other posts on this site on Cytotoxicity and Cancer include

Novel Approaches to Cancer Therapy [11.1]

Misfolded Proteins – from Little Villains to Little Helpers… Against Cancer

Multiple Lung Cancer Genomic Projects Suggest New Targets, Research Directions for Non-Small Cell Lung Cancer

A Synthesis of the Beauty and Complexity of How We View Cancer

Good and Bad News Reported for Ovarian Cancer Therapy

Read Full Post »

 

Horizon Discovery Group plc In-licenses Oncology Programme from Servier and Enters Option Agreement

Reporter: Aviva-Lev Ari, Ph.D., R.N.; Stephen J. Williams, Ph.D.

Press release

7 October 2015

 

Horizon Discovery Group plc In-licenses Oncology Programme from Servier and Enters Option Agreement

 

  • License programme with milestone payments of up to £50 million plus royalties on product sales
  • Horizon to in-license novel kinase inhibitor programme from Servier
  • Horizon will exploit its translational genomics and combination sciences platform to define optimum approaches to treatment and identify cancer patient populations most likely to respond

 

Cambridge, UK, 7 October 2015: Horizon Discovery Group plc (LSE: HZD) (“Horizon” or “the Company”), the international life science company supplying research tools and services that power genomics research and the development of personalised medicines, announces today that its leveraged business unit has signed a programme in-licensing and option agreement with Servier, the independent French research-based pharmaceutical company. The agreement is potentially worth over £50 million to Horizon in preclinical and clinical milestones, payments linked to net sales, and tiered royalties on future product sales.

 

Horizon has in-licensed novel kinase inhibitors from Servier that exhibit great promise based on pre-clinical data for treatment of a range of cancer types but do not currently have a biomarker to define a sensitive patient population. Horizon will use its world-leading platform, comprising isogenic cell lines and in vivo models, CRISPR-Cas9 mediated gene editing technology and ultra-high-throughput combination screening, to identify the population of cancer patients most likely to respond to the in-licensed compounds, whether as single agents or in combinations with other drugs. Horizon also has the option to explore the use of the inhibitors in other therapeutic indications.

 

Under the terms of the agreement, Servier has a first option to license back the assets. Should Servier take up this option, Horizon would receive up to £50 million in milestone payments plus royalties on product sales. If Servier does not take up its option, Horizon will be free to seek another pharma partner and Horizon and Servier would then share in the success of the progression of the programme as it advances into the clinic and registration.

 

Horizon will evaluate the mechanism of action of the candidate compounds, and will verify the patient stratification hypothesis by both in vitro and in vivo preclinical experiments. Horizon will also define a path towards the development of biomarkers for both patient stratification and drug efficacy.

 

Dr. Darrin M. Disley, Chief Executive Officer of Horizon Discovery Group plc, said: “The in-license of assets with a strong pre-clinical pedigree but do not yet have a clear clinical development strategy, represents a great opportunity for companies like Horizon. Demonstrating our scientific leadership through our translational genomics, drug combination and biomarker discovery platforms; we seek to identify genetic markers that predict drug sensitivity enabling programmes like this one to be progressed rapidly into the clinic for defined patient populations. This innovative deal, as part of our strategy to drive accelerated growth, offers significant upside potential for our investors built upon the leverage of our intellectual property, technology platforms and know-how.”

 

Mr Jean Pierre Abastado, Director of Oncology Innovation, Servier, commented: “The long standing collaboration between Servier and Vernalis has led to the discovery of novel kinase inhibitors. Horizon’s technology portfolio and expertise makes them ideally positioned to progress these drug candidates into the clinic and to investigate their potential for therapeutic efficacy both alone and in combination therapies. Servier is committed to driving therapeutic progress for the benefit of patients, with partnerships such as this playing a key role.”

ENDS

 

For further information from Horizon Discovery Group plc, please contact:

 

Zyme Communications (Trade and Regional Media)

Katie Odgaard

Tel: +44 (0)7787 502 947

Email: katie.odgaard@zymecommunications.com

 

Consilium Strategic Communications (Financial Media and Investor Relations)

Amber Fennell / Jessica Hodgson / Matthew Neal / Laura Thornton

Tel: +44 (0) 20 3709 5701

Email: horizon@consilium-comms.com

 

Panmure Gordon & Co. (NOMAD)

Corporate Finance: Freddy Crossley / Duncan Monteith / Fabien Holler

Broking: Tom Salvesen

Tel: +44 20 7886 2500

 

Notes for Editors

 

About Horizon Discovery Group plc www.horizondiscovery.com/

Horizon is a revenue-generating life science group supplying research tools to organisations engaged in genomics research and the development of personalised medicines. Horizon has a diverse and international customer base of over 1,200 organisations across more than 50 countries, including major pharmaceutical, biotechnology and diagnostic companies as well as leading academic research centers. The Group supplies its products and services into multiple markets, estimated to total in excess of £29 billion by 2015.

 

Horizon’s core capabilities are built around its proprietary translational genomics platform, a high-precision and flexible suite of gene editing tools able to alter almost any endogenous gene sequence of human or mammalian cell-lines. Horizon offers over 20,000 catalogue products, almost all of which are based on the application of gene editing to generate cell lines that accurately model the disease-causing mutations found in genetically based diseases. These ‘patients-in-a-test-tube’ are being used by customers to identify the effect of individual or compound genetic mutations on drug activity, patient responsiveness, and resistance, which may lead to the successful prediction of which patient sub-groups will respond to currently available and future drug treatments.

In addition, Horizon provides custom cell line and in vivo model generation services for research and bioproduction applications, quantitative molecular reference standards, in vivo disease models, and contract research and custom screening services.

 

Horizon is headquartered in Cambridge, UK, and is listed on the London Stock Exchange’s AIM market under the ticker “HZD”, for further information please visit: www.horizondiscovery.com.

 

About Servier

Servier is an independent French research-based pharmaceutical company. Its development is driven by the pursuit of innovation in the therapeutic areas of cardiovascular, metabolic, central nervous system, psychiatric, bone, muscle and joint diseases, as well as cancer.

  • In 2014, the company recorded a turnover of 4 billion euros.
  • 92% of Servier medicines are prescribed outside of France.
  • 28% of turnover from Servier drugs was reinvested in Research and Development in 2014.

With a strong international presence in 146 countries, Servier employs more than 21,400 people worldwide.

 

Oncology is one of the key priorities of Servier in terms of research and development with currently 8 new molecular entities in clinical development in breast cancer, lung cancer, other solid tumours and various types of lymphomas and leukaemias. This portfolio of innovative cancer treatments is being developed with various partners worldwide, and covers different hallmarks of cancer including cytotoxics, pro-apoptotic, targeted, immune and cellular therapies. Hence, Servier aims at delivering a significant and positive impact on cancer patients’ lives.

Read Full Post »

Development of Chemoresistance to Targeted Therapies: Alterations of Cell Signaling, & the Kinome [11.4.1.2]

 

Curator, Reporter: Stephen J. Williams, Ph.D.

The advent of molecular targeted therapies like Imatinib (Gleevec), and other tyrosine kinase inhibitors (TKI) has been transformative to cancer therapy. However, as with all chemotherapeutics, including radiation therapy, the development of chemo-resistance toward personalized, molecular therapies has been disastrous to the successful treatment of cancer. The fact that chemo-resistance develops to personalized therapies was a serious disappointment to clinicians (although most expected this to be the case) but more surprisingly it was the rapidity of onset and speed of early reported cases which may have been the biggest shocker.

A post on resistance to other TKIs (to EGFR and ALK) can be seen here: http://pharmaceuticalintelligence.com/2013/11/01/resistance-to-receptor-of-tyrosine-kinase/

History of Development of Resistance to Imatinib (Gleevec)

The Melo group published a paper in Blood showing that short exposure to STI571 (imatinib; trade name Gleevec®) could result in drug resistant clones

Selection and characterization of BCR-ABL positive cell lines with differential sensitivity to the tyrosine kinase inhibitor STI571: diverse mechanisms of resistance. Blood. 2000 Aug 1;96(3):1070-9.

Mahon FX1, Deininger MW, Schultheis B, Chabrol J, Reiffers J, Goldman JM, Melo JV.

Abstract

Targeting the tyrosine kinase activity of Bcr-Abl with STI571 is an attractive therapeutic strategy in chronic myelogenous leukemia (CML). A few CML cell lines and primary progenitors are, however, resistant to this compound. We investigated the mechanism of this resistance in clones of the murine BaF/3 cells transfected with BCR-ABL and in 4 human cell lines from which sensitive (s) and resistant (r) clones were generated by various methods. Although the resistant cells were able to survive in the presence of STI571, their proliferation was approximately 30% lower than that of their sensitive counterparts in the absence of the compound. The concentration of STI571 needed for a 50% reduction in viable cells after a 3-day exposure was on average 10 times higher in the resistant (2-3 micromol/L) than in the sensitive (0.2-0.25 micromol/L) clones. The mechanism of resistance to STI571 varied among the cell lines. Thus, in Baf/BCR-ABL-r, LAMA84-r, and AR230-r, there was up-regulation of the Bcr-Abl protein associated with amplification of the BCR-ABL gene. In K562-r, there was no Bcr-Abl overexpression, but the IC(50) for the inhibition of Bcr-Abl autophosphorylation was increased in the resistant clones. Sequencing of the Abl kinase domain revealed no mutations. The multidrug resistance P-glycoprotein (Pgp) was overexpressed in LAMA84-r, indicating that at least 2 mechanisms of resistance operate in this cell line. KCL22-r showed neither Bcr-Abl up-regulation nor a higher threshold for tyrosine kinase inhibition by STI571. We conclude that BCR-ABL-positive cells can evade the inhibitory effect of STI571 by different mechanisms, such as Bcr-Abl overexpression, reduced intake mediated by Pgp, and, possibly, acquisition of compensatory mutations in genes other than BCR-ABL.

mellobcrablresistamplification

FISH analysis of AR230 and LAMA84 sensitive and resistant clones, with probes for the ABL (red signal) and theBCR (green signal) genes. BCR-ABL is identified as a red–green or yellow fused signal. Adapted from Mahon et al., Blood 2000; 96(3):1070-9.

This rapid onset of imatinib resistance also see in the clinic and more prominent in advance disease

From NCCN 2nd Annual Congress: Hematologic Malignancies – Update on Primary Therapy, Second-Line Therapy, and New Agents for Chronic Myelogenous Leukemia (Slides with Transcript)

http://www.medscape.org/viewarticle/564097

There is some evidence that even looking earlier makes some sense in determining what the prognosis is. This is from Timothy Hughes’ group in Adelaide, and he is looking at an earlier molecular time point, 3 months after initiation of therapy. And what you have done here is you have taken the 3-month mark and you have said, “Well, based on your response at 3 months, what is your likelihood that in the future you will either get a major molecular response or become resistant?”

3monthimitanibresist

If you look at the accumulation of imatinib resistance to find if it is either initially not responding or becoming resistant after a good response, it goes up with type of disease and phase of disease. So if you look at patients who have early chronic phase disease — that is, they start getting imatinib less than a year from the diagnosis — their chance of failure is pretty low. With later disease — they are in a chronic phase but they have had disease more than a year before they get imatinib — it is higher. If you see patients with accelerated phase or blast crisis, the chances are that they will fail sometime in the future.

speed of imitinib resistance

Therefore, because not all resistant samples show gene amplification of Bcr/Abl and the rapidity of onset of resistance, many feel that there are other mechanisms of resistance at play, like kinome plasticity.

Kinome Plasticity Contributes to TKI resistance

Beyond gene amplification, other mechanisms of imitanib and other tyrosine kinase inhibitors (TKI) include alterations in compensatory signaling pathways. This can be referred to as kinome plasticity and is explained in the following abstracts from the AACR 2015 meeting.

Systems-pharmacology dissection of a drug synergy in imatinib-resistant CML

Georg E Winter, Uwe Rix, Scott M Carlson, Karoline V Gleixner, Florian Grebien, Manuela Gridling, André C Müller, Florian P Breitwieser, Martin Bilban, Jacques Colinge, Peter Valent, Keiryn L Bennett, Forest M White & Giulio Superti-Furga. Nature Chemical Biology 8,905–912(2012)

Occurrence of the BCR-ABLT315I gatekeeper mutation is among the most pressing challenges in the therapy of chronic myeloid leukemia (CML). Several BCR-ABL inhibitors have multiple targets and pleiotropic effects that could be exploited for their synergistic potential. Testing combinations of such kinase inhibitors identified a strong synergy between danusertib and bosutinib that exclusively affected CML cells harboring BCR-ABLT315I. To elucidate the underlying mechanisms, we applied a systems-level approach comprising phosphoproteomics, transcriptomics and chemical proteomics. Data integration revealed that both compounds targeted Mapk pathways downstream of BCR-ABL, resulting in impaired activity of c-Myc. Using pharmacological validation, we assessed that the relative contributions of danusertib and bosutinib could be mimicked individually by Mapk inhibitors and collectively by downregulation of c-Myc through Brd4 inhibition. Thus, integration of genome- and proteome-wide technologies enabled the elucidation of the mechanism by which a new drug synergy targets the dependency of BCR-ABLT315I CML cells on c-Myc through nonobvious off targets.

nchembio.1085-F2kinomegleevecresistance

Please see VIDEO and SLIDESHARE of a roundtable Expert Discussion on CML

Curated Content From the 2015 AACR National Meeting on Drug Resistance Mechanisms and tyrosine kinase inhibitors

Session Title: Mechanisms of Resistance: From Signaling Pathways to Stem Cells
Session Type: Major Symposium
Session Start/End Time: Tuesday, Apr 21, 2015, 10:30 AM -12:30 PM
Location: Terrace Ballroom II-III (400 Level), Pennsylvania Convention Center
CME: CME-Designated
CME/CE Hours: 2
Session Description: Even the most effective cancer therapies are limited due to the development of one or more resistance mechanisms. Acquired resistance to targeted therapies can, in some cases, be attributed to the selective propagation of a small population of intrinsically resistant cells. However, there is also evidence that cancer drugs themselves can drive resistance by triggering the biochemical- or genetic-reprogramming of cells within the tumor or its microenvironment. Therefore, understanding drug resistance at the molecular and biological levels may enable the selection of specific drug combinations to counteract these adaptive responses. This symposium will explore some of the recent advances addressing the molecular basis of cancer cell drug resistance. We will address how tumor cell signaling pathways become rewired to facilitate tumor cell survival in the face of some of our most promising cancer drugs. Another topic to be discussed involves how drugs select for or induce the reprogramming of tumor cells toward a stem-like, drug resistant fate. By targeting the molecular driver(s) of rewired signaling pathways and/or cancer stemness it may be possible to select drug combinations that prevent the reprogramming of tumors and thereby delay or eliminate the onset of drug resistance.
Presentations:
Chairperson
Tuesday, Apr 21, 2015, 10:30 AM -12:30 PM
David A. Cheresh. UCSD Moores Cancer Center, La Jolla, CA
Introduction
Tuesday, Apr 21, 2015, 10:30 AM -10:40 AM
Resistance to tyrosine kinase inhibitors: Heterogeneity and therapeutic strategies.
Tuesday, Apr 21, 2015, 10:40 AM -10:55 AM
Jeffrey A. Engelman. Massachusetts General Hospital, Boston, MA
Discussion
Tuesday, Apr 21, 2015, 10:55 AM -11:00 AM
NG04: Clinical acquired resistance to RAF inhibitor combinations in BRAF mutant colorectal cancer through MAPK pathway alterations
Tuesday, Apr 21, 2015, 11:00 AM -11:15 AM
Ryan B. Corcoran, Leanne G. Ahronian, Eliezer Van Allen, Erin M. Coffee, Nikhil Wagle, Eunice L. Kwak, Jason E. Faris, A. John Iafrate, Levi A. Garraway, Jeffrey A. Engelman. Massachusetts General Hospital Cancer Center, Boston, MA, Dana-Farber Cancer Institute, Boston, MA
Discussion
Tuesday, Apr 21, 2015, 11:15 AM -11:20 AM
SY27-02: Tumour heterogeneity and therapy resistance in melanoma
Tuesday, Apr 21, 2015, 11:20 AM -11:35 AM
Claudia Wellbrock. Univ. of Manchester, Manchester, United Kingdom
Discussion
Tuesday, Apr 21, 2015, 11:35 AM -11:40 AM
SY27-03: Breast cancer stem cell state transitions mediate therapeutic resistance
Tuesday, Apr 21, 2015, 11:40 AM -11:55 AM
Max S. Wicha. University of Michigan, Comprehensive Cancer Center, Ann Arbor, MI
Discussion
Tuesday, Apr 21, 2015, 11:55 AM -12:00 PM
SY27-04: Induction of cancer stemness and drug resistance by EGFR blockade
Tuesday, Apr 21, 2015, 12:00 PM -12:15 PM
David A. Cheresh. UCSD Moores Cancer Center, La Jolla, CA
Discussion
Tuesday, Apr 21, 2015, 12:15 PM -12:20 PM
General Discussion
Tuesday, Apr 21, 2015, 12:20 PM -12:30 PM

Targeting Macromolecular Signaling Complexes 
Room 115, Pennsylvania Convention Center

Drug Resistance 
Hall A (200 Level), Pennsylvania Convention Center
Resistance to Pathway-Targeted Therapeutics 1 
Section 33

Molecular Mechanisms of Sensitivity or Resistance to Pathway-Targeted Agents 
Room 118, Pennsylvania Convention Center

Targeting Signaling Pathways in Cancer 
Room 204, Pennsylvania Convention Center
Exploiting the MAPK Pathway in Cancer 
Room 115, Pennsylvania Convention Center

PLEASE see the attached WORD file which includes ALL abstracts, posters, and talks on this subject from the AACR 2015 national meeting BELOW

 AACR2015resistancekinome

Other posts related to, Cancer, Chemotherapy, Gleevec and Resistance on this Open Access Journal Include

Imatinib (Gleevec) May Help Treat Aggressive Lymphoma: Chronic Lymphocytic Leukemia (CLL)

Treatments for Acute Leukemias [2.4.4A]

Therapeutic Implications for Targeted Therapy from the Resurgence of Warburg ‘Hypothesis’

Hematologic Malignancies [6.2]

Overview of Posttranslational Modification (PTM)

Novel Modeling Methods for Genomic Data Analysis & Evolutionary Systems Biology to Design Dosing Regimens to Minimize Resistance

Mechanisms of Drug Resistance

Using RNA-seq and targeted nucleases to identify mechanisms of drug resistance in acute myeloid leukemia

An alternative approach to overcoming the apoptotic resistance of pancreatic cancer

Resistance to Receptor of Tyrosine Kinase

Read Full Post »

« Newer Posts - Older Posts »