Funding, Deals & Partnerships: BIOLOGICS & MEDICAL DEVICES; BioMed e-Series; Medicine and Life Sciences Scientific Journal – http://PharmaceuticalIntelligence.com
Lesson 4 Cell Signaling And Motility: G Proteins, Signal Transduction: Curations and Articles of reference as supplemental information: #TUBiol3373
Curator: Stephen J. Williams, Ph.D.
Updated 7/15/2019
Below please find the link to the Powerpoint presentation for lesson #4 for #TUBiol3373. The lesson first competes the discussion on G Protein Coupled Receptors, including how cells terminate cell signals. Included are mechanisms of receptor desensitization. Please NOTE that desensitization mechanisms like B arrestin decoupling of G proteins and receptor endocytosis occur after REPEATED and HIGH exposures to agonist. Hydrolysis of GTP of the alpha subunit of G proteins, removal of agonist, and the action of phosphodiesterase on the second messenger (cAMP or cGMP) is what results in the downslope of the effect curve, the termination of the signal after agonist-receptor interaction.
New Studies toward Understanding Alzheimer Disease
Curators: Larry H. Bernstein, MD, FCAP and Aviva Lev-Ari, PhD, RN
There is no unifying concept of Alzheimer Disease beyond the Tau and beta amyloid roles. Recently, Ingenbleek and Bernstein (journal AD) made the connection between the age related decline of liver synthesis of plasma transthyretin and the more dramatic decline of transthyretin at the blood brain barrier, and the relationship to inability to transfer vitamin A via retinol binding protein to the brain. Related metabolic events are reported by several groups.
They show that Aβ oligomerization, a behavior traditionally viewed as intrinsically pathological, may be necessary for the antimicrobial activities of the peptide. Collectively, our data are consistent with a model in which soluble Aβ oligomers first bind to microbial cell wall carbohydrates via a heparin-binding domain. Developing protofibrils inhibited pathogen adhesion to host cells. Propagating β-amyloid fibrils mediate agglutination and eventual entrapment of unatttached microbes….Salmonella Typhimurium bacterial infection of the brains of transgenic 5XFAD mice resulted in rapid seeding and accelerated β-amyloid deposition, which closely colocalized with the invading bacteria.
This is quite interesting in that infection drives the production of acute phase reactants resulting in decreased production of transthyretin. Whether this also has ties to chronic disease in the elderly and risk of AD is not known.
Through whole-genome sequencing of 1345 individuals from 410 families with late-onset AD (LOAD), they identified three highly penetrant variants in PRKCA, the gene that encodes protein kinase Cα (PKCα), in five of the families. All three variants linked with LOAD displayed increased catalytic activity relative to wild-type PKCα as assessed in live-cell imaging experiments using a genetically encoded PKC activity reporter. Deleting PRKCA in mice or adding PKC antagonists to mouse hippocampal slices infected with a virus expressing the Aβ precursor CT100 revealed that PKCα was required for the reduced synaptic activity caused by Aβ. In PRKCA(-/-) neurons expressing CT100, introduction of PKCα, but not PKCα lacking a PDZ interaction moiety, rescued synaptic depression, suggesting that a scaffolding interaction bringing PKCα to the synapse is required for its mediation of the effects of Aβ. Thus, enhanced PKCα activity may contribute to AD, possibly by mediating the actions of Aβ on synapses.
Kim BM, You MH, Chen CH, Suh J, Tanzi RE, Ho Lee T.
Hum Mol Genet. 2016 Apr 19. pii: ddw114.
Extracellular deposition of amyloid-beta (Aβ) peptide, a metabolite of sequential cleavage of amyloid precursor protein (APP), is a critical step in the pathogenesis of Alzheimer’s disease (AD). While death-associated protein kinase 1 (DAPK1) is highly expressed in AD brains and its genetic variants are linked to AD risk, little is known about the impact of DAPK1 on APP metabolism and Aβ generation. This study demonstrated a novel effect of DAPK1 in the regulation of APP processing using cell culture and mouse models. DAPK1, but not its kinase deficient mutant (K42A), significantly increased human Aβ secretion in neuronal cell culture models. Moreover, knockdown of DAPK1 expression or inhibition of DAPK1 catalytic activity significantly decreased Aβ secretion. Furthermore, DAPK1, but not K42A, triggered Thr668 phosphorylation of APP, which may initiate and facilitate amyloidogenic APP processing leading to the generation of Aβ. In Tg2576 APPswe-overexpressing mice, knockout of DAPK1 shifted APP processing toward non-amyloidogenic pathway and decreased Aβ generation. Finally, in AD brains, elevated DAPK1 levels showed co-relation with the increase of APP phosphorylation. Combined together, these results suggest that DAPK1 promotes the phosphorylation and amyloidogenic processing of APP, and that may serve a potential therapeutic target for AD.
The “amyloid β hypothesis” of Alzheimer’s disease (AD) has been the reigning hypothesis explaining pathogenic mechanisms of AD over the last two decades. However, this hypothesis has not been fully validated in animal models, and several major unresolved issues remain. Our 3D human neural cell culture model system provides a premise for a new generation of cellular AD models that can serve as a novel platform for studying pathogenic mechanisms and for high-throughput drug screening in a human brain-like environment.
The two key pathological hallmarks of AD are senile plaques (amyloid plaques) and neurofibrillary tangles (NFTs), which develop in brain regions responsible for memory and cognitive functions (i.e. cerebral cortex and limbic system) 3. Senile plaques are extracellular deposits of amyloid-β (Aβ) peptides, while NFTs are intracellular, filamentous aggregates of hyperphosphorylated tau protein 4.
The identification of Aβ as the main component of senile plaques by Drs. Glenner and Wong in 1984 5 resulted in the original formation of the “amyloid hypothesis.” According to this hypothesis, which was later renamed the “amyloid-β cascade hypothesis” by Drs. Hardy and Higgins 6, the accumulation of Aβ is the initial pathological trigger in the disease, subsequently leading to hyperphosphorylation of tau, causing NFTs, and ultimately, neuronal death and dementia 4,7–10. Although the details have been modified to reflect new findings, the core elements of this hypothesis remain unchanged: excess accumulation of the pathogenic forms of Aβ, by altered Aβ production and/or clearance, triggers the vicious pathogenic cascades that eventually lead to NFTs and neuronal death.
Over the last two decades, the Aβ hypothesis of AD has reigned, providing the foundation for numerous basic studies and clinical trials 4,7,10,11. According to this hypothesis, the accumulation of Aβ, either by altered Aβ production and/or clearance, is the initial pathological trigger in the disease. The excess accumulation of Aβ then elicits a pathogenic cascade including synaptic deficits, altered neuronal activity, inflammation, oxidative stress, neuronal injury, hyperphosphorylation of tau causing NFTs and ultimately, neuronal death and dementia 4,7–10.
One of the major unresolved issues of the Aβ hypothesis is to show a direct causal link between Aβ and NFTs 12–14. Studies have demonstrated that treatments with various forms of soluble Aβ oligomers induced synaptic deficits and neuronal injury, as well as hyperphosphorylation of tau proteins, in mouse and rat neurons, which could lead to NFTs and neurodegeneration in vivo18–21. However, transgenic AD mouse models carrying single or multiple human familial AD (FAD) mutations in amyloid precursor protein (APP) and/or presenilin 1 (PS1) do not develop NFTs or robust neurodegeneration as observed in human patients, despite robust Aβ deposition 13,22,23. Double and triple transgenic mouse models, harboring both FAD and tau mutations linked with frontotemporal dementia (FTD), are the only rodent models to date displaying both amyloid plaques and NFTs. However, the NFT pathology in these models stems mainly from the overexpression of human tau as a result of the FTD, rather than the FAD mutations24,25.
Human neurons carrying FAD mutations are an optimal model to test whether elevated levels of pathogenic Aβ trigger pathogenic cascades including NFTs, since those cells truly share the same genetic background that induces FAD in humans. Indeed, Israel et al., observed elevated tau phosphorylation in neurons with an APP duplication FAD mutation 33. Blocking Aβ generation by β-secretase inhibitors significantly decreased tau phosphorylation in the same model, but γ-secretase inhibitor, another Aβ blocker, did not affect tau phosphorylation 33. Neurons with the APP V717I FAD mutation also showed an increase in levels of phospho tau and total tau levels 28. More importantly, Muratore and colleagues showed that treatments with Aβ-neutralizing antibodies in those cells significantly reduced the elevated total and phospho tau levels at the early stages of differentiation, suggesting that blocking pathogenic Aβ can reverse the abnormal tau accumulation in APP V717I neurons 28.
Recently, Moore et al. also reported that neurons harboring the APP V717I or the APP duplication FAD mutation showed increases in both total and phospho tau levels 27. Interestingly, altered tau levels were not detected in human neurons carrying PS1 FAD mutations, which significantly increased pathogenic Aβ42 species in the same cells 27. These data suggest that elevated tau levels in these models were not due to extracellular Aβ accumulation but may possibly represent a very early stage of tauopathy. It may also be due to developmental alterations induced by the APP FAD mutations.
As summarized, most human FAD neurons showed significant increases in pathogenic Aβ species, while only APP FAD neurons showed altered tau metabolism that may represent very early stages of tauopathy. However, all of these human FAD neurons failed to recapitulate robust extracellular amyloid plaques, NFTs, or any signs of neuronal death, as predicted in the amyloid hypothesis.
In our recent study, we moved one step closer to proving the amyloid hypothesis. By generating human neural stem cell lines carrying multiple mutations in APP together with PS1, we achieved high levels of pathogenic Aβ42 comparable to those in brains of AD patients 44–46.
Platform for AD drug screening in human neural progenitor cells with FAD mutations in a 3D culture system, which successfully reproduce human AD pathogenesis (amyloid plaques-driven tauopathy).
In addition to the impact on toxic Aβ species, our 3D culture model can test if these antibodies can block tau pathologies in 3D human neural cell culture systems 44–46. Human cellular AD models can also be used to determine optimal doses of candidate AD drugs to block Aβ and/or tau pathology without affecting neuronal survival (Fig. 1).
While much progress has been made, many challenges still lie on the path to creating human neural cell culture models that comprehensively recapitulate pathogenic cascades of AD. A major difficulty lies in reconstituting the brain regions most affected in AD: the hippocampus and specific cortical layers. Recent progress in 3D culture technology, such as “cerebral organoids,” may also be helpful in rebuilding the brain structures that are affected by AD in a dish 52,53. These “cerebral organoids” were able to model various discrete brain regions including human cortical areas 52, which enabled them to reproduce microcephaly, a brain developmental disorder. Similarly, pathogenic cascades of AD may be recapitulated in cortex-like structures using this model. Adding neuroinflammatory components, such as microglial cells, which are critical in AD pathogenesis, will illuminate the validity of the amyloid β hypothesis. Reconstitution of robust neuronal death stemming from Aβ and tau pathologies will be the next major step in comprehensively recapitulating AD in a cellular model.
Natunen T, Takalo M, Kemppainen S, Leskelä S, Marttinen M, Kurkinen KM, Pursiheimo JP, Sarajärvi T, Viswanathan J, Gabbouj S, Solje E, Tahvanainen E, Pirttimäki T, Kurki M, Paananen J, Rauramaa T, Miettinen P, Mäkinen P, Leinonen V, Soininen H, Airenne K, Tanzi RE, Tanila H, Haapasalo A, Hiltunen M.
Accumulation of β-amyloid (Aβ) and phosphorylated tau in the brain are central events underlying Alzheimer’s disease (AD) pathogenesis. Aβ is generated from amyloid precursor protein (APP) by β-site APP-cleaving enzyme 1 (BACE1) and γ-secretase-mediated cleavages. Ubiquilin-1, a ubiquitin-like protein, genetically associates with AD and affects APP trafficking, processing and degradation. Here, we have investigated ubiquilin-1 expression in human brain in relation to AD-related neurofibrillary pathology and the effects of ubiquilin-1 overexpression on BACE1, tau, neuroinflammation, and neuronal viability in vitro in co-cultures of mouse embryonic primary cortical neurons and microglial cells under acute neuroinflammation as well as neuronal cell lines, and in vivo in the brain of APdE9 transgenic mice at the early phase of the development of Aβ pathology. Ubiquilin-1 expression was decreased in human temporal cortex in relation to the early stages of AD-related neurofibrillary pathology (Braak stages 0-II vs. III-IV). There was a trend towards a positive correlation between ubiquilin-1 and BACE1 protein levels. Consistent with this, ubiquilin-1 overexpression in the neuron-microglia co-cultures with or without the induction of neuroinflammation resulted in a significant increase in endogenously expressed BACE1 levels. Sustained ubiquilin-1 overexpression in the brain of APdE9 mice resulted in a moderate, but insignificant increase in endogenous BACE1 levels and activity, coinciding with increased levels of soluble Aβ40 and Aβ42. BACE1 levels were also significantly increased in neuronal cells co-overexpressing ubiquilin-1 and BACE1. Ubiquilin-1 overexpression led to the stabilization of BACE1 protein levels, potentially through a mechanism involving decreased degradation in the lysosomal compartment. Ubiquilin-1 overexpression did not significantly affect the neuroinflammation response, but decreased neuronal viability in the neuron-microglia co-cultures under neuroinflammation. Taken together, these results suggest that ubiquilin-1 may mechanistically participate in AD molecular pathogenesis by affecting BACE1 and thereby APP processing and Aβ accumulation.
PHASE 1 Type 2 DIABETES, IC50 value: 0.24± 0.0019 uM (SC1.5) [1] Target: Glucokinase activator
The R stereoisomer Ro 28-1675 activated GK with a SC1.5 of 0.24 uM, while the S isomer did not activated GK up to 10 uM. Oral administration of Ro 28-1675 (50 mg/Kg) to male C57B1/6J mice caused a statistically significant reduction in fasting glucose levels and improvement in glucose tolerance relative to the vehicle treated animals [1].
Comparison of rat PK parameters indicated that Ro 28-1675 displayed lower clearance and higher oral bioavailability compared to 9a.
Following a single oral dose, Ro 28-1675 reduced fasting and postprandial glucose levels following an OGTT, was well tolerated, and displayed no adverse effects related to drug administration other than hypoglycemia at the maximum dose (400 mg).
RO-28-1675 as glucokinase activator.
Joseph Grimsbyet al., of Roche have recently discovered activators of glucokinase that increase kcat and decrease the S0.5 for glucose, and these may offer a treatment for type II diabetes. Glucokinase (GK) plays a key role in whole-body glucose homeostasis by catalyzing the phosphorylation of glucose in cells that express this enzyme, such as pancreatic β cells and hepatocytes.
By screening of a library of 120,000 structurally diverse synthetic compounds, they found one small molecule that increased the enzymatic activity of GK. Chemical optimization of this initial molecule led to the synthesis of RO-28-0450 as a lead GK activator which is a class of antidiabetic agents that act as nonessential, mixed-type GK activators (GKAs) that increase the glucose affinity and maximum velocity (Vmax) of GK. RO-28-0450 is a racemic compound.
Activation of GK was exquisitely sensitive to the chirality of the molecule: The R enantiomer, RO-28-1675, was found to be a potent GKA, whereas the S enantiomer, RO-28-1674, was inactive. RO-28-1675 also reversed the inhibitory action of the human glucokinase regulatory protein (GKRP). The activators binding in a glucokinase regulatory site originally was discovered in patients with persistent hyperinsulinemic hypoglycemi.
The result of RO-28-1675 as a potent small molecule GKA may shed light to the chemical biologists to devise strategy for developing activators. Thus for a success to this end we must focus on highly regulated enzymes, or cooperative enzymes such as glucokinase, where nature has provided binding sites that are designed to modulate catalysis.
Glucokinase (GK) is a glucose sensor that couples glucose metabolism to insulin release. The important role of GK in maintaining glucose homeostasis is illustrated in patients with GK mutations. In this publication, identification of the hit molecule 1 and its SAR development, which led to the discovery of potent allosteric GK activators 9a and21a, is described. Compound 21a (RO0281675) was used to validate the clinical relevance of targeting GK to treat type 2 diabetes.
Flash chromatography (Merck Silica gel 60, 70-230 mesh, 9/1, 3/1, and then 11/9 hexanes/ethyl acetate) afforded (2R)-3-cyclopentyl-2-(4-methanesulfonylphenyl)-N-thiazol-2-yl-propionamide (2.10 g, 74%) as a white foam. ….
Glucokinase (GK) is a glucose sensor that couples glucose metabolism to insulin release. The important role of GK in maintaining glucose homeostasis is illustrated in patients with GK mutations. In this publication, identification of the hit molecule 1 and its SAR development, which led to the discovery of potent allosteric GK activators 9a and 21a, is described. Compound 21a (RO0281675) was used to validate the clinical relevance of targeting GK to treat type 2 diabetes.
Glucokinase (GK) is a glucose sensor that couples glucose metabolism to insulin release. The important role of GK in maintaining glucose homeostasis is illustrated in patients with GK mutations. In this publication, identification of the hit molecule 1 and its SAR development, which led to the discovery of potent allosteric GK activators 9a and 21a, is described. Compound 21a (RO0281675) was used to validate the clinical relevance of targeting GK to treat type 2 diabetes.
Rhizen Pharmaceuticals is developing RP-6530, a PI3K delta and gamma dual inhibitor, for the potential oral treatment of cancer and inflammation In November 2013, a phase I trial in patients with hematologic malignancies was initiated in Italy ]\. In September 2015, a phase I/Ib study was initiated in the US, in patients with relapsed and refractory T-cell lymphoma. At that time, the study was expected to complete in December 2016
PI3K delta/gamma inhibitor RP6530 An orally active, highly selective, small molecule inhibitor of the delta and gamma isoforms of phosphoinositide-3 kinase (PI3K) with potential immunomodulating and antineoplastic activities. Upon administration, PI3K delta/gamma inhibitor RP6530 inhibits the PI3K delta and gamma isoforms and prevents the activation of the PI3K/AKT-mediated signaling pathway. This may lead to a reduction in cellular proliferation in PI3K delta/gamma-expressing tumor cells. In addition, this agent modulates inflammatory responses through various mechanisms, including the inhibition of both the release of reactive oxygen species (ROS) from neutrophils and tumor necrosis factor (TNF)-alpha activity. Unlike other isoforms of PI3K, the delta and gamma isoforms are overexpressed primarily in hematologic malignancies and in inflammatory and autoimmune diseases. By selectively targeting these isoforms, PI3K signaling in normal, non-neoplastic cells is minimally impacted or not affected at all, which minimizes the side effect profile for this agent. Check for active clinical trials using this agent. (NCI Thesaurus)
Company
Rhizen Pharmaceuticals S.A.
Description
Dual phosphoinositide 3-kinase (PI3K) delta and gamma inhibitor
Dual PI3Kδ/γ Inhibition By RP6530 Induces Apoptosis and Cytotoxicity In B-Lymphoma Cells
Swaroop Vakkalanka, PhD*,1, Srikant Viswanadha, Ph.D.*,2, Eugenio Gaudio, PhD*,3, Emanuele Zucca, MD4, Francesco Bertoni, MD5, Elena Bernasconi, B.Sc.*,3, Davide Rossi, MD, Ph.D.*,6, and Anastasios Stathis, MD*,7
1Rhizen Pharmaceuticals S A, La Chaux-de-Fonds, Switzerland, 2Incozen Therapeutics Pvt. Ltd., Hyderabad, India, 3Lymphoma & Genomics Research Program, IOR-Institute of Oncology Research, Bellinzona, Switzerland, 4IOSI Oncology Institute of Southern Switzerland, Bellinzona, Switzerland, 5Lymphoma Unit, IOSI-Oncology Institute of Southern Switzerland, Bellinzona, Switzerland, 6Italian Multiple Myeloma Network, GIMEMA, Italy, 7Oncology Institute of Southern Switzerland, Bellinzona, Switzerland
RP6530 is a potent and selective dual PI3Kδ/γ inhibitor that inhibited growth of B-cell lymphoma cell lines with a concomitant reduction in the downstream biomarker, pAKT. Additionally, the compound showed cytotoxicity in a panel of lymphoma primary cells. Findings provide a rationale for future clinical trials in B-cell malignancies.
Blood 2013 122:4411; published ahead of print December 6, 2013
Swaroop Vakkalanka, Srikant Viswanadha, Eugenio Gaudio, Emanuele Zucca, Francesco Bertoni, Elena Bernasconi, Davide Rossi, Anastasios Stathis
Dual PI3K delta/gamma Inhibition By RP6530 Induces Apoptosis and Cytotoxicity
RP6530, a novel, small molecule PI3K delta/gamma
Activity and selectivity of RP6530 for PI3K delta and gamma isoforms
Introduction Activation of the PI3K pathway triggers multiple events including cell growth, cell cycle entry, cell survival and motility. While α and β isoforms are ubiquitous in their distribution, expression of δ and γ is restricted to cells of the hematopoietic system. Because these isoforms contribute to the development, maintenance, transformation, and proliferation of immune cells, dual targeting of PI3Kδ and γ represents a promising approach in the treatment of lymphomas. The objective of the experiments was to explore the therapeutic potential of RP6530, a novel, small molecule PI3Kδ/γ inhibitor, in B-cell lymphomas.
Methods Activity and selectivity of RP6530 for PI3Kδ and γ isoforms and subsequent downstream activity was determined in enzyme and cell-based assays. Additionally, RP6530 was tested for potency in viability, apoptosis, and Akt phosphorylation assays using a range of immortalized B-cell lymphoma cell lines (Raji, TOLEDO, KG-1, JEKO, OCI-LY-1, OCI-LY-10, MAVER, and REC-1). Viability was assessed using the colorimetric MTT reagent after incubation of cells for 72 h. Inhibition of pAKT was estimated by Western Blotting and bands were quantified using ImageJ after normalization with Actin. Primary cells from lymphoid tumors [1 chronic lymphocytic leukemia (CLL), 2 diffuse large B-cell lymphomas (DLBCL), 2 mantle cell lymphoma (MCL), 1 splenic marginal zone lymphoma (SMZL), and 1 extranodal MZL (EMZL)] were isolated, incubated with 4 µM RP6530, and analyzed for apoptosis or cytotoxicity by Annexin V/PI staining.
Results RP6530 demonstrated high potency against PI3Kδ (IC50=24.5 nM) and γ (IC50=33.2 nM) enzymes with selectivity over α (>300-fold) and β (>100-fold) isoforms. Cellular potency was confirmed in target-specific assays, namely anti-FcεR1-(EC50=37.8 nM) or fMLP (EC50=39.0 nM) induced CD63 expression in human whole blood basophils, LPS induced CD19+ cell proliferation in human whole blood (EC50=250 nM), and LPS induced CD45R+ cell proliferation in mouse whole blood (EC50=101 nM). RP6530 caused a dose-dependent inhibition (>50% @ 2-7 μM) in growth of immortalized (Raji, TOLEDO, KG-1, JEKO, REC-1) B-cell lymphoma cells. Effect was more pronounced in the DLBCL cell lines, OCI-LY-1 and OCI-LY-10 (>50% inhibition @ 0.1-0.7 μM), and the reduction in viability was accompanied by corresponding inhibition of pAKT with EC50 of 6 & 70 nM respectively. Treatment of patient-derived primary cells with 4 µM RP6530 caused an increase in cell death. Fold-increase in cytotoxicity as evident from PI+ staining was 1.6 for CLL, 1.1 for DLBCL, 1.2 for MCL, 2.2 for SMZL, and 2.3 for EMZL. Cells in early apotosis (Annexin V+/PI-) were not different between the DMSO blank and RP6530 samples.
Conclusions RP6530 is a potent and selective dual PI3Kδ/γ inhibitor that inhibited growth of B-cell lymphoma cell lines with a concomitant reduction in the downstream biomarker, pAKT. Additionally, the compound showed cytotoxicity in a panel of lymphoma primary cells. Findings provide a rationale for future clinical trials in B-cell malignancies.
While alpha and beta isoforms are ubiquitous in their distribution, expression of delta and gamma is restricted to circulating hematogenous cells and endothelial cells. Unlike PI3K-alpha or beta, mice lacking expression of gamma or delta do not show any adverse phenotype indicating that targeting of these specific isoforms would not result in overt toxicity. Dual delta/gamma inhibition is strongly implicated as an intervention strategy in allergic and non-allergic inflammation of the airways and other autoimmune diseases. Scientific evidence for PI3K-delta and gamma involvement in various cellular processes underlying asthma and COPD stems from inhibitor studies and gene-targeting approaches. Also, resistance to conventional therapies such as corticosteroids in several COPD patients has been attributed to an up-regulation of the PI3K delta/gamma pathway. Disruption of PI3K-delta/gamma signalling therefore provides a novel strategy aimed at counteracting the immuno-inflammatory response. Due to the pivotal role played by PI3K-delta and gamma in mediating inflammatory cell functionality such as leukocyte migration and activation, and mast cell degranulation, blocking these isoforms may also be an effective strategy for the treatment of rheumatoid arthritis as well.
Given the established criticality of these isoforms in immune surveillance, inhibitors specifically targeting the delta and gamma isoforms would be expected to attenuate the progression of immune response encountered in airway inflammation and rheumatoid arthritis.
Rhizen has identified an orally active Lead Molecule, RP-6530, that has an excellent pre-clinical profile. RP-6530 is currently in non-GLP Tox studies and is expected to enter Clinical Development in H2 2013.
In December 2013, Rhizen announced the start of a Phase I clinical trial. The study entitled A Phase-I, Dose Escalation Study to Evaluate Safety and Efficacy of RP6530, a dual PI3K delta /gamma inhibitor, in patients with Relapsed or Refractory Hematologic Malignancies is designed primarily to establish the safety and tolerability of RP6530. Secondary objectives include clinical efficacy assessment and biomarker response to allow dose determination and potential patient stratification in subsequent expansion studies.
Partners by Region
Rhizen’s pipeline consists of internally discovered (with 100% IP ownership) novel small molecule programs aimed at high value markets of Oncology, Immuno-inflammtion and Metabolic Disorders. Rhizen has been successful in securing critical IP space in these areas and efforts are on for further expansion in to several indications. Rhizen seeks partnerships to unlock the potential of these valuable assets for further development from global pharmaceutical partners. At present global rights on all programs are available and Rhizen is flexible to consider suitable business models for licensing/collaboration.
In 2012, Rhizen announced a joint venture collaboration with TG Therapeutics for global development and commercialization of Rhizen’s Novel Selective PI3K Kinase Inhibitors. The selected lead RP5264 (hereafter, to be developed as TGR-1202) is an orally available, small molecule, PI3K specific inhibitor currently being positioned for the treatment of hematological malignancies.
PATENT
WO2014195888, DUAL SELECTIVE PI3 DELTA AND GAMMA KINASE INHIBITORS
This scheme provides a synthetic route for the preparation of compound of formula wherein all the variables are as described herein in above
15 14 10 12 12a
REFERENCES
April 2015, preclinical data were presented at the 106th AACR Meeting in Philadelphia, PA. RP-6530 had GI50 values of 17,028 and 22,014 nM, respectively
December 2014, data were presented at the 56th ASH Meeting in San Francisco, CA.
December 2013, preclinical data were presented at the 55th ASH Meeting in New Orleans, LA.
June 2013, preclinical data were presented at the 18th Annual EHA Congress in Stockholm, Sweden. RP-6530 inhibited PI3K delta and gamma isoforms with IC50 values of 24.5 and 33.2 nM, respectively.
01 Sep 2015 Phase-I clinical trials in Hematological malignancies (Second-line therapy or greater) in USA (PO) (NCT02567656)
18 Nov 2014 Preclinical trials in Multiple myeloma in Switzerland (PO) prior to November 2014
18 Nov 2014 Early research in Multiple myeloma in Switzerland (PO) prior to November 2014
Process for preparation of optically pure and optionally substituted 2- (1 -hydroxy- alkyl) – chromen – 4 – one derivatives and their use in preparing pharmaceuticals
Biochemistry and Dysmetabolism of Aging and Serious Illness, Volume 2 (Volume Two: Latest in Genomics Methodologies for Therapeutics: Gene Editing, NGS and BioInformatics, Simulations and the Genome Ontology), Part 1: Next Generation Sequencing (NGS)
Biochemistry and Dysmetabolism of Aging and Serious Illness
Curator: Larry H. Bernstein, MD, FCAP
White Matter Lipids as a Ketogenic Fuel Supply in Aging Female Brain: Implications for Alzheimer’s Disease
•Mitochondrial dysfunction activates mechanisms for catabolism of myelin lipids to generate ketone bodies for ATP production.
•Mechanisms leading to ketone body driven energy production in brain coincide with stages of reproductive aging in females.
•Sequential activation of myelin catabolism pathway during aging provides multiple therapeutic targets and windows of efficacy.
The mechanisms underlying white matter degeneration, a hallmark of multiple neurodegenerative diseases including Alzheimer’s, remain unclear. Herein we provide a mechanistic pathway, spanning multiple transitions of aging, that links mitochondrial dysfunction early in aging with later age white matter degeneration. Catabolism of myelin lipids to generate ketone bodies can be viewed as an adaptive survival response to address brain fuel and energy demand. Women are at greatest risk of late-onset-AD, thus, our analyses in female brain address mechanisms of AD pathology and therapeutic targets to prevent, delay and treat AD in the sex most affected with potential relevance to men.
White matter degeneration is a pathological hallmark of neurodegenerative diseases including Alzheimer’s. Age remains the greatest risk factor for Alzheimer’s and the prevalence of age-related late onset Alzheimer’s is greatest in females. We investigated mechanisms underlying white matter degeneration in an animal model consistent with the sex at greatest Alzheimer’s risk. Results of these analyses demonstrated decline in mitochondrial respiration, increased mitochondrial hydrogen peroxide production and cytosolic-phospholipase-A2 sphingomyelinase pathway activation during female brain aging. Electron microscopic and lipidomic analyses confirmed myelin degeneration. An increase in fatty acids and mitochondrial fatty acid metabolism machinery was coincident with a rise in brain ketone bodies and decline in plasma ketone bodies. This mechanistic pathway and its chronologically phased activation, links mitochondrial dysfunction early in aging with later age development of white matter degeneration. The catabolism of myelin lipids to generate ketone bodies can be viewed as a systems level adaptive response to address brain fuel and energy demand. Elucidation of the initiating factors and the mechanistic pathway leading to white matter catabolism in the aging female brain provides potential therapeutic targets to prevent and treat demyelinating diseases such as Alzheimer’s and multiple sclerosis. Targeting stages of disease and associated mechanisms will be critical.
Age remains the greatest risk factor for developing AD (Hansson et al., 2006, Alzheimer’s, 2015). Thus, investigation of transitions in the aging brain is a reasoned strategy for elucidating mechanisms and pathways of vulnerability for developing AD. Aging, while typically perceived as a linear process, is likely composed of dynamic transition states, which can protect against or exacerbate vulnerability to AD (Brinton et al., 2015). An aging transition unique to the female is the perimenopausal to menopausal conversion (Brinton et al., 2015). The bioenergetic similarities between the menopausal transition in women and the early appearance of hypometabolism in persons at risk for AD make the aging female a rational model to investigate mechanisms underlying risk of late onset AD.
Findings from this study replicate our earlier findings that age of reproductive senescence is associated with decline in mitochondrial respiration, increased H2O2 production and shift to ketogenic metabolism in brain (Yao et al., 2010, Ding et al., 2013, Yin et al., 2015). These well established early age-related changes in mitochondrial function and shift to ketone body utilization in brain, are now linked to a mechanistic pathway that connects early decline in mitochondrial respiration and H2O2 production to activation of the cPLA2-sphingomyelinase pathway to catabolize myelin lipids resulting in WM degeneration (Fig. 12). These lipids are sequestered in lipid droplets for subsequent use as a local source of ketone body generation via astrocyte mediated beta-oxidation of fatty acids. Astrocyte derived ketone bodies can then be transported to neurons where they undergo ketolysis to generate acetyl-CoA for TCA derived ATP generation required for synaptic and cell function (Fig. 12).
Schematic model of mitochondrial H2O2 activation of cPLA2-sphingomyelinase pathway as an adaptive response to provide myelin derived fatty acids as a substrate for ketone body generation: The cPLA2-sphingomyelinase pathway is proposed as a mechanistic pathway that links an early event, mitochondrial dysfunction and H2O2, in the prodromal/preclinical phase of Alzheimer’s with later stage development of pathology, white matter degeneration. Our findings demonstrate that an age dependent deficit in mitochondrial respiration and a concomitant rise in oxidative stress activate an adaptive cPLA2-sphingomyelinase pathway to provide myelin derived fatty acids as a substrate for ketone body generation to fuel an energetically compromised brain.
Biochemical evidence obtained from isolated whole brain mitochondria confirms that during reproductive senescence and in response to estrogen deprivation brain mitochondria decline in respiratory capacity (Yao et al., 2009, Yao et al., 2010, Brinton, 2008a, Brinton, 2008b, Swerdlow and Khan, 2009). A well-documented consequence of mitochondrial dysfunction is increased production of reactive oxygen species (ROS), specifically H2O2 (Boveris and Chance, 1973, Beal, 2005, Yin et al., 2014, Yap et al., 2009). While most research focuses on the damage generated by free radicals, in this case H2O2 functions as a signaling molecule to activate cPLA2, the initiating enzyme in the cPLA2-sphingomyelinase pathway (Farooqui and Horrocks, 2006, Han et al., 2003, Sun et al., 2004). In AD brain, increased cPLA2 immunoreactivity is detected almost exclusively in astrocytes suggesting that activation of the cPLA2-sphingomyelinase pathway is localized to astrocytes in AD, as opposed to the neuronal or oligodendroglial localization that is observed during apoptosis (Sun et al., 2004, Malaplate-Armand et al., 2006, Di Paolo and Kim, 2011, Stephenson et al., 1996,Stephenson et al., 1999). In our analysis, cPLA2 (Sanchez-Mejia and Mucke, 2010) activation followed the age-dependent rise in H2O2 production and was sustained at an elevated level.
Direct and robust activation of astrocytic cPLA2 by physiologically relevant concentrations of H2O2 was confirmed in vitro. Astrocytic involvement in the cPLA2-sphingomyelinase pathway was also indicated by an increase in cPLA2 positive astrocyte reactivity in WM tracts of reproductively incompetent mice. These data are consistent with findings from brains of persons with AD that demonstrate the same striking localization of cPLA2immunoreactivity within astrocytes, specifically in the hippocampal formation (Farooqui and Horrocks, 2004). While neurons and astrocytes contain endogenous levels of cPLA2, neuronal cPLA2 is activated by an influx of intracellular calcium, whereas astrocytic cPLA2 is directly activated by excessive generation of H2O2 (Sun et al., 2004, Xu et al., 2003, Tournier et al., 1997). Evidence of this cell type specific activation was confirmed by the activation of cPLA2 in astrocytes by H2O2 and the lack of activation in neurons. These data support that astrocytic, not neuronal, cPLA2 is the cellular mediator of the H2O2 dependent cPLA2-sphingomyelinase pathway activation and provide associative evidence supporting a role of astrocytic mitochondrial H2O2 in age-related WM catabolism.
The pattern of gene expression during the shift to reproductive senescence in the female mouse hippocampus recapitulates key observations in human AD brain tissue, specifically elevation in cPLA2, sphingomyelinase and ceramidase (Schaeffer et al., 2010, He et al., 2010, Li et al., 2014). Further, up-regulation of myelin synthesis, lipid metabolism and inflammatory genes in reproductively incompetent female mice is consistent with the gene expression pattern previously reported from aged male rodent hippocampus, aged female non-human primate hippocampus and human AD hippocampus (Blalock et al., 2003, Blalock et al., 2004, Blalock et al., 2010, Blalock et al., 2011, Kadish et al., 2009, Rowe et al., 2007). In these analyses of gene expression in aged male rodent hippocampus, aged female non-human primate hippocampus and human AD hippocampus down regulation of genes related to mitochondrial function, and up-regulation in multiple genes encoding for enzymes involved in ketone body metabolism occurred (Blalock et al., 2003, Blalock et al., 2004, Blalock et al., 2010, Blalock et al., 2011, Kadish et al., 2009, Rowe et al., 2007). The comparability across data derived from aging female mouse hippocampus reported herein and those derived from male rodent brain, female nonhuman brain and human AD brain strongly suggest that cPLA2-sphingomyelinase pathway activation, myelin sheath degeneration and fatty acid metabolism leading to ketone body generation is a metabolic adaptation that is generalizable across these naturally aging models and are evident in aged human AD brain. Collectively, these data support the translational relevance of findings reported herein.
Data obtained via immunohistochemistry, electron microscopy and MBP protein analyses demonstrated an age-related loss in myelin sheath integrity. Evidence for a loss of myelin structural integrity emerged in reproductively incompetent mice following activation of the cPLA2-sphingomyelinase pathway. The unraveling myelin phenotype observed following reproductive senescence and aging reported herein is consistent with the degenerative phenotype that emerges following exposure to the chemotherapy drug bortezomib which induces mitochondrial dysfunction and increased ROS generation (Carozzi et al., 2010, Cavaletti et al., 2007,Ling et al., 2003). In parallel to the decline in myelin integrity, lipid droplet density increased. In aged mice, accumulation of lipid droplets declined in parallel to the rise in ketone bodies consistent with the utilization of myelin-derived fatty acids to generate ketone bodies. Due to the sequential relationship between WM degeneration and lipid droplet formation, we posit that lipid droplets serve as a temporary storage site for myelin-derived fatty acids prior to undergoing β-oxidation in astrocytes to generate ketone bodies.
Microstructural alterations in myelin integrity were associated with alterations in the lipid profile of brain, indicative of WM degeneration resulting in release of myelin lipids. Sphingomyelin and galactocerebroside are two main lipids that compose the myelin sheath (Baumann and Pham-Dinh, 2001). Ceramide is common to both galactocerebroside and sphingomyelin and is composed of sphingosine coupled to a fatty acid. Ceramide levels increase in aging, in states of ketosis and in neurodegeneration (Filippov et al., 2012, Blazquez et al., 1999, Costantini et al., 2005). Specifically, ceramide levels are elevated at the earliest clinically recognizable stage of AD, indicating a degree of WM degeneration early in disease progression (Di Paolo and Kim, 2011,Han et al., 2002, Costantini et al., 2005). Sphingosine is statistically significantly elevated in the brains of AD patients compared to healthy controls; a rise that was significantly correlated with acid sphingomyelinase activity, Aβ levels and tau hyperphosphorylation (He et al., 2010). In our analyses, a rise in ceramides was first observed early in the aging process in reproductively incompetent mice. The rise in ceramides was coincident with the emergence of loss of myelin integrity consistent with the release of myelin ceramides from sphingomyelin via sphingomyelinase activation. Following the rise in ceramides, sphingosine and fatty acid levels increased. The temporal sequence of the lipid profile was consistent with gene expression indicating activation of ceramidase for catabolism of ceramide into sphingosine and fatty acid during reproductive senescence. Once released from ceramide, fatty acids can be transported into the mitochondrial matrix of astrocytes via CPT-1, where β-oxidation of fatty acids leads to the generation of acetyl-CoA (Glatz et al., 2010). It is well documented that acetyl-CoA cannot cross the inner mitochondrial membrane, thus posing a barrier to direct transport of acetyl-CoA generated by β-oxidation into neurons. In response, the newly generated acetyl-CoA undergoes ketogenesis to generate ketone bodies to fuel energy demands of neurons (Morris, 2005,Guzman and Blazquez, 2004, Stacpoole, 2012). Because astrocytes serve as the primary location of β-oxidation in brain they are critical to maintaining neuronal metabolic viability during periods of reduced glucose utilization (Panov et al., 2014, Ebert et al., 2003, Guzman and Blazquez, 2004).
Once fatty acids are released from myelin ceramides, they are transported into astrocytic mitochondria by CPT1 to undergo β-oxidation. The mitochondrial trifunctional protein HADHA catalyzes the last three steps of mitochondrial β-oxidation of long chain fatty acids, while mitochondrial ABAD (aka SCHAD—short chain fatty acid dehydrogenase) metabolizes short chain fatty acids. Concurrent with the release of myelin fatty acids in aged female mice, CPT1, HADHA and ABAD protein expression as well as ketone body generation increased significantly. These findings indicate that astrocytes play a pivotal role in the response to bioenergetic crisis in brain to activate an adaptive compensatory system that activates catabolism of myelin lipids and the metabolism of those lipids into fatty acids to generate ketone bodies necessary to fuel neuronal demand for acetyl-CoA and ATP.
Collectively, these findings provide a mechanistic pathway that links mitochondrial dysfunction and H2O2generation in brain early in the aging process to later stage white matter degeneration. Astrocytes play a pivotal role in providing a mechanistic strategy to address the bioenergetic demand of neurons in the aging female brain. While this pathway is coincident with reproductive aging in the female brain, it is likely to have mechanistic translatability to the aging male brain. Further, the mechanistic link between bioenergetic decline and WM degeneration has potential relevance to other neurological diseases involving white matter in which postmenopausal women are at greater risk, such as multiple sclerosis. The mechanistic pathway reported herein spans time and is characterized by a progression of early adaptive changes in the bioenergetic system of the brain leading to WM degeneration and ketone body production. Translationally, effective therapeutics to prevent, delay and treat WM degeneration during aging and Alzheimer’s disease will need to specifically target stages within the mechanistic pathway described herein. The fundamental initiating event is a bioenergetic switch from being a glucose dependent brain to a glucose and ketone body dependent brain. It remains to be determined whether it is possible to prevent conversion to or reversal of a ketone dependent brain. Effective therapeutic strategies to intervene in this process require biomarkers of bioenergetic phenotype of the brain and stage of mechanistic progression. The mechanistic pathway reported herein may have relevance to other age-related neurodegenerative diseases characterized by white matter degeneration such as multiple sclerosis.
Cancer cells are characterized by perturbations of their metabolic processes. Recent observations demonstrated that the fatty acid oxidation (FAO) pathway may represent an alternative carbon source for anabolic processes in different tumors, therefore appearing particularly promising for therapeutic purposes. Because the carnitine palmitoyl transferase 1a (CPT1a) is a protein that catalyzes the rate-limiting step of FAO, here we investigated the in vitro antileukemic activity of the novel CPT1a inhibitor ST1326 on leukemia cell lines and primary cells obtained from patients with hematologic malignancies. By real-time metabolic analysis, we documented that ST1326 inhibited FAO in leukemia cell lines associated with a dose- and time-dependent cell growth arrest, mitochondrial damage, and apoptosis induction. Data obtained on primary hematopoietic malignant cells confirmed the FAO inhibition and cytotoxic activity of ST1326, particularly on acute myeloid leukemia cells. These data suggest that leukemia treatment may be carried out by targeting metabolic processes.
Suppression of detachment-induced cell death, known as anoikis, is an essential step for cancer metastasis to occur. We report here that expression of KLF12, a member of the Kruppel-like family of transcription factors, is downregulated in lung cancer cell lines that have been selected to grow in the absence of cell adhesion. Knockdown of KLF12 in parental cells results in decreased apoptosis following cell detachment from matrix. KLF12 regulates anoikis by promoting the cell cycle transition through S phase and therefore cell proliferation. Reduced expression levels of KLF12 results in increased ability of lung cancer cells to form tumours in vivo and is associated with poorer survival in lung cancer patients. We therefore identify KLF12 as a novel metastasis-suppressor gene whose loss of function is associated with anoikis resistance through control of the cell cycle.
Mol Cell. 2015 Oct 14. pii: S1097-2765(15)00764-9. doi: 10.1016/j.molcel.2015.09.025. [Epub ahead of print] PEPCK Coordinates the Regulation of Central Carbon Metabolism to Promote Cancer Cell Growth.
Phosphoenolpyruvate carboxykinase (PEPCK) is well known for its role in gluconeogenesis. However, PEPCK is also a key regulator of TCA cycle flux. The TCA cycle integrates glucose, amino acid, and lipid metabolism depending on cellular needs. In addition, biosynthetic pathways crucial to tumor growth require the TCA cycle for the processing of glucose and glutamine derived carbons. We show here an unexpected role for PEPCK in promoting cancer cell proliferation in vitro and in vivo by increasing glucose and glutamine utilization toward anabolic metabolism. Unexpectedly, PEPCK also increased the synthesis of ribose from non-carbohydrate sources, such as glutamine, a phenomenon not previously described. Finally, we show that the effects of PEPCK on glucose metabolism and cell proliferation are in part mediated via activation of mTORC1. Taken together, these data demonstrate a role for PEPCK that links metabolic flux and anabolic pathways to cancer cell proliferation.
BACKGROUND: Mutations in the Fe-S cluster-containing SDHB subunit of succinate dehydrogenase cause familial cancer syndromes. Recently the tripeptide motif L(I)YR was identified in the Fe-S recipient protein SDHB, to which the cochaperone HSC20 binds.
METHODS: In order to characterize the metabolic basis of SDH-deficient cancers we performed stable isotope-resolved metabolomics in a novel SDHB-deficient renal cell carcinoma cell line and conducted bioinformatics and biochemical screening to analyze Fe-S cluster acquisition and assembly of SDH in the presence of other cancer-causing SDHB mutations.
RESULTS:
We found that the SDHB(R46Q) mutation in UOK269 cells disrupted binding of HSC20, causing rapid degradation of SDHB. In the absence of SDHB, respiration was undetectable in UOK269 cells, succinate was elevated to 351.4±63.2 nmol/mg cellular protein, and glutamine became the main source of TCA cycle metabolites through reductive carboxylation. Furthermore, HIF1α, but not HIF2α, increased markedly and the cells showed a strong DNA CpG island methylator phenotype (CIMP). Biochemical and bioinformatic screening revealed that 37% of disease-causing missense mutations in SDHB were located in either the L(I)YR Fe-S transfer motifs or in the 11 Fe-S cluster-ligating cysteines.
CONCLUSIONS:
These findings provide a conceptual framework for understanding how particular mutations disproportionately cause the loss of SDH activity, resulting in accumulation of succinate and metabolic remodeling in SDHB cancer syndromes.
Targeting cancer cell metabolism is a promising strategy against cancer. Here, we confirmed that the anti-cancer drug carboxyamidotriazole (CAI) inhibited mitochondrial respiration in cancer cells for the first time and found a way to enhance its anti-cancer activity by further disturbing the energy metabolism. CAI promoted glucose uptake and lactate production when incubated with cancer cells. The oxidative phosphorylation (OXPHOS) in cancer cells was inhibited by CAI, and the decrease in the activity of the respiratory chain complex I could be one explanation. The anti-cancer effect of CAI was greatly potentiated when being combined with 2-deoxyglucose (2-DG). The cancer cells treated with the combination of CAI and 2-DG were arrested in G2/M phase. The apoptosis and necrosis rates were also increased. In a mouse xenograft model, this combination was well tolerated and retarded the tumor growth. The impairment of cancer cell survival was associated with significant cellular ATP decrease, suggesting that the combination of CAI and 2-DG could be one of the strategies to cause dual inhibition of energy pathways, which might be an effective therapeutic approach for a broad spectrum of tumors.
Yun, E. Mullarky, C. Lu, K. N. Bosch, A. Kavalier, K. Rivera, J. Roper, Chio, II, E. G. Giannopoulou, C. Rago, A. Muley, J. M. Asara, J. Paik, O. Elemento, Z. Chen, D. J. Pappin, L. E. Dow, N. Papadopoulos, S. S. Gross and L. C. Cantley.
The G protein-coupled receptor kinase-2 (GRK2) is upregulated in the injured heart and contributes to heart failure pathogenesis. GRK2 was recently shown to associate with mitochondria but its functional impact in myocytes due to this localization is unclear. This study was undertaken to determine the effect of elevated GRK2 on mitochondrial respiration in cardiomyocytes. Sub-fractionation of purified cardiac mitochondria revealed that basally GRK2 is found in multiple compartments. Overexpression of GRK2 in mouse cardiomyocytes resulted in an increased amount of mitochondrial-based superoxide. Inhibition of GRK2 increased oxygen consumption rates and ATP production. Moreover, fatty acid oxidation was found to be significantly impaired when GRK2 was elevated and was dependent on the catalytic activity and mitochondrial localization of this kinase. Our study shows that independent of cardiac injury, GRK2 is localized in the mitochondria and its kinase activity negatively impacts the function of this organelle by increasing superoxide levels and altering substrate utilization for energy production.
Br J Pharmacol. 2015 Oct 27. doi: 10.1111/bph.13377. [Epub ahead of print]
All-trans retinoic acid protects against doxorubicin-induced cardiotoxicity by activating the Erk2 signalling pathway.
Doxorubicin (Dox) is a powerful antineoplastic agent for treating a wide range of cancers. However, doxorubicin cardiotoxicity of the heart has largely limited its clinical use. It is known that all-trans retinoic acid (ATRA) plays important roles in many cardiac biological processes, however, the protective effects of ATRA on doxorubicin cardiotoxicity remain unknown. Here, we studied the effect of ATRA on doxorubicin cardiotoxicity and underlying mechanisms.
EXPERIMENTAL APPROACHES:
Cellular viability assays, western blotting and mitochondrial respiration analyses were employed to evaluate the cellular response to ATRA in H9c2 cells and primary cardiomyocytes. Quantitative PCR (Polymerase Chain Reaction) and gene knockdown were performed to investigate the underlying molecular mechanisms of ATRA’s effects on doxorubicin cardiotoxicity.
KEY RESULTS:
ATRA significantly inhibited doxorubicin-induced apoptosis in H9c2 cells and primary cardiomyocytes. ATRA was more effective against doxorubicin cardiotoxicity than resveratrol and dexrazoxane. ATRA also suppressed reactive oxygen species (ROS) generation, and restored the expression level of mRNA and proteins in phase II detoxifying enzyme system: Nrf2 (nuclear factor-E2-related factor 2), MnSOD (manganese superoxide dismutase), HO-1 (heme oxygenase1) as well as mitochondrial function (mitochondrial membrane integrity, mitochondrial DNA copy numbers, mitochondrial respiration capacity, biogenesis and dynamics). Both Erk1/2 (extracellular signal-regulated kinase1/2) inhibitor (U0126) and Erk2 siRNA, but not Erk1 siRNA, abolished the protective effect of ATRA against doxorubicin-induced toxicity in H9c2 cells. Remarkably, ATRA did not compromise the anticancer efficacy of doxorubicin in gastric carcinoma cells.
CONCLUSION AND IMPLICATION:
ATRA protected cardiomyocytes against doxorubicin-induced toxicity by activating the Erk2 pathway without compromising the anticancer efficacy of doxorubicin. Therefore, ATRA may be a promising candidate as a cardioprotective agent against doxorubicin cardiotoxicity.
Colak, O. Pougovkina, L. Dai, M. Tan, H. Te Brinke, H. Huang, Z. Cheng, J. Park, X. Wan, X. Liu, W. W. Yue, R. J. Wanders, J. W. Locasale, D. B. Lombard, V. C. de Boer and Y. Zhao.
Mol Cell Proteomics. 2015 Nov 1, 14 (11): 3056-71.
The insulin/insulin-like growth factor (IGF)-1 signaling pathway (ISP) plays a fundamental role in long term health in a range of organisms. Protein kinases including Akt and ERK are intimately involved in the ISP. To identify other kinases that may participate in this pathway or intersect with it in a regulatory manner, we performed a whole kinome (779 kinases) siRNA screen for positive or negative regulators of the ISP, using GLUT4 translocation to the cell surface as an output for pathway activity. We identified PFKFB3, a positive regulator of glycolysis that is highly expressed in cancer cells and adipocytes, as a positive ISP regulator. Pharmacological inhibition of PFKFB3 suppressed insulin-stimulated glucose uptake, GLUT4 translocation, and Akt signaling in 3T3-L1 adipocytes. In contrast, overexpression of PFKFB3 in HEK293 cells potentiated insulin-dependent phosphorylation of Akt and Akt substrates. Furthermore, pharmacological modulation of glycolysis in 3T3-L1 adipocytes affected Akt phosphorylation. These data add to an emerging body of evidence that metabolism plays a central role in regulating numerous biological processes including the ISP. Our findings have important implications for diseases such as type 2 diabetes and cancer that are characterized by marked disruption of both metabolism and growth factor signaling.
Skeletal muscle mitochondrial content and oxidative capacity are important determinants of muscle function and whole-body health. Mitochondrial content and function are enhanced by endurance exercise and impaired in states or diseases where muscle function is compromised, such as myopathies, muscular dystrophies, neuromuscular diseases, and age-related muscle atrophy. Hence, elucidating the mechanisms that control muscle mitochondrial content and oxidative function can provide new insights into states and diseases that affect muscle health. In past studies, we identified Perm1 (PPARGC1- and ESRR-induced regulator, muscle 1) as a gene induced by endurance exercise in skeletal muscle, and regulating mitochondrial oxidative function in cultured myotubes. The capacity of Perm1 to regulate muscle mitochondrial content and function in vivo is not yet known. In this study, we use adeno-associated viral (AAV) vectors to increase Perm1 expression in skeletal muscles of 4-wk-old mice. Compared to control vector, AAV1-Perm1 leads to significant increases in mitochondrial content and oxidative capacity (by 40-80%). Moreover, AAV1-Perm1-transduced muscles show increased capillary density and resistance to fatigue (by 33 and 31%, respectively), without prominent changes in fiber-type composition. These findings suggest that Perm1 selectively regulates mitochondrial biogenesis and oxidative function, and implicate Perm1 in muscle adaptations that also occur in response to endurance exercise.-Cho, Y., Hazen, B. C., Gandra, P. G., Ward, S. R., Schenk, S., Russell, A. P., Kralli, A. Perm1 enhances mitochondrial biogenesis, oxidative capacity, and fatigue resistance in adult skeletal muscle.
Exposure to metabolic disease during fetal development alters cellular differentiation and perturbs metabolic homeostasis, but the underlying molecular regulators of this phenomenon in muscle cells are not completely understood. To address this, we undertook a computational approach to identify cooperating partners of the myocyte enhancer factor-2 (MEF2) family of transcription factors, known regulators of muscle differentiation and metabolic function. We demonstrate that MEF2 and the serum response factor (SRF) collaboratively regulate the expression of numerous muscle-specific genes, including microRNA-133a (miR-133a). Using tandem mass spectrometry techniques, we identify a conserved phosphorylation motif within the MEF2 and SRF Mcm1 Agamous Deficiens SRF (MADS)-box that regulates miR-133a expression and mitochondrial function in response to a lipotoxic signal. Furthermore, reconstitution of MEF2 function by expression of a neutralizing mutation in this identified phosphorylation motif restores miR-133a expression and mitochondrial membrane potential during lipotoxicity. Mechanistically, we demonstrate that miR-133a regulates mitochondrial function through translational inhibition of a mitophagy and cell death modulating protein, called Nix. Finally, we show that rodents exposed to gestational diabetes during fetal development display muscle diacylglycerol accumulation, concurrent with insulin resistance, reduced miR-133a, and elevated Nix expression, as young adult rats. Given the diverse roles of miR-133a and Nix in regulating mitochondrial function, and proliferation in certain cancers, dysregulation of this genetic pathway may have broad implications involving insulin resistance, cardiovascular disease, and cancer biology.
Advinus Therapeutics announced that it has successfully completed a 14-day POC study in 60 Type II diabetic patients on its lead molecule, GKM-001, a glucokinase activator. The results of the trial show effective glucose lowering across all doses tested without any incidence of hypoglycemia or any other clinically relevant adverse events.
GKM-001 is differentiated from most other GK molecules that are in development, or have been discontinued, due to its novel liver selective mechanism of action.
GKM-001 belongs to a novel class of molecules for treatment of type II diabetes. It is an activator of Glucokinase (GK), a glucose-sensing enzyme found mainly in the liver and pancreas. Being liver selective, GKM-001 mostly activates GK in the liver and not in pancreas, which is its key differentiation from most competitor molecules that activate GK in pancreas as well.
A glucokinase activator for treatment of type II diabetes, currently in PI. Advinus is actively exploring partnership options to expedite further development and WW marketing of GKM-001.
Examples C2 and C3 were prepared in analogues manner of example (Cl) from the appropriate chiral intermediate:
Example Dl : (+)-{5-Chloro-2-[2-(4-cyclopropanesulfonylphenyl)-2-(2,4- difluorophenoxy)acetylamino]thiazol-4-yl}acetic acid, ethyl ester
Advinus’ GK-activator Achieves Early POC for Diabetes
November 29 2011
Partnership Dialog Actively Underway
Advinus Therapeutics, a research-based pharmaceutical company founded by globally experienced industry executives and promoted by the TATA Group, announced that it has successfully completed a 14-day POC study in 60 Type II diabetic patients on its lead molecule, GKM-001, a glucokinase activator. The results of the trial show effective glucose lowering across all doses tested without any incidence of hypoglycemia or any other clinically relevant adverse events.
The clinical trials on GKM-001 validate the company’s pre-clinical hypothesis that a liver selective Glucokinase activator would not cause hypoglycemia (very low blood sugar), while showing robust efficacy.
“GKM-001 is differentiated from most other GK molecules that are in development, or have been discontinued, due to its novel liver selective mechanism of action. GKM-001 has a prolonged pharmacological effect and a half-life that should support a once a day dosing as both mono and combination therapy.” said Dr. Rashmi Barbhaiya, MD & CEO, Advinus Therapeutics. He added that Advinus is actively exploring partnership options to expedite further development and global marketing of GKM-001.
GKM-001 belongs to a novel class of molecules for treatment of type II diabetes. It is an activator of Glucokinase (GK), a glucose-sensing enzyme found mainly in the liver and pancreas. Being liver selective, GKM-001 mostly activates GK in the liver and not in pancreas, which is its key differentiation from most competitor molecules that activate GK in pancreas as well. The resulting increase in insulin secretion creates a potential for hypoglycemia-a risk GKM-001 is designed to avoid. Advinus has the composition of matter patent on GKM-001 for all major markets globally. Both the Single Ascending Dose data, in healthy and type II diabetics, and the Multiple Ascending Dose Study in Type II diabetics has shown that the molecule shows effective glucose lowering in a dose dependent manner and has excellent safety and tolerability profile over a 40-fold dose range. The pharmacokinetic properties of the molecule support once a day dosing. GKM-001 has the potential to be “First-in-Class” drug to address this large, growing and yet poorly addressed market.
Advinus also has identified a clinical candidate as a back-up to GKM-001, which is structurally different. In its portfolio, the company has a growing pipeline for COPD, sickle cell disease, inflammatory bowel disease, type 2 diabetes, acute and chronic pain and rheumatoid arthritis in various stages of late discovery and pre-clinical development.
The first glucokinase modulator discovered and developed in India
A new concept for the management of diabetes for patients, globally
100 per cent ‘made in India’ molecule for the treatment of diabetes
IND approved by DGCI, Phase I clinical trial shows excellent safety and tolerance profiles with efficacy
Bangalore: Advinus Therapeutics (Advinus), the research-based pharmaceutical company founded by leading global pharmaceutical executives and promoted by the Tata group, today, announced the discovery of a novel molecule for the treatment of type II diabetes — GKM-001.The molecule is an activator of glucokinase; an enzyme that regulates glucose balance and insulin secretion in the body.
GKM-001 is a completely indigenously developed molecule and the initial clinical trials have shown excellent results for both safety and efficacy.
“Considering past failures of other companies on this target, our discovery programme primarily focused on identifying a molecule that would be efficacious without causing hypoglycaemia; a side effect associated with most compounds developed for this target.
“Recently completed Phase I data indicate that Advinus’ GKM–001 is a liver selective molecule that has overcome the biggest clinical challenge of hypoglycaemia. GKM-001 is differentiated from most other GK molecules in development due to this novel mechanism of action,” said Dr Rashmi Barbhaiya, MD and CEO, Advinus Therapeutics.
He further added, “We are very proud that GKM-001 is 100 per cent Indian. Advinus’s discovery team in Pune discovered the molecule and entire preclinical development was carried out at our centre in Bangalore. The Investigational New Drug (IND) application was filed with the DGCI for approval to initiate clinical trials in India within 34 months of initiation of the discovery programme. Subsequent to the approval of the IND, we have completed the Phase I Single Ascending Dose study in India within two months.”
GKM-001 is a novel molecule for the treatment of type II diabetes. It is the first glucokinase modulator discovered and developed in India and has potential to be both first or best in class. The success in discovering GKM-001 is attributed to the science-driven efforts in Advinus laboratories and ‘breaking the conventional mold’ for selection of a drug candidate. Advinus has ‘composition of matter’ patent on the molecule for all major markets globally. Glucokinase as a class of target is considered to be novel as currently there is no product in the market or in late clinical trials. The strategy for early clinical development revolved around assessing safety (particularly hypoglycaemia) and early assessment of therapeutic activity (glucose lowering and other biomarkers) in type II diabetics. The Phase I data, in both healthy and type II diabetics, shows excellent safety and tolerability over a 40-fold dose range and desirable pharmacokinetic properties consistent with ‘once a day’ dosing. The next wave of clinical studies planned continues on this strategy of early testing in type II diabetics.
Right behind the lead candidate GKM-001, Advinus has a rich pipeline of back up compounds on the same target. These include several structurally different compounds with diverse potency, unique pharmacology and tissue selectivity. Having discovered the molecule with early indication of wide safety margins, desired efficacy and pharmacokinetic profiles, the company now seeks to out-licence GKM-001 and its discovery portfolio.
This e-Book is a comprehensive review of recent Original Research on METABOLOMICS and related opportunities for Targeted Therapy written by Experts, Authors, Writers. This is the first volume of the Series D: e-Books on BioMedicine – Metabolomics, Immunology, Infectious Diseases. It is written for comprehension at the third year medical student level, or as a reference for licensing board exams, but it is also written for the education of a first time baccalaureate degree reader in the biological sciences. Hopefully, it can be read with great interest by the undergraduate student who is undecided in the choice of a career. The results of Original Research are gaining value added for the e-Reader by the Methodology of Curation.The e-Book’s articles have been published on the Open Access Online Scientific Journal, since April 2012. All new articles on this subject, will continue to be incorporated, as published with periodical updates.
We invite e-Readers to write an Article Reviews on Amazon for this e-Book on Amazon.
All forthcoming BioMed e-Book Titles can be viewed at:
Leaders in Pharmaceutical Business Intelligence, launched in April 2012 an Open Access Online Scientific Journal is a scientific, medical and business multi expert authoring environment in several domains of life sciences, pharmaceutical, healthcare & medicine industries. The venture operates as an online scientific intellectual exchange at their website http://pharmaceuticalintelligence.com and for curation and reporting on frontiers in biomedical, biological sciences, healthcare economics, pharmacology, pharmaceuticals & medicine. In addition the venture publishes a Medical E-book Series available on Amazon’s Kindle platform.
Analyzing and sharing the vast and rapidly expanding volume of scientific knowledge has never been so crucial to innovation in the medical field. WE are addressing need of overcoming this scientific information overload by:
delivering curation and summary interpretations of latest findings and innovations on an open-access, Web 2.0 platform with future goals of providing primarily concept-driven search in the near future
providing a social platform for scientists and clinicians to enter into discussion using social media
compiling recent discoveries and issues in yearly-updated Medical E-book Series on Amazon’s mobile Kindle platform
This curation offers better organization and visibility to the critical information useful for the next innovations in academic, clinical, and industrial research by providing these hybrid networks.
Table of Contents forMetabolic Genomics & Pharmaceutics, Vol. I
Chapter 1: Metabolic Pathways
Chapter 2: Lipid Metabolism
Chapter 3: Cell Signaling
Chapter 4: Protein Synthesis and Degradation
Chapter 5: Sub-cellular Structure
Chapter 6: Proteomics
Chapter 7: Metabolomics
Chapter 8: Impairments in Pathological States: Endocrine Disorders; Stress
Hypermetabolism and Cancer
Chapter 9: Genomic Expression in Health and Disease
Motesanib, also known as AMG-706, is an orally administered multikinase inhibitor that selectively targets VEGF receptors, platelet-derived growth factor receptors, and Kit receptors.
When it has reached 1 to 2mm across, a tumor needs to grow its own blood vessels in order to continue to get bigger. Some cancer cells make a protein called vascular endothelial growth factor (VEGF). The VEGF protein attaches to receptors on cells that line the walls of blood vessels within the tumour.
Drugs that block blood vessel growth factor
Some drugs block vascular endothelial growth factor (VEGF) from attaching to the receptors on the cells that line the blood vessels. This stops the blood vessels from growing.
A drug that blocks VEGF is bevacizumab (Avastin). It is also a monoclonal antibody.
Sunitinib (Sutent) is a type of TKI that blocks the growth signals inside blood vessel cells. It is used to treat kidney cancer and a rare type of stomach cancer called gastrointestinal stromal tumour (GIST).
Drugs that affect signals between cells
Some drugs act on the chemicals that cells use to signal to each other to grow. This can block the formation of blood vessels. Drugs that works in this way include thalidomide and lenalidomide (Revlimid).
Each drug has different side effects. You can look up the name of your drug in our cancer drug section to find out about the side effects you may have.
To find trials using anti angiogenesis treatment go to our clinical trials database and type ‘angiogenesis’ into the search box.
Tumors can cause their blood supply to form by giving off chemical signals that stimulate angiogenesis. Tumors can also stimulate nearby normal cells to produce angiogenesis signaling molecules. The resulting new blood vessels “feed” growing tumors with oxygen and nutrients, allowing the cancer cells to invade nearby tissue, to move throughout the body, and to form colonies of cancer cells, called metastases. Because tumors cannot grow beyond a certain size or spread without a blood supply, scientists are trying to find ways to block tumor angiogenesis.
Angiogenesis requires the binding of signaling molecules, such as vascular endothelial growth factor (VEGF), to receptors on the surface of normal endothelial cells. When VEGF and other endothelial growth factors bind to their receptors on endothelial cells, signals within these cells are initiated that promote the growth and survival of new blood vessels.
Angiogenesis inhibitors interfere with various steps in this process. For example, bevacizumab (Avastin®) is a monoclonal antibody that specifically recognizes and binds to VEGF (1). When VEGF is attached to bevacizumab, it is unable to activate the VEGF receptor. Other angiogenesis inhibitors, including sorafenib and sunitinib, bind to receptors on the surface of endothelial cells or to other proteins in the downstream signaling pathways, blocking their activities (2).
The U.S. Food and Drug Administration (FDA) has approved bevacizumab to be used alone forglioblastoma that has not improved with other treatments and to be used in combination with other drugs to treat metastatic colorectal cancer, some non-small cell lung cancers, and metastatic renal cell cancer. Bevacizumab was the first angiogenesis inhibitor that was shown to slow tumor growth and, more important, to extend the lives of patients with some cancers.
The FDA has approved other drugs that have antiangiogenic activity, including sorafenib (Nexavar®), sunitinib(Sutent®), pazopanib (Votrient®), and everolimus (Afinitor®). Sorafenib is approved for hepatocellular carcinoma and kidney cancer, sunitinib and everolimus for both kidney cancer and neuroendocrine tumors, and pazopanib for kidney cancer.
Angiogenesis inhibitors are unique cancer-fighting agents because they tend to inhibit the growth of blood vessels rather than tumor cells. In some cancers, angiogenesis inhibitors are most effective when combined with additional therapies, especially chemotherapy. It has been hypothesized that these drugs help normalize the blood vessels that supply the tumor, facilitating the delivery of other anticancer agents, but this possibility is still being investigated.
Angiogenesis inhibitor therapy does not necessarily kill tumors but instead may prevent tumors from growing. Therefore, this type of therapy may need to be administered over a long period.
Initially, it was thought that angiogenesis inhibitors would have mild side effects, but more recent studies have revealed the potential for complications that reflect the importance of angiogenesis in many normal body processes, such as wound healing, heart and kidney function, fetal development, and reproduction. Side effects of treatment with angiogenesis inhibitors can include problems with bleeding, clots in the arteries (with resultant stroke or heart attack), hypertension, and protein in the urine (3–5). Gastrointestinal perforation and fistulas also appear to be rare side effects of some angiogenesis inhibitors.
In addition to the angiogenesis inhibitors that have already been approved by the FDA, others that target VEGF or other angiogenesis pathways are currently being tested in clinical trials (research studies involving patients). If these angiogenesis inhibitors prove to be both safe and effective in treating human cancer, they may be approved by the FDA and made available for widespread use.
In addition, phase I and II clinical trials are testing the possibility of combining angiogenesis inhibitor therapy with other treatments that target blood vessels, such as tumor-vascular disrupting agents, which damage existing tumor blood vessels (6).
9.5.3 Recent Advances in Anti-Angiogenic Therapy of Cancer
More than forty anti-angiogenic drugs are being tested in clinical trials all over the world. This review discusses agents that have approved by the FDA and are currently in use for treating patients either as single-agents or in combination with other chemotherapeutic agents.
Tumor angiogenesis is generation of a network of blood vessels within the cancerous growth. This process can occur two ways: The more accepted model involves the release of signaling molecules by the tumor cells; these molecules activate the surrounding tissue to promote growth of new blood vessels. This stimulates vascular endothelial cells to divide rapidly [9, 10]. The other model proposes the generation of new vasculature by vasculogenic mimicry. This model argues that the tumor cells trans-differentiate in endothelial-like cells and create structures from inside of the tumor tapping into a nearby blood vessel [4].
Escape of the tumor cell from the confines of the primary tumor to distant body parts is the pre-requisite for hematogenous metastasis. This escape route is provided by the tumor vasculature. Thus, it was envisioned that inhibition of angiogenesis will also lead to inhibition of metastasis. This phenomenon was demonstrated by very elegant mouse model studies using angiostatin [11, 12]. Angiostatin was also demonstrated to be secreted by some primary tumors leading to restricted growth of the metastasis leading to “dormancy” of the metastasis. Mice deficient in angiogenesis (Id1 & Id3 deficient) showed significantly less tumor take rates [13]. Independent studies showed absence of metastasis in angiogenesis deficient mice [14, 15]. Defective angiogenesis was attributed to impaired VEGF-dependent recruitment of precursor endothelial cells from the bone marrow to the newly developing tumor vasculature [16].
Metastasis of malignant tumors to regional lymph nodes is one of the early signs of cancer spread in patients, and it occurs at least as frequently as hematogenous metastasis [17]. Particularly, in cancers, such as breast cancer, lymphatic metastasis is a predominant route for tumor spread. The contribution of lymphatic system to the tumor growth is an area that is relatively less studied. However, lymphatic vessels are speculated to contribute to tumor growth and metastasis in a variety of ways. The VEGF, FGF2 and PDGF produced by vascular endothelial cells are proposed to be involved in the activation of lymphatic endothelial cells, which in turn produce matrix metalloproteases and urokinase plasminogen activator (uPA) that can promote malignant tumor growth. Thus, there exists a synergistic crosstalk between the tumor and the lymphatic vessels and blood vessels.
Angiogenesis is a complex and intricately regulated process. Like all other regulated biological phenomena, angiogenesis has activators or pro-angiogenic factors and inhibitors or anti-angiogenic factors [9].
The Activators
Tumor cells activate signaling pathways that promote uncontrolled proliferation and survival. These include the PI3K/AKT/mTOR pathway, Hedgehog pathway and, Wnt pathway [18–24] that produce pro-angiogenic signaling intermediates [25, 26]. Among the several reported activators of angiogenesis present in cells two proteins appear to be the most important for sustaining tumor growth: vascular endothelial growth factor (VEGF) and basic fibroblast growth factor (bFGF). VEGF and bFGF are secreted by the tumor into the surrounding tissue. They bind to their cognate receptors on endothelial cells. This activates a signaling cascade that transmits a nuclear signal prompting target genes to activate endothelial cell growth. Activated endothelial cells also produce matrix metalloproteinases (MMPs). These MMPs break down the extracellular matrix and allow the migration of endothelial cells. The division and migration of the endothelial cells leads to formation of new blood vessels [27, 28].
The Inhibitors
If angiogenesis is so critical for the tumor growth, then agents that inhibit angiogenesis would have great therapeutic value. With the discovery of endostatin, the concept of anti-angiogenic therapy was launched and popularized by Dr. Folkman [29]. Angiogenesis inhibitors have been discovered from a variety of sources. Some are naturally present in the human body e.g. specific fragments of structural proteins such as collagen or plasminogen (angiostatin, endostatin, tumstatin) [30]. Others are natural products in green tea, soy beans, fungi, mushrooms, tree bark, shark tissues, snake venom etc. [31]. A plethora of synthetic compounds are also characterized to have anti-angiogenic properties [32].
ANTI-ANGIOGENIC TREATMENT OF CANCER
Since angiogenesis is an event critical to primary tumor growth as well as metastasis, anti-angiogenic treatment of tumors is a highly promising therapeutic avenue [33]. Thus, for over last couple of decades, there has been a robust activity aimed towards the discovery of angiogenesis inhibitors [34, 35]. More than forty anti-angiogenic drugs are being tested in human cancer patients in clinical trials all over the world. From the several anti-angiogenic agents reported, we have focused this review on discussing those agents that have received FDA approval in the United States and are currently in use for treating patients either as a single-agent or in combination with other chemotherapeutic agents (Figure (Figure1).1). Based on functionality, the anti-angiogenic drugs can be sub-divided into three main groups:
Targets of FDA-approved angiogenesis inhibitors: Angiogenesis inhibitors impact both, the tumor as well as the endothelial cells resulting in the disruption of the effects of the microenvironment in promoting tumor growth and angiogenesis
Drugs that inhibit growth of endothelial cells
e.g. Endostatin and combretastatin A4, cause apoptosis of the endothelial cells [36]. Thalidomide is also a potent inhibitor of endothelial cell growth [37].
Drugs that block angiogenesis signaling
e.g. anti-VEGF antibodies (Avastin, FDA approved for colorectal cancer), Interferon-alpha (inhibits the production of bFGF and VEGF) [36].
ANTI-ANGIOGENIC THERAPIES THAT HAVE RECEIVED USA-FDA APPROVAL
Conventional chemotherapy is usually a systemic therapy that tries to capture a narrow therapeutic window offered by rapid proliferation of tumor cells compared to the normal cells. Chemotherapy has significant side effects such as hair loss, diarrhea, mouth ulcer, infection, and low blood counts. Anti-angiogenic therapy has several advantages over chemotherapy as it is mostly not directed towards directly killing cells but stopping the blood vessel formation, an event that is rare in tissues other than growing tumor. Hence it is well tolerated by the patients and has fewer side effects [29]. There are currently seven approved anti-cancer therapies in two primary categories:
Monoclonal antibodies directed against specific pro-angiogenic growth factors and/or their receptors
Small molecule tyrosine kinase inhibitors (TKIs) of multiple pro-angiogenic growth factor receptors.
Besides these, inhibitors of mTOR (mammalian target of rapamycin), proteasome inhibitors and thalidomide have also been reported to indirectly inhibit angiogenesis through mechanisms that are not completely understood.
MONOCLONAL ANTIBODY THERAPIES
Four monoclonal antibody therapies are approved to treat several tumor types:
Bevacizumab (Avastin®)
The first FDA approved angiogenesis inhibitor, Avastin is a humanized monoclonal antibody that binds biologically active forms of vascular endothelial growth factor (VEGF) and prevents its interaction with VEGF receptors (VEGFR-1 and VEGFR-2), thereby inhibiting endothelial cell proliferation and angiogenesis. Bevacizumab has been tested in phase I studies in combination with chemotherapy with a good safety profile [39]. This treatment is approved for metastatic colorectal cancer and non-small cell lung cancer [40–43]. Bevacizumab has also evolved as a first line of treatment in combination with paclitaxel in breast cancer patients by virtue of its ability to double median progression-free survival (PFS) [44]. In combination with chemoendocrine therapy (including capecitabine and vinorelbine, and letrozole) bevacizumab treatment significantly decreased the percentage of viable circulating endothelial cells and prevented the chemotherapy-induced mobilization of circulating progenitors [45]. In combination with irinotecan, bevacizumab significantly increased PFS in glioma patients [46, 47]. VEGF has emerged as a compelling therapeutic target for leukemias. Inhibition of angiogenesis in hematological malignancies interdicts the angiogenesis within the bone marrow ecosystem comprised of multiple cell types, including fibroblasts, endothelial progenitor cells, endothelial cells, dendritic cells and, malignant cells, blocking the availability of nutrients to cancer cells and disrupting crosstalk between the various cell types to curtail the malignant phenotype [48].
Cetuximab (Erbitux®)
This is a monoclonal antibody that binds the extracellular domain of epidermal growth factor receptor (EGFR), preventing ligand binding and activation of the receptor resulting in internalization and degradation of the receptor culminating in inhibition of cell proliferation and angiogenesis. Cetuximab downregulated VEGF expression in a dose-dependent manner in a human colorectal carcinoma (CRC) cell line and in human CRC mouse xenografts [49]. The xenografts also showed a significant reduction in blood vessel counts following several rounds of cetuximab treatment [49], indicating that the tumor-promoting effects of EGFR overexpression may be mediated through VEGF stimulation and tumor angiogenesis. This treatment is approved for metastatic CRC and head and neck cancer [50] in patients who are refractory to irinotecan-based chemotherapy. In combination with irinotecan (an inhibitor of topoisomerase I), cetuximab is the first monoclonal antibody that has been approved by the FDA as second-line treatment for metastatic colorectal cancer [51, 52]. In Phase I and Phase III trials [53, 54] cetuximab significantly improved the effects of radiotherapy in patients with unresectable (cannot be removed by surgery) squamous cell carcinoma of the head and neck (SCCHN). Cetuximab has also been shown to sensitize cells to radiation and chemotherapy, potentially through blocking EGFR nuclear import and the associated activation of DNA protein kinase enzymes necessary for repairing radiation- and chemotherapy-induced DNA damage [55]. Compared to radiation alone, cetuximab plus radiation therapy can nearly double the median survival in patients with a certain kind of head and neck cancer that has not spread to other parts of the body [54] making cetuximab the only drug achieving interesting response rate in second line treatment of advanced SCCHN [56]. Cetuximab was also found to be tolerated well in combination with cisplatin, or carboplatin, and fluorouracil [57, 58].
Panitumumab (Vectibix™)
It is a fully humanized anti-EGFR monoclonal antibody that binds specifically to the human EGFR. Panitumumab is a recombinant human monoclonal antibody [59]; therefore, the risk of an infusion reaction is minimized. Vectibix® is indicated as a single agent for the treatment of EGFR-expressing, metastatic colorectal carcinoma with disease progression on or following fluoropyrimidine-, oxaliplatin-, and irinotecan-containing chemotherapy regimens [60–62]. The effectiveness of Vectibix® as a single agent for the treatment of EGFR-expressing, metastatic CRC is based on progression-free survival [63, 64]. Panitumumab is used in patients who are not responding to regimens containing fluorouracil, oxaliplatin, and irinotecan [60]. Patients often receive panitumumab after receiving bevacizumab or cetuximab. Panitumumab can be given with FOLFOX (oxaliplatin, leucovorin, and fluorouracil) or FOLFIRI (irinotecan, leucovorin, and fluorouracil) regimens, or as a single agent. Currently no data are available that demonstrate an improvement in disease-related symptoms or increased survival with Vectibix® in colon cancer [65]. This drug is also being tested for aerodigestive track and head and neck cancer [66, 67].
Trastuzumab (Herceptin®)
Is a humanized monoclonal antibody that binds the extracellular domain of HER-2, which is overexpressed in 25-30% of invasive breast cancer tumors [68]. HER2-positive breast cancer is highly aggressive disease with high recurrence rate, poorer prognosis with decreased survival compared with HER2-negative breast cancer [69]. Herceptin® is designed to target and block the function of HER2 protein overexpression. This is the first humanized antibody is approved for Breast cancer [70]. Herceptin® is approved by the FDA to treat HER2 positive breast cancer that has metastasized after treatment with other anticancer drugs [71]. It is also approved to be used with other drugs to treat HER2-positive breast cancer that has spread to the lymph nodes to be used after surgery. The FDA first approved Herceptin in September 1998 [71–73]. In November 2006, the FDA approved Herceptin as part of a treatment regimen containing doxorubicin, cyclophosphamide and paclitaxel, for the adjuvant treatment of patients with HER2-positive, node-positive breast cancer (http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/default.htm). In January 2008, the FDA approved Herceptin as a single agent for the adjuvant treatment of HER2-overexpressing node-negative (ER/PR-negative or with one high-risk feature) or node-positive breast cancer, following multi-modality anthracycline-based therapy (http://biopharminternational.findpharma.com/biopharm/News/FDA-Approves-Expanded-Adjuvant-Indications-for-Her/ArticleStandard/Article/detail/518867). Trastuzumab is also being studied in the treatment of other types of cancers such as pancreatic [74], endometrial [75], lung [76], cervical [77] and ovarian cancer [78]
SMALL MOLECULE TYROSINE KINASE INHIBITORS (TKIs)
Protein tyrosine kinases have emerged as crucial targets for therapeutic intervention in cancer especially because they play an important role in the modulation of growth factor signaling. As per ClinicalTrials.gov (www.clinicaltrials.gov), there are 43 ongoing studies on tyrosine kinase inhibitors in angiogenesis. Since discussing all of them is beyond the scope of this article, we have focused our discussion on the three TKIs that are currently approved as anti-cancer therapies:
Erlotinib (Tarceva®)
Erlotinib hydrochloride (originally coded as OSI-774) is an orally available, potent, reversible, and selective inhibitor of the EGFR (ErbB1) tyrosine kinase activity. Erlotinib hydrochloride has been approved by FDA for treatment of patients with locally advanced or metastatic NSCLC after failure of at least one prior chemotherapy regimen [79, 80]. Interesting recent studies have demonstrated that since Erlotinib and Bevacizumab act on two different pathways critical to tumor growth and dissemination, administering these drugs concomitantly may confer additional clinical benefits to cancer patients with advanced disease. This combination therapy may prove to be a viable second-line alternative to chemotherapy in patients with NSCLC [81]. Also, for patients with locally advanced, unresectable or metastatic pancreatic carcinoma, Erlotinib has received FDA approval for the treatment in combination with gemcitabine [82, 83]. Erlotinib is also being studied in the treatment of other types of cancers. For example combination of Erlotinib with Bevacizumab has been evaluated in metastatic breast cancer [84], hepatocellular carcinoma [85] and in metastatic renal cancer [86] as phase II trials. Outcomes for prostate, cervical and colorectal cancers treated with Erlotinib are cautiously optimistic [87–89].
Sorafenib (Nexavar®)
Sorafenib is an orally active inhibitor of VEGFR-1, VEGFR-2, VEGFR-3, PDGFR-β, and Raf-1 tyrosine kinase activity [90]. It has received the approval of FDA for the treatment of patients with unresectable hepatocellular carcinoma [91] and advanced renal cell carcinoma [92]. However, not all advanced hepatocellular carcinoma patients were able to tolerate sorafenib and some patients experienced tumor progression [91]. Sorafenib has shown improvements in PFS in patients with renal cell carcinoma [93]. It is one of the aggressively studied drugs. According to the NCI clinical trials search results, there are about 168 active clinical trials involving sorafenib in a variety of cancers.
Sunitinib (Sutent®)
Sunitinib targets activity of multiple tyrosine kinases such as VEGFR-1, VEGFR-2, VEGFR-3, PDGFR- β, and RET [94]. It is approved by FDA as Sunitinib malate for treating advanced (metastatic) renal cell carcinoma [95]. It is also approved by FDA for gastrointestinal stromal tumor (GIST) in patients whose disease has progressed or who are unable to tolerate treatment with imatinib (Gleevec), the current treatment for GIST patients [95, 96]. Sunitinib has shown early evidence of anti-tumor activity in Phase II trials in US, European and Asian patients with locally advanced, unresectable and metastatic hepatocellular carcinoma. A Phase III trial of sunitinib in hepatocellular carcinoma is ongoing [97]. According to the NCI clinical trials search results, Sunitinib is currently evaluated in about 150 active clinical trials. It is evaluated for ovarian [98], breast [99] and non small cell lung cancer [100] among others [101].
Inhibitors of mTOR
mTOR plays a part in the PI3 kinase/AKT pathway involved in tumor cell proliferation and angiogenesis [102]. Rapamycin and related mTOR inhibitors inhibit endothelial cell VEGF expression, as well as VEGF-induced endothelial cell proliferation [103]. Inhibitors of mTOR are an important class of anti-angiogenic agents. These include: deforolimus, everolimus, rapamycin (sirolimus), and temsirolimus [104, 105]. Temsirolimus (Toricel™) is a small molecule inhibitor of mTOR, approved for treating advanced renal cell carcinoma [106]. It is a type of rapamycin analog and a type of serine/threonine kinase inhibitor, it is also called CCI-779. In pre-clinical models combination therapy for treating breast cancer using anti-estrogen, ERA-923, and temsirolimus has been successfully tested [107]. It is found to be highly effective against human melanoma when tested in combination with cisplatin and DTIC (in independent studies) in a SCID mouse xenotranplantation model [108, 109]. There are over 41 active studies of Temsirolimus for a variety of solid tumors [110]. mTOR inhibition has also been strongly advocated in as a putative cancer therapeutic strategy for urologic malignancies [111]. In a pilot study (6 patients) with imatinib-resistant CML, rapamycin induced major and minor leukocyte responses, with an observed decrease in the mRNA levels of VEGFA in circulating leukaemic cells [112]. Combination treatments for breast cancer with aromatase inhibitor [113] and letrozol [114] are also being evaluated. Rapamycin treatment brought partial responses (>50% reduction in the absolute number of blood blasts) and stable disease in adult refractory/relapsed AML [115]. In a recent report, Deforolimus was studied in a Phase 2 trial in pretreated patients with various hematological malignancies, including ALL, AML, CLL, CML, MDS, agnogenic myeloid metaplasia, mantle cell lymphoma and T-cell leukemia/lymphoma [116]. Overall, 40% of deforolimus-treated patients experienced hematological improvement or stable disease.
OTHER ANGIOGENIC AGENTS
Bortezomib (Velcade®)
Is a proteasome inhibitor that disrupts signaling of cancer cells, leading to cell death and tumor regression. It is the first compound in its class to be used in clinical practice. It has indirect anti-angiogenic properties [117]. While its exact mechanism is not understood, it induces the pro-apoptotic BH3-only family member NOXA in a p53 independent fashion triggering of a caspase cascade culminating in apoptosis in melanoma and myeloma cells [118]. It is FDA-approved for the treatment of myeloma that has relapsed after two prior treatments (or where resistance has developed following the last treatment). It was also found to induce high quality responses as third line salvage therapy with acceptable toxicity in a significant proportion of homogeneously pre-treated myeloma patients with progressive disease after autologous transplantation and thalidomide. [119]. In a Phase 3 trial involving 669 myeloma patients treated with at least one prior therapy, bortezomib increased median, improved overall survival, and increased response rate, compared with high-dose dexamethasone [120]. In combination with doxorubicin and gemcitabine, bortezomib was also found to be effective in heavily pretreated, advanced Cutaneous T cell Lymphomas (CTCL) [121]. Bortezomib was also reported to be active as a single agent for patients with relapsed/refractory CTCL and Peripheral T Cell Lymphoma (PTCL) with skin involvement [122]. On the contrary, the use of bortezomib was discouraged after a phase II study revealed that found in combination with dexamethasone, bortezomib is not active in heavily pre-treated patients with relapsed Hodgkin’s lymphoma [123, 124].
Thalidomide (Thalomid®)
Possesses immunomodulatory, anti-inflammatory, and anti-angiogenic properties, although the precise mechanisms of action are not fully understood. Thalidomide was the first angiogenesis inhibitor to demonstrate clinical efficacy in multiple myeloma [37, 125]. Specifically in myeloma, thalidomide down-regulated VEGF secretion from bone marrow endothelial cells obtained from patients with active disease. In a landmark Phase 2 clinical trial, 169 previously treated patients with refractory myeloma received thalidomide monotherapy [126]. Partial response, was achieved in 30% of patients, and 14% achieved a complete or nearly complete remission. The survival rate at 2 years was 48%. These results led to many subsequent clinical studies of thalidomide in myeloma, leading ultimately to FDA approval of the drug in 2006, for the treatment of newly diagnosed multiple myeloma, in combination with dexamethasone. In the pivotal Phase 3 trial, the response rate in patients receiving thalidomide plus dexamethasone was 63% compared to 41% with dexamethasone alone [127]. Long-term outcome measures, including time-to-progression (TTP) and PFS, were recently reported for a 470 patient randomized, placebo-controlled Phase 3 clinical trial of a similar protocol in newly diagnosed multiple myeloma, with comparable overall response rates [128]. Significant increases resulted in both median TTP and median PFS for the thalidomide plus dexamethasone group versus dexamethasone alone.
Thalidomide was found to be moderately tolerated and minimally effective in patients with histologically proven advanced hepatocellular carcinoma [129]. Thalidomide provided no survival benefit for patients with multiple, large, or midbrain metastases when combined with WBRT (whole-brain radiation therapy) [130]. On the contrary, thalidomide did not significantly add to the efficacy of the fludarabine, carboplatin, and topotecan (FCT) regimen in poor prognosis AML patients [131] and was also ineffective in improving prognosis or decreasing plasma VEGF levels in patients with persistent or recurrent leiomyosarcoma of the uterus [132].
METRONOMIC THERAPY
While conventional anti-angiogenic therapy is based on Maximum Tolerated Doses (MTD), the cells involved in angiogenesis may regenerate during the three- to four-week interval between cycles of the chemotherapy. Taking advantage of the fact that endothelial cells are about 10–100 times more susceptible to chemotherapeutic agents than cancer cells, therapy based on daily, oral, low-dose chemotherapeutic drugs was designed. Metronomic chemotherapy refers to the close, rhythmic administration of low doses of cytotoxic drugs, with minimal or no drug-free breaks, over prolonged periods. Metronomic therapy appears promising mainly due to the fact that its anti-angiogenic and anti-tumorigenic effects are accompanied by low toxicity, limited side effects, no need for hospitalization and allowing for feasible combinations with selective inhibitors of angiogenesis. There are several foreseeable advantages and opportunities for metronomic chemotherapy: activity against the parenchymal and stromal components, pro-apoptotic activity, reduction of the likelihood of emergence of acquired resistance, feasibility of long term administration and acceptable systemic side effects [133]. In a pilot phase II study conducted by Correale et al [134] to investigate the toxicity and activity of the novel metronomic regimen of weekly cisplatin and oral etoposide in high-risk patients with NSCLC, the objective response rate was 45.2%, disease control was 58.1%, meantime to progression and survival were 9 and 13 months, respectively. Pharmacokinetic analysis showed that this regimen allowed a greater median monthly area under the curve of the drugs than conventional schedules. In a Phase I trial of metronomic dosing of docetaxel and thalidomide, of the 26 patients with advanced tumors enrolled, prolonged freedom from disease progression was observed in 44.4% of the evaluable patients [135].
Circulating endothelial progenitor cells (EPCs) also participate in tumor angiogenesis. In a study comparing the effects of metronomic chemotherapy over conventional dose-dense chemotherapy, it was found that the numbers of circulating EPCs and the plasma levels of VEGF increased sharply, doubling pre-therapeutic levels at day 21 after conventional chemotherapy, whereas under low-dose metronomic chemotherapy, the numbers of circulating EPCs decreased significantly and VEGF plasma concentrations remained unchanged. These observations provide evidence that conventional dose-dense chemotherapy leads to rebound EPC mobilization even when given with adjuvant intention, while low-dose metronomic scheduling of cytotoxic substances such as trofosfamide may sharply reduce EPC release into the circulation. [136].
Combined bevacizumab and metronomic oral cyclophosphamide was also discovered to be a safe and effective regimen for heavily pre-treated ovarian cancer patients [137]. Treatment with metronomic capecitabine and cyclophosphamide in combination with bevacizumab was shown to be effective in advanced breast cancer and additionally was minimally toxic [138]. Metronomic treatment with carboplatin and vincristine associated with fluvastatin and thalidomide significantly increased survival of pediatric brain stem tumor patients. Tumor volume showed a significant reduction accompanied by increased quality of life [139]. Thus, given the fact that the most evident effect of selective anti-angiogenic agents (i.e. bevacizumab) is the significant prolonging of the duration of response obtainable by chemotherapy alone, with minimal possible side effects of cytotoxic agents given in association metronomic chemotherapy should be considered both as novel up-front or maintenance treatment in patients with biologically poorly aggressive advanced cancer diseases [140].
Overall, metronomic chemotherapy was able to induce tumor stabilization and prolong the duration of clinical benefit, without much associated toxicity. Emerging evidence suggests that metronomic chemotherapy could also activate the host immune system and potentially induce tumor dormancy [141–143].
CONCLUSIONS AND FUTURE PERSPECTIVES
While angiogenesis as a hallmark of tumor development and metastasis is now a validated target for cancer treatment, the overall benefits of anti-angiogenic drugs from the perspective of impacting survival have left much to desire, endorsing a need for developing more effective therapeutic regimens e.g., combining anti-angiogenic drugs with established chemotherapeutic drugs [144, 145]. There are now several agents that target the tumor vasculature through different pathways, either by inhibiting formation of the tumor neovasculature or by directly targeting the mature tumor vessels. The main body of evolving evidence suggests that their effects are compounded by their synergistic use with conventional chemotherapy rather than individual agents. Anti-angiogenic drugs such as bevacizumab can bring about a transient functional normalization of the tumor vasculature. This can have an additive effect when co-administered with chemo/radiotherapy. But long term inhibition of angiogenesis reduces tumor uptake of co-administered chemotherapeutic agents. This underscores the need for discovering new targets for anti-angiogenic therapy in order to effectively prohibit angiogenesis and circumvent mechanisms that contribute to resistance mechanisms that emerge with long term use of anti-angiogenic therapies. It also warrants a need to define reliable surrogate indicators of effectiveness of the anti-angiogenic therapy as well as dependable markers for identifying the patients who are most likely to benefit from the combination of anti-angiogenic therapy and conventional chemotherapy.
Several new frontiers are emerging. New advances in understanding endothelial cells, which constitute the tumor vasculature, towards developing antiangiogenic strategies are one of the important ones [146, 147]. Novel cellular targets such as integrins and microRNAs and novel treatment options such as possible use of pharmaconutrients to modulate angiogenic pathways need careful testing and evaluation [148–151]. Finally, the administration of these drugs in a metronomic schedule is likely to improve the overall response to anti-angiogenic drugs making it feasible to administer them with conventionally toxic chemotherapeutic drugs, thus increasing the armamentarium of drug combinations that can be employed for treatment.
9.5.4 Angiogenesis inhibitors in cancer therapy: mechanistic perspective on classification and treatment rationales
Angiogenesis, a process of new blood vessel formation, is a prerequisite for tumor growth to supply the proliferating tumor with oxygen and nutrients. The angiogenic process may contribute to tumour progression, invasion and metastasis, and is generally accepted as an indicator of tumor prognosis. Therefore, targeting tumor angiogenesis has become of high clinical relevance. The current review aimed to highlight mechanistic details of anti-angiogenic therapies and how they relate to classification and treatment rationales. Angiogenesis inhibitors are classified into either direct inhibitors that target endothelial cells in the growing vasculature or indirect inhibitors that prevent the expression or block the activity of angiogenesis inducers. The latter class extends to include targeted therapy against oncogenes, conventional chemotherapeutic agents and drugs targeting other cells of the tumor micro-environment. Angiogenesis inhibitors may be used as either monotherapy or in combination with other anticancer drugs. In this context, many preclinical and clinical studies revealed higher therapeutic effectiveness of the combined treatments compared with individual treatments. The proper understanding of synergistic treatment modalities of angiogenesis inhibitors as well as their wide range of cellular targets could provide effective tools for future therapies of many types of cancer.
Two major processes of blood vessel formation are implicated in the development of vascular system: vasculogenesis and angiogenesis. Vasculogenesis prevails in the embryo and refers to the formation ofde novo blood vessels by in situ differentiation of the mesoderm-derived angioblasts and endothelial precursors. Angiogenesis is the formation of new capillaries from pre-existing vessels and circulating endothelial precursors (Polverini, 2002; Chung et al., 2010; Ribatti and Djonov, 2012). Angiogenesis is a tightly controlled dynamic process that can occur physiologically in those tissues that undergo active remodeling in response to stress and hypoxia (Carmeliet, 2003; Folkman, 2007). However, it can be aberrantly activated during many pathological conditions such as cancer, diabetic retinopathy as well as numerous ischemic, inflammatory, infectious and immune disorders (Carmeliet, 2003; Ali and El-Remessy, 2009; Willis et al., 2011). Although the concept of proposing angiogenesis inhibitors as anticancer drugs received considerable skepticism when first presented by Dr. Folkman in the early 1970s (Folkman, 1971), active research in the field and subsequent clinical trials eventually resulted in US Food and Drug Administration (FDA) approval of bevacizumab for colorectal cancer in 2004 (Cohen et al., 2007). Since then, several angiogenic inhibitors have been identified. This review will provide an overview of the key mechanisms involved in tumor angiogenesis, classification of angiogenesis inhibitors as well as treatment rationales from the mechanistic point of view.
Sustained angiogenesis as a hallmark of cancer
Proliferating tumours tend to activate an angiogenic phenotype to fulfil their increased demand of oxygen and nutrients (Hanahan and Folkman, 1996; Carmeliet, 2005). Additionally, paracrine release of anti-apoptotic factors from activated endothelial cells in the newly formed vasculature supplies tumour cells with a survival privilege (Folkman, 2003). Consequently, in order to progress, tumors tend to activate an event called ‘angiogenic switch’ by shifting the balance of endogenous angiogenesis inducers and inhibitors towards a pro-angiogenic outcome. As a result, dormant lesion progresses into outgrowing vascularized tumor and eventually into a malignant phenotype (Hanahan and Folkman, 1996; Baeriswyl and Christofori, 2009). Hypoxia drives such imbalance through up-regulation of the transcription factor hypoxia inducible factor-1α (HIF-1α), which in turn increases the expression of many angiogenesis inducers as well as suppresses the expression of endogenous angiogenesis inhibitors (Pugh and Ratcliffe, 2003). In spite of that, accumulating evidence indicates that angiogenic cascade can be also driven by alternative HIF-1-independent pathways (Mizukami et al., 2007; Arany et al., 2008; Lee, 2013).
As summarized in Table 1, the angiogenesis inducers are a wide range of mediators that include many growth factors, a plethora of cytokines, bioactive lipids, matrix-degrading enzymes and a number of small molecules (Folkman, 1995; Folkman, 2003; Lopez-Lopez et al., 2004; Bouis et al., 2006; El-Remessy et al., 2007; Bid et al., 2011; MacLauchlan et al., 2011; Murakami, 2011; Fagiani and Christofori, 2013; Qin et al., 2013). Pro-angiogenic growth factors mostly activate a series of surface receptors in a series of paracrine and autocrine loops with the VEGF-A signaling representing the critical rate-limiting step, physiologically and pathologically. VEGF-A (traditionally known as VEGF) is the most potent VEGF isoform that acts mainly on VEGF receptor 2 (VEGFR2) to mediate vascular permeability, endothelial proliferation, migration and survival (Takahashi and Shibuya, 2005; Bouis et al., 2006). In spite of the well-established master roles of VEGF signaling in literature, those processes are probably accomplished through a highly regulated interplay between VEGF and other pro-angiogenic factors. In this context, basic fibroblast growth factor (bFGF) activation of the endothelium is required for maintenance of VEGFR2 expression and the ability to respond to VEGF stimulation (Murakami et al., 2011). Similarly, sphingosine-1-phosphate (S1P), a pleiotropic bioactive lipid that can directly contribute to tumor angiogenesis (reviewed in Sabbadini, 2011), is needed for VEGF-induced blood vessel formation, indicating the cooperation between S1P and VEGF in tumor angiogenesis (Visentin et al., 2006). As a net result, the pro-angiogenic interplay of those ligands and others dominates over the activities of two dozen endogenous angiogenesis inhibitors that can be either matrix-derived inhibitors or non–matrix-derived inhibitors (Nyberg et al., 2005).
Table 1. Pro-angiogenic mediators implicated in tumor angiogenesis
The multistep angiogenic process starts with vasodilation and increased permeability of existing vessels in response to tumor cell-secreted VEGF. This is accompanied by loosening of pericytes covering mediated by angiopoietin-2 (ANG2), a ligand of tyrosine kinase with immunoglobulin-like and EGF-like domains 2 (TIE2) receptor (Bergers and Benjamin, 2003; Jain, 2003; Fagiani and Christofori, 2013). Meanwhile, many secreted matrix-degrading enzymes, such as MMPs and heparanases, function in concert to dissolve the basement membrane and to remodel the extracellular matrix (ECM) as well as to liberate more pro-angiogenic growth factors (bFGF and VEGF) from matrix heparan sulfate proteoglycans (HSPGs) respectively (Houck et al., 1992; Whitelock et al., 1996; Vlodavsky and Friedmann, 2001; Tang et al., 2005; van Hinsbergh and Koolwijk, 2008). The overall chemotactic angiogenic stimuli guide endothelial cells to migrate, to align into tube-like structures and to eventually form new blood vessels. However, such blood vessels are characterized by being disorganized, chaotic, hemorrhagic and poorly functioning (Bergers and Benjamin, 2003).
The angiogenic phenotype in tumor micro-environment can further be sustained and extravagated by the recruitment of other types of stromal cells. Stromal cells such as fibroblasts, mesenchymal stem cells and various bone marrow-derived myeloid cells including macrophages, TIE2-expressing monocytes, neutrophils and mast cells contribute to tumor angiogenesis through their production of growth factors, cytokines and proteases (Murdoch et al., 2008; Joyce and Pollard, 2009; Cirri and Chiarugi, 2011). For example, in response to cancer cell-derived TGF-β, PDGF or bFGF, fibroblasts are transformed to an activated phenotype with a higher proliferative activity and myofibroblastic characteristics (Kalluri and Zeisberg, 2006; Cirri and Chiarugi, 2011). Such carcinoma-associated fibroblasts (CAFs) were shown to promote angiogenesis and metastasis by secreting large amounts of MMP-2 and MMP-9 as well as by expressing many cytokines and chemokines that resulted in immune cell infiltration (Gerber et al., 2009; Giannoni et al., 2010). Furthermore, it has been shown that PDGF-C produced by CAFs is able to elicit VEGF production from tumor cells, thereby sustaining the angiogenic shift (Crawford et al., 2009). Similarly, tumor-associated macrophages (TAMs), one of the bone marrow myeloid-derived cells, are induced to develop into polarized type II (alternatively activated or M2 macrophages), upon exposure to tumor hypoxia and tumor cell-derived cytokines (Leek et al., 2002; Rogers and Holen, 2011). M2 macrophages tend to produce many pro-angiogenic growth factors, cytokines and matrix-degrading enzymes such as VEGF, PDGF, bFGF, TNF-α, COX-2, MMP-9, MMP-7 and MMP-12 (Lewis and Pollard, 2006).
From another perspective, angiogenesis may be dispensable for progression of some malignancies. For example, some tumours may co-opt pre-existent vessels as an alternative way to obtain blood supply. Vessel co-option was first described in the brain, one of the most densely vascularized organs, in which tumours may develop in earlier stages without the activation of angiogenic response (Holashet al., 1999; Leenders et al., 2002; Bergers and Benjamin, 2003; Hillen and Griffioen, 2007). In another example, hypovascularized tumors such as pancreatic ductal adenocarcinoma may involve certain adaptation to flourish in the absence of prominent angiogenesis (Bergers and Hanahan, 2008). Obviously, in both cases, tumors may be intrinsically indifferent to angiogenesis inhibitors. However, in most other cases, therapy directed towards the vasculature of solid tumors is being considered as an important direction in cancer treatment.
Classification of angiogenesis inhibitors
Growth of newly formed vessels in tumor micro-environment can be inhibited directly by targeting endothelial cells in the growing vasculature or indirectly by targeting either tumor cells or the other tumor-associated stromal cells. Therefore, angiogenesis inhibitors can be classified into direct and indirect inhibitors (Kerbel and Folkman, 2002; Folkman, 2007).
Direct endogenous inhibitors of angiogenesis
Direct endogenous inhibitors of angiogenesis, such as angiostatin, endostatin, arrestin, canstatin, tumstatin and others, are fragments released on proteolysis of distinct ECM molecules. Endogenous inhibitors prevent vascular endothelial cells from proliferating, migrating in response to a spectrum of angiogenesis inducers, including VEGF, bFGF, IL-8 and PDGF (Kerbel and Folkman, 2002; Abdollahi et al., 2004; Mundel and Kalluri, 2007; Ribatti, 2009). This direct anti-angiogenic effect may be mediated by interference with endothelial integrins along with several intracellular signaling pathways (Mundel and Kalluri, 2007). For example, the ability of tumstatin-derived active peptide to inhibit angiogenesis and tumour growth is associated with the expression of the adhesion receptor, αvβ3 integrin, on tumor endothelial cells (Eikesdal et al., 2008). Through binding αvβ3 integrin, full tumstatin was found to inhibit endothelial cell activation of focal adhesion kinase, PI3K, Akt, mammalian target of rapamycin (mTOR) and others (Maeshima et al., 2002). Direct targeting of those signaling pathways by endogenous inhibitors was thought to be the least likely to induce acquired drug resistance because they target endothelial cells with assumed genetic stability rather than unstable mutating tumour cells (Kerbel and Folkman, 2002). However, endostatin has not yet led to any documented benefit to patients in randomized phase III trials, or even modest activity in phase II trials (Ellis and Hicklin, 2008).
Indirect inhibitors of angiogenesis
Indirect inhibitors of angiogenesis classically prevent the expression or block the activity of pro-angiogenic proteins (Folkman, 2007). For example, Iressa, an EGF receptor (EGFR) TK inhibitor (TKI), blocks tumour expression of many pro-angiogenic factors; bevacizumab, a monoclonal antibody, neutralizes VEGF after its secretion from tumour cells whereas sunitinib, a multiple receptor TKI, blocks the endothelial cell receptors (VEGFR1, VEGFR2 and VEGFR3), preventing their response to the secreted VEGF (Folkman, 2007; Roskoski, 2007). In addition, this class extends to include conventional chemotherapeutic agents, targeted therapy against oncogenes and drugs targeting other cells of the tumor micro-environment (Kerbel et al., 2000; Ferrara and Kerbel, 2005).
Conventional chemotherapeutic agents
Conventional chemotherapeutic agents have been shown to have anti-angiogenic properties in addition to the ability to induce direct cancer cell death. Such chemotherapeutic agents can affect the endothelial cell population in the tumour bed during treatment cycles because they have significantly higher proliferation rates than resting endothelium outside a tumor, making them more susceptible to cytotoxic effect (Kerbel et al., 2000; Folkman, 2003). However, the cyclic treatment rationale of cytotoxic drugs allows the potential damage to the tumour vasculature to be repaired during the long breaks. Thus, continuous low doses of chemotherapeutic agents were suggested as a way to reduce side effects and drug resistance (Drevs et al., 2004). This modality is termed metronomic therapy, and clinically, it refers to the daily administration of 5–10% of the phase II-recommended dose of the chemotherapeutic agent (Penel et al., 2012). The extended use of such low doses of cytotoxic agents elicits an anti-angiogenic activity through induction of endothelial cell apoptosis and decreasing the level of circulating endothelial precursors (Hamano et al., 2004; Shahrzad et al., 2008). In clinical investigations, metronomic dosing of cyclophosphamide and others showed promising efficacy in patients with advanced, multiple metastasized and/or multiple pretreated solid tumours (Lord et al., 2007; Fontana et al., 2010; Nelius et al., 2011; Gebbia et al., 2012; Briasoulis et al., 2013; Navid et al., 2013).
VEGF-targeted therapy
VEGF-targeted therapy includes neutralizing antibodies to VEGF (e.g. bevacizumab) or VEGFRs (e.g. ramucirumab), soluble VEGFR/VEGFR hybrids (e.g. VEGF-Trap) and TKIs with selectivity for VEGFRs (e.g. sunitinib and sorafenib; Baka et al., 2006; Ellis and Hicklin, 2008; Hsu and Wakelee, 2009). Bevacizumab, a humanized monoclonal antibody against all isoforms of VEGF-A, has been approved for the treatment of colorectal, lung, glioblastoma and renal cell carcinoma (Hsu and Wakelee, 2009). Many other clinical trials with promising efficacy were also conducted in other cancers such as head and neck cancer, hepatocellular carcinoma, ovarian cancer, metastatic melanoma and gastric cancer (Argiris et al., 2011; 2013; Burger et al., 2011; Ohtsu et al., 2011; Fang et al., 2012; Minor, 2012; Schuster et al., 2012; Van Cutsem et al., 2012). However, for metastatic breast cancer, bevacizumab had been initially granted an accelerated FDA approval, which was later withdrawn due to lack of improvement evidence in disease-related symptoms or overall survival (Burstein, 2011; Montero et al., 2012). Similarly, clinical trials showed that the addition of bevacizumab to the treatment regimens of advanced pancreatic cancer did not extend overall survival (Chiu and Yau, 2012). The neutralization of VEGF-A can also be achieved by soluble receptor construct (VEGF-Trap) that monomerically ‘traps’ the different isoforms of VEGF-A, in addition to VEGF-B and placental growth factor (Rudge et al., 2007). VEGF-Trap showed clinical benefit in a phase III trial of oxaliplatin pretreated metastatic patients with colorectal cancer, and is currently being investigated in a prostate cancer phase III trial (Gaya and Tse, 2012). TKIs are small molecules with different chemical structures that have the ability to interact physically with the highly conserved kinase domain shared by different VEGFRs as well as PDGF receptors (PDGFRs), FGF receptors (FGFRs), EGFR, Raf kinases and c-Kit (a receptor of the pluripotent cell growth factor, stem cell factor). Such interaction directly inhibits tyrosine phosphorylation and the subsequent many downstream pro-angiogenic signaling networks (Baka et al., 2006; Ivy et al., 2009). Those multi-targeted TKIs demonstrated efficacy against various solid malignancies in different clinical trials, some of which have lead eventually to FDA approval of sunitinib and sorafenib. Sunitinib, known to inhibit several receptor TKs (RTKs) including VEGFR1–3, PDGFR-α, PDGFR-β, c-Kit, colony-stimulating factor-1 receptor (CSF-1R) and Flt-3, was approved for the treatment of renal cell carcinoma and gastrointestinal stromal cell tumours. Sorafenib that acts also by inhibiting VEGFR1–3 and PDGFR-β in addition to the serine–threonine kinases Raf-1, B-Raf, was approved for hepatocellular carcinoma in addition to renal cell carcinoma (Llovet et al., 2008; Ivy et al., 2009; Huang et al., 2010).
FGF-targeted therapies
FGF-targeted therapies were recently reconsidered as promising anti-angiogenic and anti-tumor agents after a long period of little attention for drug development, partly due to redundancy (Bono et al., 2013). The FGFR superfamily with its 18 ligands and four receptors has been involved in endothelial cell migration, proliferation and differentiation (Presta et al., 2005). Therapeutic targeting of FGF/FGFR signalling was accomplished by either monoclonal antibodies that inhibit FGFs binding, small molecules that inhibit FGFR TK activity or allosteric modulators that bind the extracellular FGFR domain. Monoclonal antibodies against bFGF displayed potent anti-tumor and anti-angiogenic effects in different preclinical cancer models, which warrant further clinical evaluation (Zhao et al., 2010; Wang et al., 2012). Pan inhibitors of the FGFR TKs such as AZD4547 (blocks the activity of FGFR1–3) and ponatinib (blocks all the FGFR isoforms) elicited potent anti-tumor activities in preclinical investigations so they are currently being evaluated in clinical trials. Those inhibitors displayed the greatest potency in FGFR-driven cancer models, which may be attributed to the interference with the oncogenic functions of either amplified or constitutively active FGFR (Dutt et al., 2011; Zhao et al., 2011; Gavine et al., 2012; Gozgit et al., 2012). Accordingly, further studies are needed to evaluate the relative contribution of angiogenic versus oncogenic inhibitory mechanisms towards the overall anti-tumor activity. The allosteric antagonist of the FGFR, SSR128129E, showed a strong anti-angiogenic activity in addition to tumour growth and metastasis inhibitory effects in animal models of arthritis and cancer respectively. Because allosteric modulators leave a residual level of baseline signalling, they have the ability to fine-tune target biological responses. As a result, allosteric multi-FGFR inhibitors may have an improved benefit/risk ratio that is not attainable with the other TKIs (Bonoet al., 2013; Herbert et al., 2013). However, preclinical findings suggest that long-term clinical outcomes may improve with blockade of additional pro-angiogenic RTKs that may also reduce the risk of drug resistance (Hilberg et al., 2008). For example, dual inhibition of VEGFRs and FGFRs using brivanib produced enduring tumour stasis and angiogenic blockade following the failure of VEGF-targeted therapies (Allen et al., 2011). Furthermore, triple inhibition of FGFRs, VEGFRs and PDGFR(s) using dovitinib (TKI258) or nintedanib (BIBF 1120) displayed broad-spectrum anti-tumour activities in several tumour xenograft models as well as promising data in clinical trials. Combined inhibition of FGFR/VEGFR/PDGFR targets not only tumour cells, but also endothelial cells, pericytes and smooth muscle cells, resulting in an effective inhibition of tumour growth, angiogenesis and metastasis even in advanced tumour stages (Hilberg et al., 2008; Ledermann et al., 2011; Taeger et al., 2011; Chenet al., 2012; Angevin et al., 2013).
Oncogene-targeted therapy
Oncogenes, genes that cause the transformation of normal cells into cancerous cells, are thought to up-regulate many pro-angiogenic proteins. Therefore, anticancer drugs that were developed for their capacity to block an oncogene also have an indirect anti-angiogenic activity (Kerbel et al., 2000; Bergers and Benjamin, 2003; Folkman, 2003). For example, dasatinib and other inhibitors of sarcoma (Src), an aberrantly activated non-RTK associated with many human malignancies, showed potent anti-angiogenic effects through the down-regulation of VEGF and IL-8 (Summy et al., 2005; Han et al., 2006; Haura et al., 2010). Another example is to target the oncogenic Ras using farnesyl transferase (FT) inhibitors, which inhibit post-translational farnesylation of Ras that governs the latter’s activity (Awada et al., 2002). FT inhibitors were found to inhibit tumor VEGF expression and block FTase-dependent Ras activation, which is critically involved in VEGF-elicited angiogenic signal transduction and angiogenesis (Han et al., 2005; Izbicka et al., 2005; Kim et al., 2010). In addition to classical oncogenes inhibition, interference with other tumor-deregulated signaling pathways would offer another approach in targeting angiogenesis. For example, inhibitors of heat shock protein 90 (HSP90), a chaperone molecule known to protect oncoproteins from misfolding and degradation in the protein-rich intracellular environment, were found to prevent VEGF production and to disrupt multiple pro-angiogenic signalling pathways in numerous cancer cells. They were also shown to inhibit tumour growth and vascularity of different human tumor xenografts (Sanderson et al., 2006; Langet al., 2007; Eccles et al., 2008; Trepel et al., 2010; Moser et al., 2012). Proteasome inhibitors, such as bortezomib (PS-341) or MG-132, were also shown to reduce tumour growth and vascularity of squamous cell carcinoma and pancreatic cancer xenograft probably through inhibition of NF–κB-dependent release of pro-angiogenic gene products, VEGF and IL-8 (Sunwoo et al., 2001; Nawrocki et al., 2002; Matsuo et al., 2009). Similarly, inhibition of B-cell lymphoma 2 (Bcl-2), a prosurvival protein that regulates apoptosis by preventing the mitochondrial release of pro-apoptogenic factors, was shown to prevent NF-κB-mediated release of the pro-angiogenic factors IL-8 and CXC chemokine ligand 1 (CXCL1) as well as VEGF in tumor-associated endothelial cells and pancreatic cell lines respectively (Karl et al., 2005; Wang et al., 2008). Moreover, (−)-gossypol, a natural BH3 mimetic that inhibits BH3 domain of Bcl-2 as well as related prosurvival proteins (Bcl-xL and Mcl-1), was shown to remarkably decrease microvessel density in human prostate tumour PC-3 xenografts through decrease of VEGF and IL-8 release as well as blocking multiple steps in VEGF-activated biological events (Karaca et al., 2008; Pang et al., 2011).
Matrix degrading and remodelling-targeted therapy
Matrix degrading and remodelling are activated by tumors to modify local micro-environment, which in turn promote their angiogenic potential (Bergers et al., 2000; Vlodavsky and Friedmann, 2001). Up-regulation of expression and activity of several endogenous MMPs including MMP-2, MMP-9 as well as MMP-3 and MMP-7 have been identified in invasive tumors (for a review, see Bourboulia and Stetler-Stevenson, 2010). Consequently, inhibitors of MMPs were extensively pursued as a therapeutic strategy for treating cancer. Unfortunately, MMPs intervention strategies had met with limited clinical success because of severe toxicities and associated metastasis-promoting effect (Coussens et al., 2002; Devy et al., 2009). Furthermore, the paradoxical roles of tissue inhibitors of metalloproteinases (TIMPs) may contribute to such failure depending on the net balance of TIMPs and MMPs in tumour stroma (Jiang et al., 2002). As a result, efforts were directed at therapies exploiting endogenous MMP inhibitors, TIMPs or monoclonal antibodies against individual MMPs (Martens et al., 2007; Jarvelainen et al., 2009). For example, DX-2400, a highly selective fully human MMP-14 inhibitory antibody, was found to block pro-MMP-2 processing on tumor and endothelial cells, inhibited angiogenesis, and slowed tumor progression and formation of metastatic lesions (Devy et al., 2009). Alternatively, in order to reduce toxicity and enhance drug delivery, polymeric nanoparticulate delivery systems could be used to target individual components of ECM. For example, targeted delivery of antisense inhibitors of laminin-8, a vascular basement membrane component, by conjugation to the natural drug carrier β-poly(L-malic acid) significantly reduced tumour microvessel density and increased animal survival in an experimental model of glioblastoma (Fujita et al., 2006). Similarly, a nano delivery system that incorporate peptides against proteolytically processed type IV collagen significantly accumulated in tumors and blocked angiogenesis in experimental models (Mueller et al., 2009). However, the highly sulfated oligosaccharides, Heparan (HS) mimetics highly sulfated oligosaccharides, were shown to have a heparanase-inhibiting effect sequestering, in turn, many heparan sulfate proteoglycan (HSPG)-binding factors (Johnstone et al., 2010; Dredge et al., 2011). In preclinical studies, HS mimetics have effectively targeted multiple HSPG-dependent functions and have resulted in decreased in vivo tumor growth, tumor invasion, tumor metastasis and angiogenesis (Johnstone et al., 2010; Dredge et al., 2011; Zhou et al., 2011). Clinically, the heparanase inhibitor PI-88 showed preliminary efficacy as an adjunct therapy for post-operative hepatocellular carcinoma (Liu et al., 2009).
Tumour-associated stromal cell-targeted therapy
Tumour-associated stromal cells crosstalk is a perquisite for the formation of a tumour vasculature, an essential step for tumour progression (Lorusso and Ruegg, 2008). Interference with those crosstalk circuits through intervention of cellular adhesion (highlighted in next paragraph) or tumor-induced recruitment of different stromal cells may be considered as an indirect way of anti-angiogenic therapy (Ferrara and Kerbel, 2005). The latter can be supported by studies in which inhibition of macrophage infiltration, for example, by either genetic ablation of the macrophage CSF-1 or liposomal clodronate-induced macrophage depletion, was shown to delay the angiogenic switch and malignant transition (Giraudo et al., 2004; Lin et al., 2006). Furthermore, CSF-1R kinase inhibitors were found to reduce tumor-associated vascularity in two different tumor mouse models (Kubota et al., 2009; Mantheyet al., 2009). In addition, clodronate and other related bisphosphonates, originally used to treat skeletal complications in patients with tumour-induced osteolysis, were shown to exert potent anti-tumour and anti-angiogenic effects in many other studies (Fournier et al., 2002; Santini et al., 2003; Stathopoulos et al., 2008). Zoledronic acid, a third-generation bisphosphonate, was also found to reduce a number of tumour-associated macrophages and shift their phenotype from M2 to M1, resulting in a reduction in TAM-associated production of VEGF in murine models of spontaneous mammary carcinogenesis and mesothelioma (Coscia et al., 2010; Veltman et al., 2010). Clinically, repeated low-dose therapy with zoledronic acid, which maintains active drug plasma concentration, was able to induce an early remarkable and long-lasting decrease of VEGF levels in patients with cancer (Santini et al., 2007). In another example, inhibition of mobilization of neutrophils, from bone marrow and their infiltration into tumour, using neutralizing anti–prokineticin-2, an antibody against a secreted protein known also as BV8, was shown to impair the initial angiogenic switch in a multistage pancreatic beta cell tumorigenesis model (Shojaei et al., 2008). Furthermore, the neutralizing anti-BV8 was found to prevent myeloid cell-dependent tumour angiogenesis in several xenograft models (Shojaei et al., 2007). Cancer-associated fibroblasts (CAF) can also be targeted with thapsigargin analogue coupled with peptides specific for fibroblast activation protein (FAP), a CAF membrane-bound protease whose catalytic site has access to the peritumoural fluid of the tumor micro-environment. This extracellular activation results in the death of CAFs as well as pericytes and endothelial cells within milieu of different human tumor xenografts (Brennen et al., 2012).
Cell adhesion molecules (CAMs)-targeted therapy
CAMs are cell surface proteins known to be involved in binding with other counter-receptors on adjacent cells or surrounding ECM macromolecules (Aplin et al., 1998). Many CAMs, such as αv-integrins, E-selectin, N-cadherin and VE-cadherin, have been implicated in tumour angiogenesis (Bischoff, 1997; Tei et al., 2002; Nakashima et al., 2003; Weis and Cheresh, 2011). For example, αv-integrins are expressed on surface of endothelial cells and can determine whether cells can adhere to and survive in a particular micro-environment. A number of matrix-derived fragments have the ability to act as endogenous angiogenesis inhibitors through binding to integrins on endothelial cells, disrupting physical connections and suppressing signalling events associated with cell survival, migration and proliferation (Nyberg et al., 2005). Consequently, integrins antagonism using peptidomimetics (e.g. cilengitide), monoclonal antibodies (e.g. volociximab) or oral small-molecule compounds have been investigated in a wide range of malignancies (Huveneers et al., 2007). Cilengitide is a cyclized pentapeptide peptidomimetic designed to compete for the arginine-glycine-aspartic acid (RGD) peptide sequence, thereby blocking the ligation of the αvβ3 and αvβ5 integrins to matrix proteins (Hariharan et al., 2007). Cilengitide is mainly under clinical development for glioblastoma; however, clinical trials of other malignancies such as head and neck cancer as well as lung cancer were also initiated (Reardon and Cheresh, 2011; Vermorken et al., 2012; Manegold et al., 2013). Alternatively, cyclic peptides containing RGD motif could guide nanoparticulate delivery system, which incorporates anti-angiogenic cytotoxic agents such as doxorubicin, paclitaxel or combretastatin A4, to accumulate specifically in tumor vasculature with no overt systemic toxicity (Murphy et al., 2008; Ruoslahti et al., 2010; Wang et al., 2011). Volociximab, a chimeric humanized monoclonal antibody that selectively inhibits the αvβ1 integrin interaction with fibronectin, has been evaluated also in clinical trials for solid tumours such as renal cell carcinoma, recurrent ovarian cancer, advanced non–small-cell lung cancer and metastatic pancreatic cancer (Figlin et al., 2006; Evans et al., 2007; Jarvelainen et al., 2009; Vergote et al., 2009; Besse et al., 2013). Cadherins constitute a superfamily of molecules that mediate calcium-dependent cell–cell adhesions. The intracellular domains of cadherins directly bind to β-catenin and link with cytoskeletal components, providing the molecular basis for stable cell–cell adhesion (Zhang et al., 2010). Targeting cadherin signalling may also represent another way for tumor angiogenesis intervention. For example, ADH-1, a cyclic pentapeptide containing the cell adhesion recognition site (His-Ala-Val) required for N-cadherin adhesion, was shown to possess anti-angiogenic and anti-tumour activity (Blaschuk et al., 2005; Blaschuk, 2012). Similarly, monoclonal antibody directed against specific region of VE-cadherin was able to inhibit tumor angiogenesis and growth with no side effects on normal vasculature (Corada et al., 2002; May et al., 2005).
Inflammatory angiogenesis-targeted therapy
Targeting inflammatory angiogenesis, responsible for a substantial part of tumour vascularization initiated by infiltrating leukocytes, may be considered as another indirect anti-angiogenic strategy (Albini et al., 2005). Moreover, as mentioned before, tumour-infiltrating leukocytes contribute into malignant progression through production of many pro-inflammatory cytokines, chemokines and enzymes that can mostly induce angiogenic cascade (Balkwill et al., 2005). Such vital roles have been supported by the early observation that nonsteroidal anti-inflammatory drugs can inhibit tumour angiogenesis and, in turn, tumor progression (Albini et al., 2005). For example, ibuprofen was found to decrease tumor growth and metastatic potential in mice models through modulation of angiogenesis (Yao et al., 2005). Moreover, selective inhibitors of COX-2, an inducible enzyme that catalyses the production of prostanoids from arachidonic acid, were also shown to inhibit angiogenesis (Tsujii et al., 1998; Wei et al., 2004). The anti-angiogenic effect of COX-2 inhibitors may be contributed, in part, by decreasing the COX-2 metabolic product PGE2, the predominant PG in solid tumors known to stimulate cancer cells to produce pro-angiogenic factors such as VEGF and bFGF as well as many other factors belonging to CXC chemokines family (Strieter et al., 2004; Wang et al., 2006; Wang and Dubois, 2010). Members of the CXC chemokine family are heparin-binding proteins that possess disparate regulative roles in angiogenesis. For example, the ELR+ CXC chemokines, characterized by highly conserved three amino acid motifs (Glu-Leu-Arg; ‘ELR’ motif), are potent promoters of angiogenesis, whereas the IFN-inducible (ELR−) CXC chemokines are inhibitors of angiogenesis (Strieter et al., 2004). The use of repertaxin, originally designed to target the ELR+ CXC chemokine receptors CXCR1 and CXCR2 on neutrophils to prevent their migration to sites of inflammation, was found to inhibit tumor angiogenesis, thereby suppressing tumour progression in a genetic model of pancreatic ductal adenocarcinoma (Ijichi et al., 2011). It would be beneficial to explore other small-molecule CXCR2 antagonists that have already been developed for the treatment of inflammatory diseases in different preclinical models of cancer, especially inflammation-associated cancers (refer to Chapman et al., 2009 for a list of newly developed CXCR2 antagonists used in the treatment of inflammatory diseases of the lung).
Mechanisms of enhanced therapeutic efficacy
Dual targeting of tumor vasculature
Targeting different cell types of tumor micro-environment
Normalization of tumor vasculature
Chemosensitization of tumor cells
Interference with the repair of cytotoxic drug-induced damage and resistance mechanisms
Consequences of anti-angiogenic therapy with other anticancer therapy
Contrary to initial expectations, treatment with angiogenesis inhibitors was associated with unexpected toxicities. The toxicity profiles of those inhibitors reflect the systemic disturbance of growth factor signalling pathways that mediate their anti-angiogenic activity (Elice and Rodeghiero, 2010; 2012). In this context, disturbance of the tight endothelial cell-platelet interaction that maintains vascular integrity results in bleeding complications, gastrointestinal perforations, and disturbed wound and ulcer healing (Verheul and Pinedo, 2007). In general, the incidence of those adverse effects increases when anti-angiogenic agent is combined with chemotherapy. For example, bleeding complications have been observed in patients with colorectal cancer treated with chemotherapy in combination with bevacizumab (Kabbinavar et al., 2003; Giantonio et al., 2006). In non–small-cell lung cancer, some patients treated with bevacizumab in combination with carboplatin and paclitaxel experienced severe or fatal pulmonary haemorrhage (Johnson et al., 2004). Furthermore, a higher incidence of gastrointestinal perforation was observed in patients with colorectal cancer given bevacizumab in combination with chemotherapy compared with chemotherapy alone (Hurwitz et al., 2004). Similarly, thrombotic events have been observed in patients treated with angiogenesis inhibitors, especially when these agents are given in combination with chemotherapy (Verheul and Pinedo, 2007). Treatment of patients with cancer with angiogenesis inhibitors is frequently associated with hypertension, which may require the addition of regular anti-hypertensive agent (Izzedine et al., 2009).
Summary and future directions
Angiogenesis is a critical process that occurs pathologically in many malignancies due to changing balance of endogenous angiogenesis inducers and inhibitors, leading to the activation of nearby endothelial cells to form new vasculature. Consequently, angiogenesis can be targeted to restrict initiation, growth and progression of most of angiogenesis-dependent malignancies. Numerous angiogenic inhibitors have been identified, some of which are currently being investigated in clinical trials and some others were even approved for cancer therapies. These angiogenesis inhibitors were classified based on their target into two main classes: direct and indirect inhibitors. Indirect angiogenesis inhibitors can be further subclassified based on their interference mechanisms with the angiogenic cascade. A list of major categories and molecular targets for angiogenesis inhibitors is shown in Table 2.
Most angiogenesis inhibitors conferred clinical benefits mainly when combined with other chemotherapeutic/targeted therapies rather than being used as monotherapy. Unfortunately, many anti-angiogenic agents were shown to be associated with overt systemic toxicity as well as resistance emergence and disease recurrence. Drug resistance in anti-angiogenic therapy may result from a plethora of pro-angiogenic factors released by inappropriately functioning host cells in the tumor micro-environment as a compensatory mechanism. Therefore, the strategy of targeting endothelial cells alone may not be enough as explained in the previous texts, requiring the proposal of different rationales in which other cellular compartments of tumor micro-environment are targeted to attain proper anti-angiogenic and anti-tumor response. That highlights the importance of considering tumor micro-environment as a dynamic system, as depicted in Figure 1 in which interference with any of its components may be an approach to interfere with cancer hallmarks, including angiogenesis.
9.5.5 LUCITANIB a VEGFR/FGFR dual kinase inhibitor in Phase 2 trials
Spiro Substituted Compounds As Angiogenesis Inhibitors [US8163923]
2008-09-18
2012-04-24
A 4-(3-methoxypropoxy)-3-methylpyridinyl derivative of timoprazole that is used in the therapy of STOMACH ULCERS and ZOLLINGER-ELLISON SYNDROME. The drug inhibits H(+)-K(+)-EXCHANGING ATPASE which is found in GASTRIC PARIETAL CELLS.
For in advanced solid tumors.
Lucitanib (E-3810): Lucitanib, also known as E-3810, is a novel dual inhibitor targeting human vascular endothelial growth factor receptors (VEGFRs) and fibroblast growth factor receptors (FGFRs) with antiangiogenic activity. VEGFR/FGFR dual kinase inhibitor E-3810 inhibits VEGFR-1, -2, -3 and FGFR-1, -2 kinases in the nM range, which may result in the inhibition of tumor angiogenesis and tumor cell proliferation, and the induction of tumor cell death. Both VEGFRs and FGFRs belong to the family of receptor tyrosine kinases that may be upregulated in various tumor cell type
Lucitanib is an oral, potent inhibitor of the tyrosine kinase activity of fibroblast growth factor receptors 1 through 3 (FGFR1-3), vascular endothelial growth factor receptors 1 through 3 (VEGFR1-3) and platelet-derived growth factor receptors alpha and beta (PDGFR α-ß). We own exclusive development and commercial rights to lucitanib on a global basis, excluding China. Lucitanib rights to markets outside of the U.S. and Japan have been sublicensed to Les Laboratoires Servier (Servier). We are collaborating with Servier on the global clinical development of lucitanib.
11.1.1 Electrically-driven modulation of surface-grafted RGD peptides for .. cell adhesion
11.1.2 The metabolic state of cancer stem cells—a target for cancer therapy
11.1.3 Regulation of tissue morphogenesis by endothelial cell-derived signals
11.1.4 Novel approach to bis(indolyl)methanes. De novo synthesis of 1-hydroxyimino-methyl derivatives with anti-cancer properties
11.1.5 Synthesis and Biological Evaluation of New 1,3-Thiazolidine-4-one Derivatives of 2-(4-Isobutylphenyl)propionic Acid molecules
11.1.6 Targeting pyruvate kinase M2 contributes to radiosensitivity of NSCLC cells
11.1.7 The tyrosine kinase inhibitor nilotinib has antineoplastic activity in prostate cancer cells but up-regulates the ERK survival signal—Implications for targeted therapies
11.1.8 PAF and EZH2 Induce Wnt.β-Catenin Signaling Hyperactivation
11.1.9 PAF Makes It EZ(H2) for β-Catenin Transactivation
11.1.10 PI3K.AKT.mTOR pathway as a therapeutic target in ovarian cancer
11.1.11 Endogenous, hyperactive Rac3 controls proliferation of breast cancer cells by a p21-activated kinase-dependent pathway
11.1.12 Curcumin-could-reduce-the-monomer-of-ttr-with-tyr114cys-mutationvia autophagy in cell model of familial amyloid polyneuropathy.
11.1.1 Electrically-driven modulation of surface-grafted RGD peptides for .. cell adhesion
Reported herein is a switchable surface that relies on electrically-induced conformational changes within surface-grafted arginine–glycine–aspartate (RGD) oligopeptides as the means of modulating cell adhesion
Stimuli-responsive surfaces that are capable of modulating their biological properties in response to an external stimuli, including temperature,1,2 light,3 magnetic field4 and electrical potential,5–9 are of growing interest for a variety of biological and medical applications.10,11 Switchable surfaces that can be controlled on-demand are playing an increasingly important part in the development of highly sensitive biosensors,12–15novel drug delivery systems16–18 and functional microfluidic, bioanalysis, and bioseparation systems.19–22Additionally, dynamic, synthetic surfaces that can control the presentation of regulatory signals to a cell are expected to have a significant impact in the field of tissue engineering and regenerative medicine, and to provide unprecedented opportunities in fundamental studies of cell biology.23,24 The availability of sophisticated and functional switchable surfaces is expected to emulate more complex in vivo like extracellular environments, and provide a powerful means to probe and control the dynamic interactions between the cell and its external environments.
The majority of studies on stimuli-responsive surfaces reported to date either rely25–29 on controlling non-specific interactions (i.e., hydrophobic/hydrophilic and electrostatic) of the biomolecules with the active surface, or have focused30–32 on demonstrating modulation of specific biomolecular interactions using relatively simple biological systems (e.g. biotin–streptavidin) and conditions (i.e. water or buffer solutions). For example, Zareie et al.30 fabricated a mixed self-assembled monolayer (SAM) on gold comprising oligo(ethylene glycol) (OEG) thiol molecules and shorter disulfides carrying biotin end-groups that regulated the interaction between biotin and streptavidin in water. The OEG thiols were able to switch in response to a change in temperature below and above their lower critical solution temperature (LCST = 37 °C). At 23 °C the structure of the OEG molecules was fully extended hindering the shorter biotin disulfide components. On the contrary, at 45 °C the OEG backbone collapsed, thus allowing the specific interaction between the biotin molecule on the surface and the protein streptavidin in solution. In our previous work,7–9 electrically controlled switching has been applied to regulate the conformational changes of modified positively charged oligolysine peptides tethered to a gold surface, such that biotin moieties incorporated into the oligolysines could be reversibly exposed or concealed on demand, as a function of surface potential. Switchable SAMs used to control biomolecular interactions via an electrical stimulus are particularly appealing because of their fast response times, ease of creating multiple individually addressable switchable regions on the same surface, as well as low-drive voltage and electric fields, which are compatible with biological systems.33 Our previous reported electrically switchable surface was able to control directly the biomolecular interactions between biotin and neutravidin in phosphate buffer saline (PBS) solution.
However, switchable surfaces have been scarcely used, thus far, to control biomolecular interactions on more complex systems such as those involving modulation of cell responsiveness.34–37 Jonkheijm and co-workers35 have reported a cucurbit[8]uril-based SAM system to electrochemically control the release of cells. Charged end groups on SAM surfaces have been exploited to electrically control the early stages of bacterial cell adhesion37 and form patterned surfaces with two independent dynamic functions for inducing cell migration.36 In spite of these efforts, given cellular complexity and diversity, such studies are very limited in number, as are the opportunities to further understand and control the complex interplay of events and interactions occurring within living cells.
Herein, we report on a stimuli-responsive surface that relies on electrically-induced conformational changes within surface-grafted arginine–glycine–aspartate (RGD) oligopeptides as the means of modulating cell adhesion. RGD, which is present in most of the adhesive ECM proteins (e.g. fibronectin, vitronectin, laminin and collagen), is specific for integrin-mediated cell adhesion.38 The RGD modified electrode is used here to dynamically regulate the adhesion of immune macrophage cells. The stimuli-responsive surface is fabricated on a gold surface and comprises a mixed SAM consisting of two components (Fig. 1): (i) an oligopeptide containing a terminal cysteine for attachment to the gold surface, three lysine residues as the main switching unit, and a glycine–arginine–glycine–aspartate–serine (GRGDS) as the recognition motif for cell adhesion –C3K-GRGDS, and (ii) an ethylene glycol-terminated thiol (C11TEG) to space out the oligopeptides. Since the charged backbone of the oligopeptide can be potentially harnessed7–9 to induce its folding on the surface upon an application of an electrical potential, we reasoned that such conformational changes can be employed to selectively expose under open circuit (OC) conditions (bio-active state) or conceal under negative potential (bio-inactive state) the RGD to the cell and dynamically regulate cell adhesion.
Fig. 1 Schematic of the dynamic RDG oligopeptide SAM utilised for controlling specific cellular interactions. The electrically switchable SAM exposes the RGD peptide and supports cell adhesion under open circuit (OC) conditions (no applied potential), while …
Mixed SAMs of C3K-GRGDS : C11TEG were formed from a solution ratio of 1 : 40 and characterised by X-ray photoelectron spectroscopy (XPS) (Fig. S2, ESI†). XPS analysis confirmed the formation of the C3K-GRGDS:C11TEG mixed monolayer and displayed signals from S, N, C and O. The chemical state of the sulphur atom was probed using the XPS spectra of the S 2p emission (Fig. S2, ESI†). The S 2p spectrum (Fig. S2a, ESI†) consists of two doublet peaks, with one doublet peak at 162.0 eV (S 2p3/2) and 163.2 eV (S 2p1/2), indicating that the sulphur is chemisorbed on the gold surface.39 A second small doublet peak can be observed at 163.8 eV and 165.0 eV, which can be attributed to the S–H bond, indicating a small presence of unbound sulphur. No sulphur peaks above 166 eV were observed, indicating that no oxidised sulphur is present at the surface. The N 1s spectrum (Fig. S2b, ESI†) can be de-convoluted into two peaks, which support the presence of the peptide on the surface. The first peak centred at 400.5 eV is attributed to amino (NH2) and amide (CONH) moieties. The second peak centred at 402.8 eV is ascribed to protonated amino groups.40 Note that no nitrogen peak was observed for pure C11TEG SAMs. The C 1s spectrum (Fig. S2c, ESI†) can be de-convoluted into three peaks, which are attributed to five different binding environments. The peak at 285.0 eV is attributed to C–C bonds,41 while the peak at 286.7 eV corresponds to C 1s of the three binding environments of C–S, C–N and C–O.41 The third and smaller peak (288.6 eV) is assigned to the C 1s photoelectron of the carbonyl moiety, C O.41 The O 1s spectrum (Fig. S2d, ESI†) is de-convoluted into two different peaks, corresponding to two different binding environments, arising from the C–O (533.3 eV) and C O (532.0 eV) bonds.41 From integrating the area of the S 2 p and N 1s peaks and taking into consideration that the C3K-GRGDS oligopeptide consists of 15 N atoms and 1 S atom and C11TEG has no N and 1 S atom only, it was possible to infer that the ratio of C3K-GRGDS:C11TEG on the surface is 1 : 10 ± 2. The presence of C11TEG was utilised not only to ensure sufficient spatial freedom for molecular reorientation of the surface bound oligopeptide, but also to stop non-specific binding to the surface.
The C3K-GRGDS:C11TEG mixed SAMs were shown to support adhesion of immune macrophage cells as determined by cell counting42,43 (Fig. 2). When RAW 264.7 mouse macrophages were cultured on theC3K-GRGDS:C11TEG mixed SAM in supplemented Dulbecco’s Modified Eagle Medium (DMEM), the number of cells adhered to the surface increased with incubation time, reaching 1792 ± 157 cells per mm2after 24 hours. This is in contrast with the weak cell adhesion observed in two control surfaces, pureC11TEG SAMs and clean gold, in which the number of cells that adhere was 60% and 50% lower, respectively, after 24 hours (Fig. 2).
Fig. 2Microscopic images and density of adhered cells on C3K-GRGDS:C11TEG mixed SAM, pure C11TEG SAM and bare gold surfaces that were normalized against the density of cells adherent onto the C3K-GRGDS:C11TEG mixed SAM. The surfaces were cultured in RAW 264.7 mouse macrophage cells under OC conditions for 24 hours.
In order to demonstrate that the C3K-GRGDS:C11TEG mixed SAMs can support or resist cell adhesion on demand, the macrophage cells were cultured on the C3K-GRGDS:C11TEG mixed SAM in DMEM medium under OC conditions and applied negative potential (–0.4 V) for a period of 1 h. Note that DMEM contains a mixture of inorganic salts, amino acids, glucose and vitamins. On application of the applied potential of –0.4 V the number of adherent cells was 70% less compared to the C3K-GRGDS:C11TEGmixed SAMs under OC conditions, Fig. 3. Similar switching efficiencies have been observed in another oligopeptide system using different DMEM solutions.44 These findings suggest that the negative potential induces the conformational changes in the C3K moiety of C3K-GRGDS in the SAM which in turn leads to the RGD moiety being concealed and hence reducing the binding of the cells.
Fig. 3 Density of adhered cells on C3K-GRGDS:C11TEG, C11TEG, C6EG-GRGDS:C11TEG mixed SAMs that were normalized against the density of cells adherent onto the C3K-GRGDS:C11TEG mixed SAM. The surfaces were cultured in RAW 264.7 for 1 h under OC conditions or while applying –0.4 V.
Previous studies have shown that small conformational and orientational changes in proteins and peptides modulate the availability and potency of the active sites for cell surface receptors.45–47 Thus, in a similar manner, small changes in the conformation/orientation of the RGD peptide on the surface induced by application of an electrical potential are able to affect the binding activity of the peptide. Recently, we have conducted detailed theoretical8 and experimental9 studies aimed at understanding the switching mechanism of oligopeptide-based switchable surfaces, that similarly as in the case of the C3K-GRGDS:C11TEG mixed SAMs, use lysine residues to act as the switching unit. These previous studies unraveled that the surface-appended oligolysines undergo conformational changes between fully extended, partially extended and collapsed conformer structures in response to an applied positive potential, open circuit conditions and negative electrical potential, respectively. Thus, these previous findings allow us to propose that when a negative potential is applied to the GRGDS:C11TEG mixed SAM surface, the oligopeptide chain adopts a collapsed conformation on the surface and the RGD binding motif is partially embedded on the C11TEGmatrix, thus showing no bioactivity (“OFF” state).
In order to verify that the changes in adhesion upon application of a negative surface potential occur due to changes in the conformational orientation of the RGD instead of cell repulsion or cell damage due to the presence of an electrical potential, control mixed SAMs were also prepared using C11TEG and a peptide where the 3 lysine residues as the switching unit were replaced by 6 non-switchable ethylene glycol units –C6EG-GRGDS (Fig. S1, ESI†). Fig. 3 demonstrates that cells adhered in similar numbers to the C11TEGand C6EG-GRGDS:C11TEG mixed SAMs under OC conditions and an applied negative potential. These results provide strong evidence that control over cell adhesion using the C3K-GRGDS:C11TEG mixed SAM is due to a conformational behaviour of the lysine-containing oligopeptide that can either expose or conceal the RGD moiety.
Cell viability was checked following application of –0.4 V for 1 h by performing a trypan blue assay. Cells that were dead were stained blue due to a break down in membrane integrity. Incubation of the cells under a negative potential had negligible effect on cell viability, which was greater than 98%. Cyclic voltammetric studies (outlined in detail in the Fig. S3, ESI†) were also performed to demonstrate that no significant faradaic process occur over the potential range studied, and thus ions are not participating in redox reactions and consequently redox chemistry is not being significantly affected by application of the potential used. In agreement with other studies,35,36,48 we conclude that the electrical modulation of the surface neither affected cell viability nor induced any redox process in the medium that could have had an effect on cells.
We then addressed the question of whether the C3K-GRGDS:C11TEG surfaces could be switched between different cell adhesive states (cell-resistant and cell-adhesive states). To begin with, we investigated the switching from a cell-adhesive state to a cell-resistant state, and the possibility to detach the cells from the substrate upon the application of a negative potential. Cells were incubated in the C3K-GRGDS:C11TEGmixed SAMs for 1 h under OC conditions, thereby exposing the RGD moiety and allowing for cell attachment. This step was followed by the application of a potential of –0.4 V for 1 h in order to detach the cells from the surface, by concealing the RGD moieties. Cell counts showed no significant differences between the pre and post application of the –0.4 V, suggesting that the electrostatic force generated by the applied negative electrical potential might not be sufficient to disrupt the RGD–integrin interaction. These results were to a certain extent expected since adherent cells are able to withstand strong detachment forces due to the adhesion being mediated by multiple RGD–integrin bonds in parallel.49
In contrast, a reversal of the switching sequence demonstrated that our surfaces can be dynamically switched from a non-adhesive to cell-adhesive state. Cells were incubated in the C3K-GRGDS:C11TEG mixed SAMs for 1 h while holding the potential at –0.4 V for 1 h making the RGD peptide inaccessible for recognition by the corresponding integrin. As above, the number of adherent cells when a negative potential of –0.4 V was applied was 70% of the number that adhered to the C3K-GRGDS:C11TEG mixed SAMs under OC conditions, Fig. 4. The potential was then shifted to open circuit conditions for 1 h on those exposed to a potential of –0.4 V, which resulted in a significant increase in the number of cells as a result of the exposure of the RGD moiety to the cells (Fig. 4). These values were similar to those obtained for the samples that were only incubated for 1 hour under OC conditions (Fig. 4), indicating that the surfaces were highly effective at switching from a non-adhesive to cell-adhesive state.
Fig. 4Microscopic images and density of adhered cells on C3K-GRGDS:C11TEG mixed SAMs that were incubated with cells for 1 h while applying –0.4 V and subsequently in OC conditions for 1 h. The density was normalized against the density of cells adherent onto C3K-GRGDS:C11TEG mixed SAMs that were incubated with cells in OC conditions for 1 h.
In summary, an electrically switchable surface has been devised and fabricated that is capable of efficiently exposing and concealing the RGD cell adhesion motif and dynamically regulate the adhesion of immune macrophage cells. This study will no doubt be useful in developing more realistic dynamic extracellular matrix models and is certainly applicable in a wide variety of biological and medical applications. For instance, macrophage cell adhesion to surfaces plays a key role in mediating immune response to foreign materials.50 Thus, development of such dynamic in vitro model systems that can control macrophage cell adhesion on demand are likely to provide new opportunities to understand adhesion signaling in macrophages51 and develop effective approaches for prolonging the life-span of implantable medical devices and other biomaterials.52
11.1.2 The metabolic state of cancer stem cells—a target for cancer therapy
Bulk tumor cell populations rely on aerobic glycolysis.
Cancer stem cells are in a specific metabolic state.
Cancer stem cells in breast cancer, glioblastoma, and leukemia rely on oxidative phosphorylation of glucose.
In the 1920s Otto Warburg first described high glucose uptake, aerobic glycolysis, and high lactate production in tumors. Since then high glucose uptake has been utilized in the development of PET imaging for cancer. However, despite a deepened understanding of the molecular underpinnings of glucose metabolism in cancer, this fundamental difference between normal and malignant tissue has yet to be employed in targeted cancer therapy in the clinic. In this review, we highlight attempts in the recent literature to target cancer cell metabolism and elaborate on the challenges and controversies of these strategies in general and in the context of tumor cell heterogeneity in cancer.
11.1.3 Regulation of tissue morphogenesis by endothelial cell-derived signals
Endothelial cells lining blood vessels induce organ formation and other morphogenetic processes in the embryo.
Blood vessels are also an important source of paracrine (angiocrine) signals acting on other cell types in organ regeneration.
Vascular niches and endothelial cell-derived signals generate microenvironments for stem and progenitor cells.
Endothelial cells (ECs) form an extensive network of blood vessels that has numerous essential functions in the vertebrate body. In addition to their well-established role as a versatile transport network, blood vessels can induce organ formation or direct growth and differentiation processes by providing signals in a paracrine (angiocrine) fashion. Tissue repair also requires the local restoration of vasculature. ECs are emerging as important signaling centers that coordinate regeneration and help to prevent deregulated, disease-promoting processes. Vascular cells are also part of stem cell niches and have key roles in hematopoiesis, bone formation, and neurogenesis. Here, we review these newly identified roles of ECs in the regulation of organ morphogenesis, maintenance, and regeneration.
The examples provided in this review highlight the important roles of ECs in tissue development, patterning, homeostasis, and regeneration. The endothelium often takes a central position in these processes and there are many reasons why ECs are ideally positioned as the source of important instructive, angiocrine signals. The vascular transport network extends into every organ system and needs to be embedded in those tissues in a certain spacing or pattern, which places ECs in central and, therefore, strategic positions for the regulation of morphogenesis and organ homeostasis.
Given that ECs and other cell types frequently form functional units, such as kidney glomeruli, liver lobules, or lung alveoli, the assembly, differentiation, and function of the different cellular components needs to be tightly coordinated. In addition, because circulating blood cells extensively rely on the vascular conduit system and frequently interact with the endothelium, it is perhaps not surprising that ECs contribute to niche microenvironments. During tissue repair, proliferative cell expansion processes are sometimes temporally separated from cell differentiation and tissue patterning events. The latter has to involve the restoration of a fully functional vascular network so that ECs appear ideally suited as the source of molecular signals that can trigger or suppress processes in the surrounding tissue.
11.1.4 Novel approach to bis(indolyl)methanes. De novo synthesis of 1-hydroxyimino-methyl derivatives with anti-cancer properties
A versatile and broad range approach to previously unknown bis(indolyl)methane oximes based on two consecutive hetero Diels-Alder cycloaddition reactions of electrophilic conjugated nitrosoalkenes with indoles is disclosed. The cytotoxic properties and selectivity of some adducts against several human cancer cell lines pointing to a promising role in the development of anti-tumoural drugs, in particular for leukemia and lymphoma.
The one-pot synthetic strategy to bis(indolyl)methanes is outlined in Scheme 3. The starting a,a 0-dihalogenooximes 3 were efficiently prepared from the respective ketones by known procedures [58,61]. These compounds, in the presence of base, were converted, in situ, into the corresponding transient and reactive nitrosoalkenes 4, which were intercepted bya first molecule of the appropriate indole 5 originating the intermediate indole oximes 6. The initially formed tetrahydroxazines undergo ring-opening to the corresponding oximes, under the driving force of the energy gain on rearomatisation. Subsequent dehydro-halogenation of 6 produces nitrosoalkenes 7 which reacted with a second molecule of indole, producing the target bis(indolyl)methanes 8. The results obtained are summarised in Table 1.
The reaction yields may be considered generally good, taking into account that the synthetic process involves a sequence of reactions. On the other hand, no other products could be obtained, which indicates that the reactions were regioselective. The results have shown also that both alkyl and aryl oximes can be used in the synthesis of bis(indolyl)methanes. Starting from aryl oximes 3aef the expected (E) oximes 9 were obtained as single or major products (Entries 1e11) whereas alkyl oxime 3g reacted with indole to give the (Z)-oxime 10g as the major product (Entries 12e13). The stereochemistry assignment of oximes 9 and 10 was confirmed by analysis of the NOESY spectra of 9d, 9g, 10d and 10g. In the spectra of 10d and 10g, connectivity was observed between the hydroxyl proton and the phenyl protons and the methyl protons, respectively, whereas in the case of 9d and 9g no connectivity was observed. Moreover, oximes 9 and 10 are also characterized by 1H NMR spectra with different features. The chemical shift of the methylenic proton appears at higher value for (E)-oximes 9 (9b: δ 6.81 ppm; 9d: δ = 6.82 ppm; 9g: δ = 6.39 ppm) than for the corresponding (Z) oximes 10 (10b: δ = 5.74 ppm; 10d: δ = 5.77 ppm; 10g: δ = 5.41 ppm).
The synthesis of two isomeric oximes from the reaction of arylnitrosoethylenes with pyrrole and dipyrromethanes has been previously observed [62]. The process was rationalized considering the conjugate addition of the heterocycle to the nitrosoalkene, at the s-cis or s-trans conformation, followed by rearomatization of the pyrrole unit leading to (E)- and (Z)-oxime, respectively. Thus, the synthesis of the BIM oximes via 1,4-conjugate addition of indole to the nitrosoelkene cannot be ruled out.
The use of water as solvent in Diels- Alder reactions has been shown to be advantageous, not only in environmental terms but also inducing critical improvements in reaction times, yields and selectivity [51,63]. We observed that carrying out the synthesis of bis(indolyl)methanes in water using dichloromethane as co-solvent is a valuable alternative to the use of dichloromethane as the only solvent. Generally the yields were better or comparable to those obtained in dichloromethane and reaction time significantly shorter (the reaction time was reduced from 36 h to 3 h). Clearly the efficiency of the reaction, using H2O/CH2Cl2 system, amongst the nitrosoalkenes bearing halogenated aryl substituents increases in the order F > Cl > Br > H the order of electron withdrawing ability and consequently the order of the expected effectiveness for an inverse electron demand Diels-Alder reaction (entries 2, 5, 7 and 9). However, the isolated yields from the reaction carried out in CH2Cl2 do not reflect the expected reactivity, which can be explained considering differences in the efficiency of the purification process.
The cytotoxicity of compounds 9a, 9e and 9d was evaluated in different tumorl cell lines, namely HepG2 (hepatocellular carcinoma), MDA-MB-468 (human breast carcinoma), RAW 264.7 (murine leukemic monocyte macrophages), THP1 (human acute monocytic leukaemia), U937 (human leukaemic monocytic lymphoma) and EL4 cells (murine T-lymphoma). The compounds’ selectivity towards tumoural cells was assessed determining their cytotoxicity with respect to two non-tumoural derived cell lines S17 (murine bone marrow) and N9 cells (murine microglial). Results of the half maximal concentrations (IC50) are shown in Table 2 together with the toxicity of etoposide, a known antitumoural drug. Compound 9e was considerably less cytotoxic on tumor cell lines than the other two compounds, with IC50 values ranging from 35.7 (HepG2) to 124 mM (THP1) and was not selective. Compounds 9a and 9d, however, were considerably cytotoxic to all cells tested, with IC50 values ranging from 1.62 (THP1) to 23.9 mM (RAW) and from 10.7 (MDA) to 34.1 mM (U937), respectively. Compound 9a was particularly active against non-adherent cell lines with IC50 values ranging from 1.62 in THP1 to 1.65 mM in EL4.
Some conclusions regarding structure activity relationships can be redrawn based on the biological evaluation of these bis(indolyl)methanes. There is a dramatic difference in anticancer activitybetweenN-unsubstituted bis(indolyl)methanes 9a and the Nmethyl substituted derivative 9e, the latter characterized by high IC50 values. On the other hand, the significantly lower IC50 values observed for 9a for non-adherent cell lines in comparisonwith the ones obtained for 9d demonstrates that the presence of the bromo substituent leads to higher cytotoxic activity.
The observed high cytotoxicity of compound 9a against THP1, EL4 and U937 cell lines led us to extend the study to BIMs 9c, 9g and 10g (Table 3). Compound 9c, bearing a 4-fluorophenyl substituent, showed moderate anti-cancer activity which reinforces the observation that the 4-bromophenyl group is crucial to ensure low IC50 values. On the other hand, alkyl oximes 9g and 10g were even less cytotoxic against THP1, EL4 and U937 cell lines. None of these compounds were selective towards the tumor cell lines (selectivity index calculated for non-tumour cell line S17). In addition to having displayed higher toxicity towards the nontumor cell lines than all the studied compounds, compound 9a demonstrated the highest selectivity indexes: 9.86-14.2. Further studies using 9a as scaffold in the development of anti-tumoural drugs for leukaemia and lymphoma is worth pursuing since it presents lower IC50 and higher selectivity than etoposide.
Conclusions
The reliable preparation of a variety of unknown BIMs bearing different oxime substituents at the methylene bridge was presented. This strategy, supported on the robust and proved methodology of Diels-Alder cyclo addition reactions of electrophilic nitrosoalkenes with electron rich indoles, may pave the way for the synthesis of a vast library of new compounds.
Table 1 Preparation of bis(indolyl)methane oxime
Scheme 1. Selected biological active bis(indolyl)methanes.
Scheme 2. Common methods for BIMs’ preparation [27e44].
Scheme 3. Synthetic strategy towards BIM oximes.
Synthesis of a new bis(indolyl)methane that inhibits growth and induces apoptosis in human prostate cancer cells
The synthesis and the antiproliferative activity against the human breast MCF-7, SkBr3 and the prostate LNCaP cancer cell lines of a series of bis(indolyl)methane derivatives are reported. The synthesis of new compounds was first accomplished by the reaction of different indoles with trimethoxyacetophenone in the presence of catalytic amounts of hydrochloric acid. A second procedure involving the use of oxalic acid dihydrate [(CO2H)2·2H2O] and N-cetyl-N,N,N-trimethylammonium bromide in water was carried out and led to better yields. Compound 5b significantly reduced LNCaP prostate cancer cell viability in a dose-dependent manner, with an IC50 of 0.64 ± 0.09 μM. To determine whether the growth inhibition was associated with the induction of apoptosis, treated cells were stained using DAPI. LNCaP cells treated with 1 μM of 5b showed the morphological changes characteristic of apoptosis after 24 h of incubation.
11.1.5 Synthesis and Biological Evaluation of New 1,3-Thiazolidine-4-one Derivatives of 2-(4-Isobutylphenyl)propionic Acid molecules
New thiazolidine-4-one derivatives of 2-(4-isobutylphenyl)propionic acid (ibuprofen) have been synthesized as potential anti-inflammatory drugs. The structure of the new compounds was proved using spectral methods (FR-IR, 1H-NMR, 13C-NMR, MS). The in vitro antioxidant potential of the synthesized compounds was evaluated according to the total antioxidant activity, the DPPH and ABTS radical scavenging assays. Reactive oxygen species (ROS) and free radicals are considered to be involved in many pathological events like diabetes mellitus, neurodegenerative diseases, cancer, infections and more recently, in inflammation. It is known that overproduction of free radicals may initiate and amplify the inflammatory process via upregulation of genes involved in the production of proinflammatory cytokines and adhesion molecules. The chemical modulation of acyl hydrazones of ibuprofen 3a–l through cyclization to the corresponding thiazolidine-4-ones 4a–n led to increased antioxidant potential, as all thiazolidine-4-ones were more active than their parent acyl hydrazones and also ibuprofen. The most active compounds are the thiazolidine-4-ones 4e, m, which showed the highest DPPH radical scavenging ability, their activity being comparable with vitamin E.
In order to improve the anti-inflammatory effect and safety profile of representative NSAIDs, one research strategy is derivatization of the carboxylic acid group with various heterocyclic systems (oxazole, izoxazole, pyrazole, oxadiazole, thiazole, thiadiazole, triazole, etc.) [9,10]. In the past two decades there has been considerable interest in the role of reactive oxygen species (ROS) in inflammation [11]. ROS mediate the oxidative degradation of cellular components and alteration of protease/antiprotease balance with damage to the corresponding tissue. In the early stages of the inflammatory process, ROS exert their actions through activation of nuclear factors, such as NFkB or AP-1, that induce the synthesis of cytokines. In later stages, endothelial cells are activated due to the synergy between free radicals and cytokines, promoting the synthesis of inflammatory mediators and adhesion of molecules. In the last step free radicals react with different cellular components (trypsin, collagen, LDL, DNA, lipids) inducing the death of cells [12,13].
The thiazolidine-4-one moiety is a heterocycle that has received more attention in the last years due its important biological properties [14]. Many effects have been found, including anti-inflammatory and analgesic [15], antitubercular [16], antimicrobial and antifungal [17], antiviral, especially as anti-HIV agents [18], anticancer, antioxidants [19], anticonvulsants [20] and antidiabetic activity [21]. In the present study, some new derivatives of ibuprofen that contain thiazolidine-4-one scaffolds were synthesized in order to obtain compounds with double effect—antioxidant and anti-inflammatory properties. The structures of the compounds were assigned based on their spectral data (FT-IR, 1H-NMR, 13C-NMR, MS) and the compounds were screened for their in vitro antioxidant potential.
The 1,3-thiazolidine-4-one derivatives 4a–m were synthesized in several steps using the method summarized in Scheme 1 and Table 1. First 2-(4-isobutylphenyl)propionic acid (ibuprofen, 1) was reacted with thionyl chloride, followed by treatment with dry ethanol to get 2-(4-isobutylphenyl)propionic acid ethyl ester, which was turned in 2-(4-isobutylphenyl)propionic acid hydrazide (2) by reaction with 66% hydrazine hydrate [22]. The condensation of compound 2 with various aromatic aldehydes allowed the preparation of the corresponding hydrazone derivatives 3a–l in satisfactory yields. Finally, the hydrazone derivatives of ibuprofen upon reaction with mercaptoacetic acid led to the thiazolidine-4-one derivatives 4a–l in moderate to good yields. By reduction of compound 4g in presence of tin chloride and few drops of acetic acid in ethanol, the thiazolidine-4-one 4m was obtained in 90% yield. Acetylation of 4m with acetyl chloride gave thiazolidine-4-one 4n in moderate yield.
In the acyl hydrazone series most of the the tested compounds showed a radical scavenging ability comparable with ibuprofen (Table 4). The most active compounds were 3e and 3f which are about three times and two times more active than their parent compound, respectively. The scavenging ability of the acyl hydrazones was improved by cyclization to the corresponding thiazolidine-4-one derivatives, these compounds all being more active than ibuprofen, except for compound 4j which contains a CF3 group in the metaposition of phenyl ring (Table 5). The most active compounds were 4e and 4m which contain NO2 and NH2 groups in ortho and paraposition of the phenyl ring, respectively. For these compounds the radical scavenging ability (%) was 94.42 ± 0.43 and 94.88 ± 0.57, which means that the compounds are about 23 times more active than ibuprofen (4.15 ± 0.22). The activity of these compounds is comparable with that of vitamin E used as positive control. Important radical scavenging ability was also shown by compound 4b(81.31 ± 0.55), which contains a Cl group in the para position of the phenyl ring, the compound being 20 times more active than ibuprofen.
The acyl hydrazone derivatives showed an antioxidant activity comparable with ibuprofen. The most active compound in this series was 3h, with radical scavenging activity of 13.31 ± 0.81, which means that this compound is three times more active than ibuprofen (4.42 ± 0.18). In the thiazolidine-4-one series the most active compounds were 4b, 4e and 4k, which contain Cl(4), NO2(2) and CN(4), respectively, as substituents on the phenyl ring. These compounds, which showed a scavenging ability of around 50%, are 12 times more active than ibuprofen. In comparison with the corresponding acyl hydrazones 3b, 3e and 3k the thiazolidine-4-ones were 10 times (4b), seven times (4e) and 13 times (3k) more active. The improved antiradical activity of acyl hydrazones by cyclization to form thiazolidine-4-ones was also observed for compounds 3d, 3f and 3g. The most favorable influence was observed for acyl hydrazone 4g, which contains a NO2 in the para position of the phenyl ring. The corresponding thiazolidine-4-one (4g, 37.14 ± 1.10) is 22 times more active than 3g (1.67 ± 0.35). These data strongly support the favorable influence of the thiazolidine-4-one ring on the antioxidant potential of these compounds. The tested compounds were less active than vitamin E.
In this study new heterocyclic compounds that combine the thiazolidine-4-one structure with the arylpropionic acid one have been synthesized. The structure of the new compounds was proved using spectral methods (IR, 1H-NMR, 13C-NMR, MS). The compounds were evaluated for their antioxidant effects using in vitro assays: total antioxidant activity, DPPH and ABTS radical scavenging ability. All thiazolidin-4-one derivatives 4a–n showed improved antioxidant effects in comparison with the corresponding acyl hydrazones 3a–l and ibuprofen, the parent compound. The encouraging preliminary results illustrate the antioxidant potential of the synthesized compounds and motivate our next research focused on their anti-inflammatory effects in chronic and acute inflammation models.
11.1.6 Targeting pyruvate kinase M2 contributes to radiosensitivity of NSCLC cells
Aerobic glycolysis, a metabolic hallmark of cancer, is associated with radioresistance in non-small cell lung cancer (NSCLC). Pyruvate kinase M2 isoform (PKM2), a key regulator of glycolysis, is expressed exclusively in cancers. However, the impact of PKM2 silencing on the radiosensitivity of NSCLC has not been explored. Here, we show a plasmid of shRNA-PKM2 for expressing a short hairpin RNA targeting PKM2 (pshRNA-PKM2) and demonstrate that treatment with pshRNA-PKM2 effectively inhibits PKM2 expression in NSCLC cell lines and xenografts. Silencing of PKM2 expression enhanced ionizing radiation (IR)-induced apoptosis and autophagy in vitro and in vivo, accompanied by inhibiting AKT and PDK1 phosphorylation, but enhanced ERK and GSK3β phosphorylation. These results demonstrated that knockdown of PKM2 expression enhances the radiosensitivity of NSCLC cell lines and xenografts as well as may aid in the design of new therapies for the treatment of NSCLC.
Knockdown of PKM2 expression increases the sensitivity of NSCLC cells to radiotherapy in vitro
To examine PKM2 expressions levels in the normal lung epithelial cell and the NSCLC cell lines, we evaluated the expression levels of PKM2 in normal lung bronchial epithelial cell BEAS-2B and five NSCLC cell lines including A549, H460, H1299, H292, and H520 by Western blotting assays, and our results demonstrated that PKM2 expression was elevated in almost five NSCLC cell lines examined compared to autologous normal lung bronchial epithelial cell, although the expression levels fluctuated slightly depending on the different cell lines (Fig.1A). To test the role of PKM2 in the sensitivity of NSCLC to radiotherapy, we generated plasmids of pshRNA-PKM2 and control pshRNA-Con by inserting the DNA fragment for a pshRNA specifically targeting the PKM2 or control into the pGenesil2 vector. After demonstrating the authenticity, A549 and H460 cells were transfected with the plasmid for 48h and the levels of PKM2 expression were tested by Western blot assays. Obviously, transfection with control plasmid did not significantly modulate PKM2 expression; while transfection with pshRNA-PKM2 reduced the levels of PKM2 expression (Fig.1B and Appendix: Supplementary Fig.S1A). Quantitative analysis revealed that transfection with pshRNA-PKM2 significantly reduced PKM2 expressions as compared with that in the mock-treated and control pshRNA-Con plasmid-transfected cells, respectively (p<0.05, Fig.1C). Mock-treated and pshRNA-PKM2-trasnfected A549 and H460 cells were subjected to IR (4Gy), and 12 and 24h after IR, these cells, together with un-irradiated mock-treated, pshRNA-Con-transfected, and pshRNA-PKM2-trasnfected cells, were tested for cell viability by trypan blue staining. Knockdown of PKM2 reduced the percentage of A549 viable cells by 12.6–20% and IR treatment decreased the frequency of viable cells by 17.1–28.2%. However, the percentages of viable cells in the PKM2-silencing and irradiated cells were reduced by 27.7–48.7%, which were significantly lower than that in other groups (Fig.1D, p<0.05). Furthermore, it was consistent with the above results of A549 cells that knockdown of PKM2 significantly reduced the percentage of H460 viable cells (Appendix: Supplementary Fig.S1B). In addition, to further validate PKM2 silencing on their radiosensitivity,unirradiated control, mock-treated, and pshRNA-PKM2 transfected A549 cells were subjected to IR (0, 2, 4, 6, and 8Gy), and two weeks after IR, these cells were tested for the capacity for colony formation. The results showed that the numbers of colonies formed by pshRNA-PKM2 cells were significantly decreased compared with that of mock-treated and control cells; however, there was no significant change in mock-treated cells compared with control cells. These results suggested that pshRNA-PKM2 cells were more sensitive to IR than mock-treated and control cells (Fig.1E and F). Given that IR usually causes DNA double-strand breaks [28], we characterized the frequency of γ-H2AX nuclear foci positive cells by immunofluorescent assays. While IR treatment dramatically increased the frequency of γ-H2AX+ cells, the same dose of IR further significantly increased the percentages of γ-H2AX+ cells when combined with PKM2 silencing at 12 and 24h after IR, and there was a significant difference in γ-H2AX+ cells between these two groups at 12 and 24 h after IR (Fig. 1G and H, p < 0.05).
Fig. 1. The PKM2 expression levels in the normal lung epithelial cell and the NSCLC cell lines and knockdown of PKM2 expression enhance the radiosensitivity of A549 cells in vitro. The expression levels of PKM2 in normal lung bronchial epithelial cell BEAS-2B and five NSCLC cell lines including A549, H460, H1299, H292, and H520 were determined by Western blotting assay (A). A549 cells were transfected with pshRNA-PKM2 or pshRNA-Con plasmid for 48h, and the levels of PKM2 expression were determined by Western blot assays using a PKM2-specific antibody and β-actin as an internal control (B and C). Data are representative images or expressed as mean±SD of the relative levels of PKM2 to control β-actin in individual groups of cells from three separate experiments. # p
Knockdown of PKM2 enhances IR-induced apoptosis in NSCLC cells
Next, we tested the impact of PKM2-silencing on IR-induced cell death types. One day after IR, the apoptotic cells in the irradiatedmock-treated,pshRNA-PKM2-trasnfected cells, and one group of cells that had been pre-treated with 30μM Z-VAD for 1h prior to IR, together with mock-treated, unirradiated pshRNA-Contransfected, and pshRNA-PKM2-trasnfected groups of cells were characterized by TUNEL assays and/or FACS analysis (Fig.2A and C). In comparison with that in mock-treated and control plasmid transfected cells, the frequency of apoptotic cells in the PKM2 silencing or IR-treated cells increased moderately, while the percentages of apoptotic cells in the cells receiving combined treatment with IR and PKM2-silencing were significantly greater. However, the frequency of apoptotic cells in the Z-VAD-pretreated cells was partially reduced. Apparently, knockdown of PKM2 and IR induced apoptosis in NSCLC cells in vitro (Fig. 2B and D, and Appendix: Supplementary Fig.S1C).
Fig. 2. Knockdown of PKM2 expression enhances IR-induced apoptosis in A549 cells. A549 cells were transfected with, or without, pshRNA-Con or pshRNA-PKM2 for 48h and treated with, or without, Z-VAD for 1h. Subsequently, the cells were subjected to IR, and 24h later, the frequency of apoptotic cells was determined by TUNEL assays and FACS. (A and C) TUNEL and FACS analyses of apoptotic cells. (B and D) Quantitative analysis of the percentage of apoptotic cells. Data are representative images or expressed as mean%±SD of individual groups of cells from three independent experiments. * p
Knockdown of PKM2 enhances IR-induced autophagy in NSCLC cells
The cell autophagy is characterized by the formation of numerous autophagic vacuoles, autophagosome, in the cytoplasm and elevated levels of the microtubule-associated protein 1 light chain 3 (LC3)-II [29]. To test the impact of PKM2 silencing on IR-induced autophagy, the presence of autophagosome in mock-treated, pshRNACon-transfected, pshRNA-PKM2-transfected, IR-treated alone, IR + pshRNA-PKM2-transfected, and 1 mM 3-MA-pretreated IR + pshRNA-PKM2-transfected cells was characterized by electronic microphotography (EM). Intriguingly and importantly, numerous autophagosomes were detected in the IR + pshRNAPKM2-transfected cells, and only a few were detected in the sensitivity of the NSCLC cells to radiotherapy in vitro. It was noted that pshRNA-Con had almost no effect on A549 cells, therefore, some subsequently experiments did not set this group.
Fig. 3. Knockdown of PKM2 and IR induce A549 cell autophagy. A549 cells were transfected with, or without, pshRNA-Con or pshRNA-PKM2 for 48h and treated with, or without, 3-MA for 1h. Subsequently, the cells were subjected to IR, and 2h later, the presence of autophagic vacuoles and autolysosomes in A549 cells was determined by EM and the relative levels of LC3-I, LC3-II, AKT, ERK1/2, and control β-actin expression and AKT, ERK1/2, GSK3β, PDK1 phosphorylation were determined by Western blot assays using specific antibodies. Data are representative images and expressed as mean values of the relative levels of target protein to control in individual groups of cells from three separate experiments. The relative levels of target protein to control in mock-treated cells were designated as 1. (A) EM analysis of autophagic vacuoles and autophagosomes. Black arrows point to autophagic vacuoles and autophagosomes in the cytoplasma of A549 cells. (B) Western blot analysis of LC3-I and LC3-II expression. The values indicate the ratios of the relative levels of LC3-II to LC3-I in individual groups. (C) Western blotting analysis of individual signal events. The values indicate the relative levels of target protein to control β-actin in individual groups of cell
Fig. 4. The impact of 3-MA or/and V-ZAD on cell viability, colony formation, apoptosis and autophagy in A549 cells. A549 cells were transfected with, or without, pshRNACon or pshRNA-PKM2 for 48h and pre-treated with, or without, 3-MA or V-ZAD for 1h, respectively. Subsequently, the cells were subjected to IR. Twenty-four hours later and two weeks, the viability, apoptosis, and colony formation were determined. Two hours after treatment, autophagy and the relative levels of LC3-I and LC3-II expression in different groups of cells were determined. Data are representative images and expressed as mean%±SD of individual groups of cells from three separate experiments. (A) The percentages of viable cells. (B) The capacity of cell colony formation. (C) Quantitative analysis of apoptotic cells. (D) Western blot analysis of LC3-I and LC3-II expression. The values indicate the ratios of LC3-II to LC3-I in individual groups of cells. * p
Fig. 5. Treatment with pshRNA-PKM2 enhances the IR-inhibited growth of implanted tumors in mice. The nude mice were inoculated with A549 cells and when the tumor grew at 50mm3 in one dimension, the mice were randomized and treated with vehicle (PS), plasmid of pshRNA-Con or pshRNA-PKM2 alone or IR (4Gy×7f) alone or in combination with pshRNA-PKM2 and IR, respectively. The body weights and tumor growths of individual mice were monitored longitudinally. At the end of the in vivo experiment, the tumor tissues were dissected out and the frequency of apoptotic cells, the presence of autophagosomes and the expression of PKM2 were determined by TUNEL, EM and immunohistochemistry, respectively. Data are representative images or expressed as mean±SD of individual groups of mice (n=6 per group). (A) The body weights of mice. (B and C) The tumor growth curve of implanted tumors and the log-transformed tumor growth curve of implanted tumors in mice. (D) Quantitative analysis of the frequency of apoptotic cells.(E) EM analysis of autophagy. (F)The expression of PKM2.(G) Quantitative analysis of PKM2 expression.The cells with brown cytoplasma were considered as positive anti-PKM2 staining and the percentage of PKM2-positive cells was obtained by dividing the numbers of the PKM2-positive cells by the total number of cancer cells in the same field.
11.1.7 The tyrosine kinase inhibitor nilotinib has antineoplastic activity in prostate cancer cells but up-regulates the ERK survival signal—Implications for targeted therapies
Background: Novel therapeutic options beyond hormone ablation and chemotherapy are urgently needed for patients with advanced prostate cancer. Tyrosine kinase inhibitors (TKIs) are an attractive option as advanced prostate cancers show a highly altered phosphotyrosine proteome. However, despite favorable initial clinical results, the combination of the TKI dasatinib with docetaxel did not result in improved patient survival for reasons that are not known in detail. Methods: The National Cancer Institute-Approved Oncology Drug Set II was used in a phenotypic drug screen to identify novel compounds with antineoplastic activity in prostate cancer cells. Validation experiments were carried out in vitro and in vivo. Results: We identified the TKI nilotinib as a novel compound with antineoplastic activity in hormone-refractory prostate cancer cells. However, further analyses revealed that treatment with nilotinib was associated with a significant up-regulation of the phospho-extracellular-signal-regulated kinases (ERK) survival signal. ERK blockade alone led to a significant antitumoral effect and enhanced the cytotoxicity of nilotinib when used in combination. Conclusions: Our findings underscore that TKIs, such as nilotinib, have antitumoral activity in prostate cancer cells but that survival signals, such as ERK up-regulation, may mitigate their effectiveness. ERK blockade alone or in combination with TKIs may represent a promising therapeutic strategy in advanced prostate cancer.
Identification of nilotinib as a novel antineoplastic compound in prostate cancer cells
Using the NCI-Approved Oncology Drug Panel II for a phenotypic drug screen of normal prostate epithelial cells and prostate cancer cell lines (Fig. 1) [7], we identified the TKI nilotinib as a positive hit in hormone-refractory DU-145 prostate cancer cells.
Fig. 1. Discovery of nilotinib as a novel antineoplastic agent in prostate cancer cells using a phenotypic drug screen. Overview of the drug screen procedure (see text for details).
Results were confirmed using annexin V staining, which showed a significant induction of apoptosis beginning at 24 hours (Fig. 2A). The IC50 of nilotinib against DU-145 cells was determined at 10 μM using an MTT cell viability assay (Fig. 2B). Immunoblot experiments confirmed an induction of apoptosis using PARP cleavage in DU-145 cells and in hormonerefractory PC-3 prostate cancer cells at this drug concentration (Fig. 2C). An onset of apoptosis at 24 hours was likewise confirmed using PARP cleavage at a nilotinib concentration of 10 μM(Fig. 2D). PWR-1E prostate epithelial cells and hormone-sensitive prostate LNCaP prostate cancer cells were not found to undergo enhanced apoptosis when treated with nilotinib (not shown).
Fig. 2. Antitumoral effects of nilotinib in prostate cancer cells: (A) flow cytometric analysis of DU-145 prostate cancer cells for annexin V to detect apoptotic cells after treatment with 10 μM of nilotinib for the indicated intervals; (B) cell viability (MTT) assay to determine the IC50 of nilotinib in DU-145 cells (24-h treatment); (C and D) immunoblot analysis of DU-145 and PC-3 prostate cancer cells for PARP cleavage (arrow) at nilotinib concentrations and time intervals as indicated. GAPDH is shown for protein loading; and (E) colony growth assay of DU-145 cells after drug treatment and washout as shown. Cells grown in 60-mm dishes were stained with crystal violet to visualize viable cells at the time points indicated. (Color version of figure is available online.
To further confirm the effect of nilotinib on prostate cancer cell growth, we performed a colony growth assay in which DU-145 cells were treated with nilotinib for 72 hours followed by a washout of the drug and continued culture for additional 9 days (Fig. 2E). We found that nilotinib induced significant cytotoxicity after 72 hours and that a minor regrowth of cancer cells did not occur until 6 to 9 days after the washout, which is comparable to other TKIs [8]. Next, we sought to identify the targets of nilotinib in DU-145 prostate cancer cells. Overall, 5 well-established targets, including ABL1, KIT, PDGFRA, DDR1, and NQO2, were analyzed for their role in the drug response. We found that protein expression of 3 of these targets (ABL1, KIT, and PDGFRA) was not detectable in DU-145 cells and that small interfering RNA–mediated knockdown of the remaining 2 targets, DDR1 and NQO2, did not result in apoptosis (not shown). Collectively, these results show a significant antitumoral activity of nilotinib in prostate cancer cells. However, this effect was associated with a relatively high IC50 and was independent of known nilotinib targets.
Nilotinib up-regulates the ERK survival signal in prostate cancer cells
To further investigate why relatively high concentrations of nilotinib were required to induce cytotoxicity, we analyzed 40,6-diamidino-2-phenylindole–stained DU-145 cells treated with 10 μM of nilotinib for 24 hours using fluorescence microscopy (Fig. 3A).
Fig. 3. Nilotinib up-regulates the ERK survival signal in prostate cancer cells. (A) Fluorescence microscopic analysis of DAPI-stained DU-145 cells. (B and C) Immunoblot analyses of DU-145 cells (B) or DU-145 cells in comparison with LNCaP and PC-3 cells (C) treated with nilotinib for the expression of phospho-ERK1/2 T202/Y204 and total ERK. Immunoblot for GAPDH is shown as a loading control. (D) Immunohistochemical staining of xenografted DU-145 cells after 21 days of treatment with 75 mg/kg/d of nilotinib for phospho-ERK1/2 T202/Y204 expression. It can be noted that tumors explanted from vehicle-treated mice showed mostly positivity at the tumor periphery, whereas tumors explanted from nilotinib-treated mice showed a more evenly distributed phospho-ERK immunostaining (left panels). Quantification of phospho-ERK–positive DU-145 xenografts explanted after 21 days of treatment. Mean and standard errors of positive cells per high-power field (HPF; [1]40) from at least 3 tumors are given (right panel). (E) Immunoblot analysis of DU-145 cells treated with U0126 alone or in combination with nilotinib shows abrogation of phospho-ERK1/2 T202/Y204 expression by U0126. (F) Quantification of viable cells compared with that of controls using the MTT assay after treatment with U0126 (10 μM) or nilotinib (10 μM) or both and after either pretreatment (24 h) or simultaneous treatment (72 h). DAPI ¼ 40,6-diamidino-2-phenylindole. (Color version of figure is available online.)
We found that, despite the presence of apoptotic cells, there was also a population of actively dividing tumor cells in the presence of nilotinib as well as a population of viable but multinucleated cells (Fig. 3A). We interpreted these results as evidence that a subset of tumor cells has the ability to resist TKI treatment. To reconcile these results, we analyzed the activation of ERK1/2, which is known to function as a prosurvival signal in TKI-treated tumor cells [9,10]. We detected a robust overexpression of phospho-ERK1/2 T202/Y204 in nilotinib-treated DU-145 cells (Fig. 3B). An up-regulation of phospho-ERK1/2 T202/Y204 was also detectable in nilotinib-treated LNCaP cells, albeit at a lower level, and was not found in PC-3 cells (Fig. 3C). To further corroborate the evidence of phospho-ERK upregulation in vivo, we analyzed explanted DU-145 xenografts from a representative experiment in which nilotinib was used at a 75-mg/kg/d concentration. This initial dosage was based on published animal experiments [11] but yielded no or incomplete tumor control in our experiment (data not shown).
In vivo antitumoral activity of nilotinib and ERK blockade
Our results raised 2 important questions First, can a higher dose of nilotinib induce improved tumor control, and second, is a combination of nilotinib with the MEK inhibitor U0126 to block ERK activity superior to nilotinib alone?
Fig. 4. In vivo antitumoral activity of nilotinib and ERK blockade in prostate cancer cells: (A) tumor growth curves of DU-145 xenografts in NMRI-nude mice and (B) analysis of tumor volumes on day 21. Asterisks indicate statistical significance (**P r 0.01 and ***P r 0.001). (Color version of figure is available online.)
11.1.8 PAF and EZH2 Induce Wnt.β-Catenin Signaling Hyperactivation
Fine-control of Wnt signaling is essential for various cellular and developmental decision making processes. However, deregulation of Wnt signaling leads to pathological consequences including cancer. Here, we identify a novel function of PAF, a component of translesion DNA synthesis, in modulating Wnt signaling. PAF is specifically overexpressed in colon cancer cells and intestinal stem cells, and required for colon cancer cell proliferation. In Xenopus laevis, ventrovegetal expression of PAF hyperactivates Wnt signaling, developing secondary axis with β-catenin target gene upregulation. Upon Wnt signaling activation, PAF is dissociated from PCNA, and directly binds to β-catenin. Then, PAF recruits EZH2 to β-catenin transcriptional complex, and specifically enhances Wnt target gene transactivation, independently of EZH2’s methyltransferase activity. In mouse, conditional expression of PAF induces intestinal neoplasia via Wnt signaling hyperactivation. Our studies reveal an unexpected role of PAF in regulating Wnt signaling, and propose a novel regulatory mechanism of Wnt signaling during tumorigenesis. Fine-control of Wnt signaling is essential for various cellular and developmental decision making processes. However, deregulation of Wnt signaling leads to pathological consequences including cancer. Here, we identify a novel function of PAF, a component of translesion DNA synthesis, in modulating Wnt signaling. PAF is specifically overexpressed in colon cancer cells and intestinal stem cells, and required for colon cancer cell proliferation. In Xenopus laevis, ventrovegetal expression of PAF hyperactivates Wnt signaling, developing secondary axis with β-catenin target gene upregulation. Upon Wnt signaling activation, PAF is dissociated from PCNA, and directly binds to β-catenin. Then, PAF recruits EZH2 to β-catenin transcriptional complex, and specifically enhances Wnt target gene transactivation, independently of EZH2’s methyltransferase activity. In mouse, conditional expression of PAF induces intestinal neoplasia via Wnt signaling hyperactivation. Our studies reveal an unexpected role of PAF in regulating Wnt signaling, and propose a novel regulatory mechanism of Wnt signaling during tumorigenesis.
Keywords: Wnt, β-catenin, PAF, KIAA0101, EZH2
Strict regulation of stem cell proliferation and differentiation is required for mammalian tissue homeostasis, and its repair in the setting of tissue damage. These processes are precisely orchestrated by various developmental signaling pathways, with dysregulation contributing to disease and genetic disorders, including cancer (Beachy et al., 2004). Cancer is initiated by the inactivation of tumor suppressor genes and activation of oncogenes. For instance, colon cancer cells harbor genetic mutations in Wnt/β-catenin pathway constituents such as adenomatous polyposis coli (APC), Axin, and β-catenin (Polakis, 2007). In mouse models, inactivation of APC or activation of β-catenin results in the development of intestinal hyperplasia and adenocarcinoma (Moser et al., 1990), indicating that hyperactivation of Wnt signaling promotes intestinal tumorigenesis.
In canonical Wnt signaling, Wnt ligand induces stabilization of β-catenin protein via inhibition of the protein destruction complex (glycogen synthase kinase 3, APC, casein kinase I, and Axin). Then, activated β-catenin is translocated into the nucleus and binds to its nuclear interacting partners, TCF/LEF. Finally, β-catenin-TCF/LEF transactivates the expression of its target genes (Clevers and Nusse, 2012).
Although various Wnt/β-catenin modulators have been identified (Wnt homepage; wnt.stanford.edu), the pathological relevance of these modulators to tumorigenesis remains elusive. Also, many reports have suggested that mutation-driven Wnt signaling activation can be enhanced further (Goentoro and Kirschner, 2009; He et al., 2005; Suzuki et al., 2004; Vermeulen et al., 2010), which implies the presence of an additional layer of Wnt-signaling regulation in cancer beyond genetic mutations in APC or β-catenin. Here, we unraveled a novel function of the DNA repair gene, PAF (PCNA-associated factor) /KIAA0101). PAF was shown to be involved in translesion DNA synthesis (TLS), an error-prone DNA repair process that permits DNA replication machinery to replicate DNA lesions with specialized translesion DNA polymerase (Emanuele et al., 2011; Povlsen et al., 2012; Sale et al., 2012). Our comprehensive approaches uncover that cancer-specifically expressed PAF hyperactivates Wnt/β-catenin signaling and induces intestinal tumorigenesis.
Mitogenic role of PAF via Wnt signaling
To identify colon cancer-specific Wnt signaling regulators, we analyzed multiple sets of human colon cancer tissue samples using the publicly available database (www.oncomine.org), and selected genes that are highly expressed in colon cancer cells (fold change > 2; P < 0.0001; top 10% ranked). Among several genes, we investigated the biological role of PAF, based on its significant overexpression in human colon adenocarcinoma with correlated expression of Axin2, a well-established specific target gene of β-catenin (Jho et al., 2002; Lustig et al., 2002)(Figure 1A). To validate our in silico analysis, we performed immunostaining of colon cancer tissue microarray, and confirmed that PAF was highly expressed in colon cancer cells, whereas its expression was barely detectable in normal intestine (Figure 1B). Consistently, PAF was strongly expressed and mainly localized in the nucleus of colon cancer cell lines (Figure 1C). Additionally, we found that PAF was not expressed in non-transformed cells such as NIH3T3, mouse embryonic fibroblasts, and mammary epithelial cells (data not shown). Next, to assess the relevance of PAF upregulation in colon cancer cell proliferation, we depleted endogenous PAF using short hairpin RNAs (shRNAs) in these cell lines. Intriguingly, PAF knockdown (sh-PAF) inhibited colon cancer cell proliferation (Figures 1D and 1E). Given that PAF was shown to interact with PCNA via PIP box (Yu et al., 2001), we also examined whether PAF-PCNA interaction is required for mitogenic effects of PAF. In reconstitution experiments, sh-PAF-induced cell growth inhibition was rescued by ectopic expression of both shRNA non-targetable wild-type PAF (nt-PAF) and PIP mutant PAF (mutPIP-PAF) (Figure 1F), indicating that the PAF-PCNA interaction is not necessary for PAF-mediated colon cancer cell proliferation. Interestingly, PAF knockdown downregulated cell proliferation–related genes (Cyclin D1 and c-Myc) (Figure 1G). Given that Cyclin D1 and c-Myc are β-catenin direct target genes (He et al., 1998; Tetsu and McCormick, 1999), PAF likely participates in regulating Wnt/β-catenin signaling. Interestingly, PAF depletion-induced downregulation of Cyclin D1 andc-Myc was only observed in SW620 colon cancer cells, but not in Panc-1 and MDA-MB-231 cells (Figure 1G), indicating the specific effects of PAF on Cyclin D1 and c-Myc expression in colon cancer cells. We also assessed the effects of PAF knockdown on Axin2. Indeed, PAF knockdown suppressed Axin2transcription in colon cancer cells (Figure 1H). Moreover, as nt-PAF did, β-catenin ectopic expression reverted sh-PAF–induced cell growth arrest (Figure 1I), implying that PAF might be functionally associated with Wnt/β-catenin. We also examined whether other mitogenic signaling pathways mediate PAF’s mitogenic role. Of note, except HT29, other colon cancer cell lines (SW620, HCT116, HCC2998, and HCT15) harbor oncogenic mutations in K-Ras gene. Nonetheless, PAF depletion induced the suppression of cell growth on all five colon cancer cells (Figure 1D), indicating that PAF’s mitogenic function is independent of Ras/MAPK signaling activation. Additionally, overexpression of wild-type Akt or constitutively active form of Akt (myristoylated form of Akt [Myr-Akt]) did not rescue sh-PAF-induced inhibition of cell proliferation (Figure 1I). Moreover, β-catenin activation did not revert cell proliferation suppression resulted from MAPK or PI3K inhibition (Figure 1J), indicating that β-catenin-mediated mitogenic function is independent of MAPK and PI3K signaling pathways. These results suggest that PAF contributes to colon cancer cell proliferation, possibly via Wnt/β-catenin signaling.
Figure 1Mitogenic role of PAF in colon cancer cells
PAF positively modulates Wnt signaling
Given that many cancers develop as a result of deregulation of developmental signalings (Beachy et al., 2004), analyzing PAF expression during development may provide insights into the mechanisms of PAF-mediated signaling regulation. Whole mount immunostaining of mouse embryos, showed that PAF was specifically enriched in the apical ectodermal ridge (AER) of the limb bud, midbrain, hindbrain, and somites (Figure 2A and data not shown). During limb development, AER induction is specifically coordinated by active Wnt signaling (Figure 2B)(Kengaku et al., 1998). Using, Axin2-LacZ, a β-catenin reporter (Lustig et al., 2002), mouse embryos, we confirmed the specific activation of Wnt signaling in AER (Figure 2C). Intriguingly, Wnt signaling activity as exhibited in the AER, overlapped with the pattern of PAF expression (Figures 2A and 2C). Given that (1) Wnt signaling is deregulated in most colon cancer, (2) PAF is highly overexpressed in colon cancer cells, (3) PAF is required for colon cancer cell proliferation (Figure 1D), and (4) PAF is enriched in AER where Wnt signaling is active (Figure 2A), we hypothesized that PAF modulates the Wnt signaling pathway. To test this, we first examined the impact of PAF on β-catenin transcriptional activity using TOPFLASH reporter assays. In HeLa cells, PAF knockdown decreased β-catenin reporter activation by 6-bromoindirubin-3′-oxime, a GSK3 inhibitor (Figure 2D). Similarly, Wnt3A-induced transcriptional activation of Axin2 was also inhibited by PAF depletion (Figure 2E). These data suggest that PAF might be required for Wnt target gene expression.
To gain better insight of PAF’s role in Wnt signaling regulation, we utilized Xenopus laevis embryos for axis duplication assays (Funayama et al., 1995), as previously performed (Park et al., 2009). Because of Wnt signaling’s pivotal role in vertebrate anterior-posterior axis development, the effects of Xenopus PAF (xPAF) on Wnt signaling can be monitored and quantified on the basis of secondary axis formation following injection of in vitro transcribed mRNAs. xβ-catenin mRNA, titrated to a subphenotypic level when expressed in isolation, was co-injected with xPAF mRNA into ventrovegetal blastomeres. Unlike the controls (β-catenin and β-galactosidase mRNA), the experimental group (β-catenin and xPAF mRNA) displayed axis-duplications (Figures 2F-H). Of note, the ventrovegetal injection of xPAF mRNA alone failed to induce secondary axes (data not shown), showing that PAF hyperactivates Wnt/β-catenin signaling only in the presence of active β-catenin. Consistent with the results of axis duplication assays, qRT-PCR assays showed that xPAF expression upregulated expression of Siamois and Xnr3, β-catenin targets in frogs (Figure 2I). Furthermore, we examined the specificity of PAF on Wnt/β-catenin signaling activity, using various luciferase assays. Ectopic expression of PAF hyperactivates Wnt3A or LiCl, a GSK3 inhibitor, -induced activation of β-catenin target gene reporter activity (MegaTOPFLASH, Siamois, c-Myc, and Cyclin D1). Of note, BMP/Smad pathway also plays an essential role in the developing limb AER (Ahn et al., 2001). However, PAF knockdown or overexpression did not affect BMP/Smad or FoxO signalings, respectively, (Figure 2J) indicating the specificity of PAF in regulating Wnt signaling. These results suggest that PAF positively modulates Wnt/β-catenin signaling in vitro and in vivo.
Next, we investigated the molecular mechanism underlying PAF hyperactivation of Wnt signaling. Given that stabilization of β-catenin protein is a key process in transducing Wnt signaling, we asked whether PAF affects β-catenin protein level. However, we found that the level of β-catenin protein was not altered by PAF knockdown or overexpression (Figures 2E and and3A),3A), leading us to test whether PAF controls the β-catenin/TCF transcriptional complex activity. Owing to the nuclear specific localization of PAF in colon cancer cells (Figure 1C), we tested whether PAF interacts with β-catenin transcriptional complex. Using a glutathione S-transferase (GST) pull-down assay, we found that PAF bound to β-catenin and TCF proteins (Figure 3B). Also, endogenous PAF interacted with β-catenin and TCF3 in SW620 cells that display constitutive hyperactivation of Wnt signaling by APC mutation (Figure 3C). Moreover, binding domain mapping assays showed that the Armadillo repeat domain of β-catenin was essential for its interaction with PAF (Figure 3D). Although PAF is a cell cycle-regulated anaphase-promoting complex substrate (Emanuele et al., 2011), PAF-β-catenin interaction was not affected (Figure S1). These data suggest that PAF directly binds to β-catenin transcriptional complex and this interaction is independent of cell cycle. Next, due to interaction of PAF with β-catenin and TCF, we tested whether PAF is also associated with Wnt/β-catenin target genes. First, we analyzed the subnuclear localization of PAF by chromatin fractionation. We found that PAF was only detected in the chromatin fraction of HCT116 cells (Figure 3E). Additionally, chromatin immunoprecipitation (ChIP) assays showed that both ectopically expressed and endogenous PAF occupied the TCF-binding element (TBE)-containing proximal promoter of the β-catenin targets (c-Myc and Cyclin D1) in HCT116 cells (Figures 3F and 3G). These data show that PAF is specifically associated with the promoters of Wnt/β-catenin target genes.
Figure 3PAF-EZH2-β-catenin transcriptional complex at target gene promoters
In intestine, Wnt/β-catenin signaling constitutively activates intestinal stem cells (ISCs) to give rise to progenitor cells, which replenishes intestinal epithelium (Figure 3H). Given the involvement of PAF on Wnt/β-catenin signaling regulation (Figure 2), we analyzed the spatial expression of PAF in intestinal epithelium. Immunostaining showed that PAF was specifically expressed in B lymphoma Mo-MLV insertion region 1 homolog (Bmi1) positive intestinal stem cells (ISCs)(Figures 3I and 3J). Bmi1 and its associated components in Polycomb-repressive complex 1 (PRC1) and 2 (PRC2) are shown to epigenetically regulate gene expression (Sparmann and van Lohuizen, 2006). Due to (1) specific association of PAF with TBEs of β-catenin target promoters (Figures 3F and 3G) and (2) co-localization with Bmi1 positive ISCs (Figure 3J), we asked whether PAF is associated with components of PRC1 and PRC2, using co-immunoprecipitation (co-IP) assays. Intriguingly, PAF interacted with both Bmi1 and enhancer of zeste homolog 2 (EZH2) in SW620 cells (Figure 3K), which led us to test whether either Bmi1 or EZH2 is functionally associated with PAF-mediated Wnt signaling hyperactivation. To do this, we assessed the effects of Bmi1 and EZH2 on β-catenin transcriptional activity, using β-catenin reporter assays. We observed that ectopic expression of EZH2 upregulated β-catenin transcriptional activity, but Bmi1 overexpression did not (data not shown), implying that EZH2 might be associated with Wnt signaling activation. Binding domain mapping analysis showed that EZH2 bound to PAF via the middle region of EZH2 including the CXC cysteine-rich domain (Figure 3L). In conjunction with the Bmi1-containing PRC1, EZH2-containing PRC2 catalyzes histone H3 lysine 27 trimethylation (H3K27me3) via histone methyltransferase domain. Despite the crucial role of EZH2 in H3K27me3-meidated gene regulation, we found that other core components of PRC2, EED, and Suz12 were not associated with PAF (Figure 3K). Moreover, although EZH2 overexpression in cancer induces PRC4 formation in association with the NAD+-dependent histone deacetylase Sirt1 (Kuzmichev et al., 2005), the PAF-EZH2 complex did not contain Sirt1 (Figure 3K). These data indicate that PAF-EZH2 complex is distinct from the conventional PRCs in cancer cells. Also, we questioned whether PCNA is required for PAF’s interaction with either PAF or β-catenin. Interestingly, β-catenin-PAF and EZH2-PAF complexes existed only in PCNA-free fractions (Figure 3M, compare lanes 1 and 2), which is consistent with PCNA-independent mitogenic role of PAF in colon cancer cell proliferation (Figure 1I). Due to exclusive interaction of PAF with either PCNA or β-catenin, we asked whether Wnt signaling activation affects either PAF-β-catenin or PAF-PCNA interaction. Co-IP assays showed that, in HeLa cells, PAF-β-catenin interaction was only detected upon LiCl treatment, while PAF-EZH2 interaction remained constant. Moreover, PAF-PCNA association was decreased by LiCl or Wnt3A treatment (Figures 3N and 3O, compare lanes 3 and 4). These data suggest that Wnt signaling activation is required for PAF-β-catenin interaction. Due to absence of endogenous Wnt signaling activity in HeLa cells, we also assessed the effects of active Wnt/β-catenin signaling on PAF-PCNA binding in colon cancer cell lines that exhibit hyperactivation of Wnt signaling by genetic mutations in APC or β-catenin alleles. Surprisingly, PAF-PCNA interaction was barely detectable in colon cancer cell lines, whereas 293T and HeLa cells displayed strong PAF-PCNA association (Figure 3P), implying that active β-catenin may sequester PAF from PCNA. In binding domain mapping analysis, we also found that N-terminal and PIP regions are required for β-catenin interaction (Figure S2), suggesting that β-catenin competes with PCNA for PAF interaction. These results suggest that, upon Wnt signaling activation, PAF is conditionally associated with β-catenin transcriptional complex.
PAF activates β-catenin target genes by recruiting EZH2 to promoters
Previous studies showed that EZH2 interacts with β-catenin (Li et al., 2009; Shi et al., 2007). Also, we found that PAF is physically associated with EZH2, independently of PRC2 complex (Figure 3). These evidences prompted us to ask whether EZH2 mediates PAF-induced Wnt signaling hyperactivation. Given PAF-EZH2-β-catenin complex formation, we tested whether EZH2 is also associated with the promoters of β-catenin target genes. Intriguingly, PAF, EZH2, and β-catenin steadily co-occupied the promoters of c-Myc,Cyclin D1, and Axin2 in HCT116 cells carrying β-catenin mutation, whereas PCNA, EED, and Suz12 did not (Figure 4A), which recapitulates PRC2 complex-independent association of EZH2 with PAF (see Figures 3K and 3N). Next, we asked whether PAF, EZH2, and β-catenin are recruited to β-catenin target gene promoter upon Wnt signaling activation, as PAF-β-catenin interaction was dependent of Wnt signaling activation (Figure 3N). In HeLa cells, we found that PAF, EZH2, and β-catenin conditionally bound to TBEs in the c-Myc and Axin2 promoters, only upon LiCl treatment (Figure 4B), indicating that Wnt signaling activation is a prerequisite for PAF-β-catenin-EZH2’s promoter association. To further confirm the specificity of PAF-EZH2-β-catenin’s recruitment to β-catenin target promoters, we performed ChIP promoter scanning of 10 kb of the c-Myc promoter, and found that PAF, EZH2, and β-catenin specifically co-occupied the proximal promoter containing TBEs of the c-Myc gene (at -1037 and -459 bp) (He et al., 1998) in HCT116 cells (Figure 4C). Also, the analysis of EZH2 ChIP-sequencing data from mouse embryonic stem cells showed that EZH2 was specifically enriched in the proximal promoters of β-catenin targets (Lef1, Lgr5, c-Myc, and Axin2) (Figure 4D).
Next, we asked whether EZH2 promoter recruitment is necessary for activation of β-catenin target gene transcription. Previously, depletion of EZH2 was shown to inhibit c-Myc expression in DLD-1 colon cancer cells (Fussbroich et al., 2011). Consistently, EZH2 knockdown downregulated β-catenin target genes, Axin2and Cyclin D1 in HCT116 cells (Figure 4E), and decreased LiCl-induced β-catenin reporter activation (Figure 4F), suggesting that EZH2 is required for PAF-mediated Wnt target gene hyperactivation. These results are also supported by previous finding that EZH2 enhances β-catenin transcriptional activity by connecting β-catenin with the Med1/RNA polymerase II (Pol II) complex (Shi et al., 2007). Indeed, Med1/TRAAP220 and Pol II conditionally binds to c-Myc and Axin2 promoters in LiCl-treated HeLa cells (Figure 4G). Given that PRC2-indepednent interaction between EZH2 and PAF (Figures 3K and 3N), we asked whether EZH2’s histone methyltransferase activity is dispensable in β-catenin regulation. We utilized an EZH2 point mutant (F681I) that disrupts the contact between the EZH2 hydrophobic pocket and histone lysine residue H3K27 (Joshi et al., 2008). Ectopic expression of either EZH2 or EZH2-F681I enhanced β-catenin reporter activity (Figure 4H). Also, PAF knockdown did not change the H3K27 methylation status (H3K27me3) of proximal promoters of c-Myc, Axin2, Cyclin D1, and DCC in HCT116 cells (Figure 4I). These results indicate a methyltransferase-independent role of EZH2 in transactivating β-catenin targets.
Due to PAF’s (1) small size (111 amino acids, one α-helix), (2) lack of a specific catalytic domain, and (3) binding to both β-catenin and EZH2, PAF may facilitate the interaction between EZH2 and β-catenin through recruiting EZH2 to the promoter. We tested this using ChIP assays for EZH2 in the setting of PAF depletion. Indeed, PAF-depleted HCT116 cells displayed decreased EZH2-association at the c-Myc promoter (Figure 4J), suggesting that PAF assists or is needed to recruit EZH2 to β-catenin transcriptional complex. Also, β-catenin knockdown decreased recruitment of PAF and EZH2 to promoters (Figure 4K), showing that PAF and EZH2 occupy target promoters via β-catenin. We then asked whether PAF promotes β-catenin-EZH2 binding. In vitro binding assays showed that the addition of GST-PAF protein increased EZH2-β-catenin association (Figure 4L). Moreover, ectopic expression of PAF promoted the EZH2-β-catenin interaction in HeLa cells treated with LiCl (Figure 4M). Additionally, we tested whether Wnt signaling-induced post-translational modification of either β-catenin or PAF is required for EZH2 interaction. However, in GST pull-down assays, we found that bacterially expressed either GST-β-catenin or –PAF bound to EZH2 (Figure S3). Due to the lack of post-translational modification in GST protein expression system, these data indicate that post-translation modification of either β-catenin or PAF is not necessary for EZH2 interaction. Together, these results suggest that PAF acts as a molecular adaptor to facilitate EZH2-β-catenin complex, and subsequently enhances the transcriptional activity of the β-catenin transcriptional complex at Wnt target promoters (Figure 4N).
Intestinal tumorigenesis following PAF conditional expression
Having determined that PAF is overexpressed in colon cancer cells and hyperactivates Wnt/β-catenin signaling, we aimed to determine whether mimicking PAF overexpression drives intestinal tumorigenesis, using genetically engineered mouse models. To conditionally express PAF, we generated doxycycline (doxy)-inducible PAF transgenic mice (TetO-PAF-IRES-emGFP [iPAF]). For intestine-specific expression of PAF, we used iPAF:Villin-Cre:Rosa26-LSL-rtTA mouse strains. Villin-Cre is specifically expressed in intestinal epithelial cells (IECs), including ISCs and progenitor cells. Cre removes a floxed stop cassette (loxP-STOP-loxP [LSL]) from the Rosa26 allele and induces rtTA expression. Upon doxy treatment, rtTA drives the transcriptional activation of the tetracycline-responsive element promoter, resulting in conditional transactivation of PAF selectively in IECs. We also utilized the Rosa26-rtTA strain for ubiquitous expression of PAF (Figure 5A and Figure S4). First, we examined the effects of PAF induction on IEC proliferation using a crypt organoid culture system (Figure S5A). Intriguingly, PAF conditional expression (2 weeks) induced expansion of the crypt organoids (Figures 5B and 5C), which recapitulates the mitogenic function of PAF (Figure 1). In mouse, IEC-specific PAF expression (iPAF:Villin-Cre:Rosa26-LSL-rtTA; 2 months) developed adenoma in both small intestine and colon (Figure 5D). Also, microscopic analysis using hematoxylin and eosin (H&E) staining showed aberrant IEC growth and crypt foci formation (Figures 5E and 5F), with disorganized epithelial cell arrangements (Figure S5B). Consistently, PAF-induced IEC hyperproliferation was manifested by increased Ki67 expression, a mitotic marker (Figure 5G). Importantly, these lesions exhibited the upregulation of CD44, a β-catenin target gene, whereas CD44 was expressed strictly in the crypts of normal intestine (Figure 5H). Next, we examined whether PAF directly hyperactivates Wnt/β-catenin in vivo using BAT-gal, a β-catenin reporter transgenic mouse carrying multiple TBEs followed by a LacZ reporter. To quantify the early effects of PAF on β-catenin activity, we treated mice with doxy for 1 week, and found that short-term induction of PAF increased β-catenin transcriptional activity as represented by enhanced X-gal staining (Figure 5I). Moreover, conditional PAF expression upregulated the β-catenin target genes, Axin2, Lgr5, CD44, Cyclin D1, and c-Myc in crypt organoids (Figure 5J). Additionally, mice ubiquitously expressing PAF exhibited intestinal hypertrophy (Figure S5C), which is similar to that induced by R-Spondin1, a secreted Wnt agonist (Kim et al., 2005). These data strongly suggest that PAF expression is sufficient to initiate intestinal tumorigenesis via Wnt signaling hyperactivation.
Figure 5Induction of intestinal neoplasia by PAF expression
Herein we reveal the unexpected role of PAF in modulating Wnt/β-catenin signaling. PAF enhances the transcription of Wnt targets by recruiting EZH2 to the β-catenin transcriptional complex. This is similar to the mechanism by which Lgl/BCL9 binds to β-catenin and thereby recruits the PHD-finger protein Pygopus, to bridge the β-catenin/TCF complex to Med12 and Med13 (Carrera et al., 2008). Importantly, due to specific overexpression of PAF in cancer cells, our studies identified an additional layer of the regulatory mechanism of β-catenin target gene transactivation.
In cancer cells, the upregulation of EZH2 contributes to tumorigenesis through the epigenetic repression of various genes including tumor suppressor genes, Wnt antagonists, and DNA repair genes (Chang et al., 2011; Cheng et al., 2011; Kondo et al., 2008). Our results propose a noncanonical function of EZH2 in activating β-catenin target genes in conjunction with PAF. Consistently, recent study also suggests methyltransferase activity-independent function of EZH2 in gene activation (Xu et al., 2012). Moreover, this non-canonical role of EZH2 is supported by several lines of evidence: (a) EZH2 transactivates β-catenin target genes (Li et al., 2009; Shi et al., 2007) (Figures 4E and 4F); (b) EZH2 overexpression in murine mammary epithelium induces ductal hyperplasia (Li et al., 2009), which phenocopies that in a ∆Nβ-catenin (constitutively active form of β-catenin) mouse model (Imbert et al., 2001); (c) EZH2 occupies β-catenin target promoters (Figures 4A-D); and (d) EZH2’s methyltransferase activity is dispensable for β-catenin target activation (Figures 4H and 4I). Moreover, similar to PAF expression in the AER (Figure 2A), EZH2 is also specifically expressed there to maintain of Hox cluster gene transcription (Wyngaarden et al., 2011). Thus, it is plausible that EZH2 and PAF cooperatively control Hox gene activation in the developing limb. Interestingly, despite the presence of a physical and functional connection between Bmi1 and EZH2 in H3K27me3-mediated gene repression, EZH2 is expressed only in crypt IECs including ISCs (Figure S6), whereas Bmi1 is expressed in ISCs at position 4 (Figure 3J), implying a Bmi1-independent role for EZH2 in gene regulation. These results demonstrate the novel function of EZH2 in β-catenin target gene activation, independent of the histone methyltransferase activity of EZH2.
Previously, we found that TERT, a catalytic subunit of telomerase, positively modulates Wnt signaling (Park et al., 2009), elucidating a non-telomeric function of telomerase in development and cancer. Here our results propose that one component of DNA damage bypass process also functions in regulating Wnt signaling, dependent of context. In cancer, PAF overexpression may play a dual role in inducing (a) cell hyperproliferation (via Wnt signaling hyperactivation) and (b) the accumulation of mutations arising from DNA lesion bypass (by PAF-mediated TLS) (Povlsen et al., 2012). Importantly, PAF is only expressed in cancer cells, but not in normal epithelial cells. Thus, upon DNA damage, instead of cell growth arrest to permit high-fidelity DNA repair, the PAF overexpression in cancer cells is likely to induce DNA lesion bypass by facilitating TLS. However, in the setting of Wnt signaling deregulation, nuclear β-catenin sequesters PAF from PCNA and utilize PAF as a co-factor of transcriptional complex, which induces Wnt signaling hyperactivation and possibly lead to increased mutagenesis.
We observed that PAF marked the stemness of ISCs and mouse embryonic stem cells (Figure S7), implicating its roles in stem cell regulation under physiological conditions. In a previous study, a PAFgermline knockout mouse model displayed defects in hematopoietic stem cell self-renewal (Amrani et al., 2011), suggesting a crucial role of PAF in stem cell maintenance and activation. In the intestine, β-catenin activation in Lgr5-positive or Bmi1-positive cells is sufficient to develop intestinal adenoma (Barker et al., 2009; Sangiorgi and Capecchi, 2008), suggesting an essential role of tissue stem cells in tumor initiation. Considering PAF expression in Bmi1-positive ISCs, PAF upregulation in ISCs likely hyperactivates the Wnt/β-catenin signaling and contributes to intestinal tumor initiation.
Despite the critical role of Wnt signaling in early vertebrate, development PAF germline knockout mice are viable (Amrani et al., 2011). It is noteworthy that, whereas deletion of any core component in the Wnt signaling pathway causes embryonic lethality, mice with germline knockout of Wnt signaling modulators, including Nkd1/2, Pygo1/2, and BCL9/9-2, exhibit no lethal phenotypes (Deka et al., 2010; Schwab et al., 2007; Zhang et al., 2007). This may result from the robustness of Wnt signaling during embryogenesis because of functional compensation not only via the presence of multiple Wnt signaling regulators per se but also via other types of signaling crosstalk. Therefore, as described previously in pRb studies (Sage et al., 2003), acute deletion of PAF in a conditional knockout mouse model may disrupt the developmental balance or tissue homeostasis, and then reveal the full spectrum of the physiological and pathological roles of PAF in tumorigenesis. Taken together, our findings reveal unexpected function of PAF and EZH2 in modulating Wnt signaling, and highlight the impacts of PAF-induced Wnt signaling deregulation on tumorigenesis.
11.1.9 PAF Makes It EZ(H2) for β-Catenin Transactivation
In this issue of Molecular Cell, Park and colleagues (Jung et al., 2013) show that PAF (PCNA-associatedfactor) binds to and hyperactivates transcriptional function of β-catenin in colon cancer cells by recruiting EZH2 to the coactivator complex. PAF-β-catenin and PAF-PCNA interactions are competitive, raising the question of whether β-catenin might regulate PCNA-dependent DNA replication and repair.
Wnt signaling through stabilization of transcription co-activator β-catenin plays critical roles in animal development and tissue homeostasis, and its deregulation is involved in myriad human diseases including cancer (Clevers and Nusse, 2012). Notably, most colorectal cancers (CRCs) have elevated β-catenin signaling caused by mutations of Wnt pathway components such as the tumor suppressor APC (Adenomatosis polyposis coli) and β-catenin itself (Clevers and Nusse, 2012). Much effort has focused on studying β-catenin-dependent transactivation in CRCs, including the current study by Park and colleagues that identifies PAF as an unexpected β-catenin co-activator (Jung et al., 2013).
PAF, for PCNA (proliferating cell nuclear antigen)-associated factor (also known as KIAA0101 or p15PAF), is an interacting partner of PCNA (Yu et al., 2001). PCNA has a key role in DNA replication and repair by assembling various DNA polymerase and repair complexes at the replication fork (Mailand et al., 2013). Dynamic regulation of PAF abundance and/or interaction with PCNA appears to be important for engaging DNA damage repair and bypass pathways (Emanuele et al., 2011; Povlsen et al., 2012). PAF is overexpressed in many types of cancers and required for cell proliferation (e.g., Yu et al., 2001).
In the current study (Jung et al., 2013), Jung et al. show that PAF is overexpressed in CRCs in a manner that parallels expression of Axin2, an established Wnt/β-catenin target gene. PAF knockdown inhibits CRC proliferation, and this effect is independent of PAF-PCNA interaction and can be rescued by a PAF mutant that does not binds to PCNA or by β-catenin overexpression. PAF knockdown downregulates the expression of Wnt/β-catenin target genes Cyclin D1, c-Myc, and Axin2 in a CRC line, leading the authors to hypothesize that PAF participates in Wnt/β-catenin signaling. Indeed PAF knockdown reduces, and its overexpression augments, Wnt/β-catenin responsive TOPFLASH reporter and target gene expression induced by Wnt3a or by pharmacological agents that stabilize β-catenin. In Xenopus embryos, PAF synergizes with β-catenin to induce Wnt target gene expression and axis duplication (a hallmark of Wnt/β-catenin activation). In mouse embryos, PAF is highly expressed in regions known for Wnt/β-catenin signaling such as the apical ectodermal ridge of the limb bud. Therefore PAF appears to be a positive regulator of Wnt/β-catenin signaling in CRCs and vertebrate embryos.
PAF does not affect β-catenin protein levels and is localized in the nucleus. Protein binding assays show that PAF interacts, directly or indirectly, with β-catenin (via the Armadillo-repeat domain) and its DNA-bound partner TCF (T Cell factor). Indeed PAF is associated with promoters of Wnt/β-catenin target genes in chromatin in CRC cells. Interestingly in the mouse intestine, the PAF protein is enriched in Bmi (B lymphoma Mo-MLV insertion region 1 homolog)-positive stem cells (at the “+4” position) (Sangiorgi and Capecchi, 2008). Bmi1 is a component of Polycomb Repressive Complex 1 (PRC1), which, together with the PRC2 complex that modifies Histone H3, has critical functions in transcriptional epigenetic silencing. Previous studies have suggested that a core PRC2 component, EZH2 (enhancer of zeste homolog 2), is a partner and paradoxically a co-activator of β-catenin, acting in a manner that is independent of EZH2’s methyltransferase activity (Li et al., 2009; Shi et al., 2007). Jung et al. found that PAF indeed interacts with both Bmi1 and EZH2, but not other PRC2 components, and EZH2 overexpression augments β-catenin transcriptional activity. PAF, EZH2, and β-catenin are found to co-occupy promoters of several Wnt/β-catenin target genes in CRC and mouse ES cells, and PAF depletion decreases EZH2 association with the c-Myc promoter, and β-catenin depletion decreases the association of both PAF and EZH2 with the promoter. Thus the β-catenin-PAF-EZH2 complex appears to constitute a chain of co-activators (Figure 1), and indeed PAF, which binds to both β-catenin and EZH2, enhances β-catenin-EZH2 co-immunoprecipitation. Together with an earlier study (Shi et al., 2007), these results suggest a model that PAF brings EZH2 and the associated RNA polymerase II Mediator complex to β-catenin target genes for transactivation in CRCs (Figure 1). Consistent with this model, transgenic overexpression of PAF in the mouse intestine induces β-catenin-dependent target and reporter gene expression, intestinal overgrowth, and adenoma formation in vivo and crypt organoid expansion in vitro, resembling Wnt/β-catenin signaling activation in the gastrointestinal tract.
Figure 1 β-catenin transactivation mediated by PAF and EZH2 in the G1 phase and a speculative role of β-catenin in modulating PAF-PCNA-dependent DNA replication and repair/bypass pathways in the S phase.
PAF and EZH2 represent newer additions to β-catenin’s plethora of co-activators (Mosimann et al., 2009), which offer multiple routes to engage the basal transcription apparatus. These co-activators may have partially redundant and/or context-dependent functions for numerous Wnt/β-catenin-dependent gene programs. Mouse mutants that lack an individual β-catenin co-activator are often viable (MacDonald et al., 2009; Mosimann et al., 2009). Paf−/− mice are viable but exhibit defects in hematopoietic stem cell properties (Amrani et al., 2011). PAF is also expressed in self-renewing mouse ES cells but the expression is downregulated upon ES cell differentiation (Jung et al., 2013). Whether PAF has a general role in self-renewal of embryonic and adult stem cells through its role in β-catenin signaling or DNA replication and repair pathways remains to be investigated.
PAF-β-catenin interaction is observed under Wnt stimulation, likely as a consequence of β-catenin accumulation (Jung et al., 2013). In some cell types PAF is ubiquitinated and degraded by the anaphase promoting complex and thus exhibits the lowest level in the G1 phase of the cell cycle (Emanuele et al., 2011). In these cells PAF may have a limited role as a co-activator for β-catenin-dependent transcription, which primarily occurs in G1. But in CRC and other cancers where PAF is overexpressed, PAF may have a prominent role as a β-catenin co-activator.
PAF-PCNA interaction is well documented (e.g., Yu et al., 2001). Surprisingly however, in CRCs with high levels of β-catenin, PAF-PCNA interaction is barely detectable (Jung et al., 2013). Conversely, in cells where the basal level of Wnt/β-catenin signaling is low, PAF-PCNA interaction is detected but is diminished by Wnt3a or pharmacological agents that stabilize β-catenin (Jung et al., 2013). PAF seems to interact with β-catenin and PCNA via an overlapping domain (although this remains to be better defined), offering a possible explanation why PAF-β-catenin and PAF-PCNA complexes appear to be mutually exclusive (Jung et al., 2013). This may simply reflect the fact that PAF-β-catenin (for RNA transcription) and PAF-PCNA (for DNA replication/repair) complexes act in G1 and S, respectively (Figure 1). However, when β-catenin levels are high in Wnt-stimulated cells or in CRCs, one may speculate that β-catenin accumulation could inhibit PAF-PCNA complex formation in the S phase, thereby enabling Wnt/β-catenin signaling to modulate PAF-PCNA-dependent DNA replication and repair/bypass pathways (Figure 1). This would constitute an unsuspected role for Wnt/β-catenin signaling in genomic stability beyond its established transcriptional function and could have implications to tumorigenesis.
Amrani YM, Gill J, Matevossian A, Alonzo ES, Yang C, Shieh JH, Moore MA, Park CY, Sant’Angelo DB, Denzin LK. J Exp Med. 2011;208:1757–1765. [PMC free article] [PubMed]
Clevers H, Nusse R. Cell. 2012;149:1192–1205. [PubMed]
Emanuele MJ, Ciccia A, Elia AE, Elledge SJ. Proc Natl Acad Sci USA. 2011;108:9845–9850.[PMC free article] [PubMed]
Jung H-Y, Jun S, Lee M, Kim H-C, Wang X, Ji H, McCrea PD, Park J-I. Molecular Cell. 2013 this issue, *bxs. [PMC free article] [PubMed]
Li X, Gonzalez ME, Toy K, Filzen T, Merajver SD, Kleer CG. Am J Pathol. 2009;175:1246–1254.[PMC free article] [PubMed]
Background: Ovarian cancer is one of the major causes of death in women worldwide. Despite improvements in conventional treatment approaches, such as surgery and chemotherapy, a majority of patients with advanced ovarian cancer experience relapse and eventually succumb to the disease; the outcome of patients remains poor. Hence, new therapeutic strategies are urgently required. The phosphatidylinositol 3-kinase (PI3K)/AKT/mammalian target of rapamycin (mTOR) is activated in approximately 70 % of ovarian cancers, resulting in hyperactive signaling cascades that relate to cellular growth, proliferation, survival, metabolism, and angiogenesis. Consistent with this, a number of clinical studies are focusing on PI3K pathway as an attractive target in the treatment of ovarian cancer. In this review, we present an overview of PI3K pathway as well as its pathological aberrations reported in ovarian cancer. We also discuss inhibitors of PI3K pathway that are currently under clinical investigations and the challenges these inhibitors face in future clinical utility.Methods: PubMed was searched for articles of relevance to ovarian cancer and the PI3K pathway. In addition, the ClinicalTrials.gov was also scanned for data on novel therapeutic inhibitors targeting the PI3K pathway. Results: Genetic aberrations at different levels of PI3K pathway are frequently observed in ovarian cancer, resulting in hyperactivation of this pathway. The alterations of this pathway make the PI3K pathway an attractive therapeutic target in ovarian cancer. Currently, several inhibitors of PI3K pathway, such as PI3K/AKT inhibitors, rapamycin analogs for mTOR inhibition, and dual PI3K/mTOR inhibitors are in clinical testing in patients with ovarian cancer. Conclusions: PI3K pathway inhibitors have shown great promise in the treatment of ovarian cancer. However, further researches on selection patients that respond to PI3K inhibitors and exploration of effective combinatorial therapies are required to improve the management of ovarian cancer.
Fig.1. Inputs from receptor tyrosine kinases (RTKs) and G protein-coupled receptors (GPCR) to class I PI3K.
Fig. 2. Schematic representation of the PI3K/AKT/mTOR signaling pathway.
Fig.3. PI3K/AKT/mTOR inhibitors.
AKT inhibitors
AKT inhibitors can be grouped into three classes including lipid based phosphatidylinositol (PI) analogs, ATP-competitive inhibitors, and allosteric inhibitors. Perifosine, which is the most clinically studied AKT inhibitor, is a lipid-based PIanalog that targets the pleckstrin homology domain of AKT, preventing its translocation to the cell membrane. Amongthe three classes of AKT inhibitors, allosteric AKT inhibitors display highly specific selectivity for AKT isoforms. Considering the genetic background of ovarian cancer, allosteric AKT inhibitors such as MK2206 that can target both AKT1 and AKT2 might be the best agents for treating ovarian cancer.In clinical trials, AKT inhibitors have shown similar toxicities to those caused by PI3K inhibitors, such as hyperglycemia, rashes, stomatitis, and gastrointestinal side effects [25].
mTOR inhibitors
Rapamycin and its analogs Rapamycin (sirolimus), a potent inhibitor of mTORC1, was first isolated in 1975 from the bacterium Streptomyces hygroscopicus. Rapamycin inhibits mTORC1 by first binding to the intracellular protein FK506 binding protein 12 (FKBP12). The resultant rapamycin–FKBP12 complex then binds to the FKBP12–rapamycin-binding domain (FRB) of mTORC1 and inhibits the serine/threonine kinase activity of mTORC1 via an allosteric mechanism. In contrast to mTORC1, the rapamycin–FKBP12 complex cannot interact with the FRB domain of mTORC2, and thus,mTORC2 is generally resistant to rapamycin treatment [12]. As rapamycin displays very poor water solubility, which limits its clinical use, several soluble ester analogs of rapamycin (rapalogs) have been developed [12]. Currently, these analogs include temsirolimus, everolimus, and ridaforolimus. Temsirolimus and everolimus are formulated for intravenous and oral administration, respectively. Ridaforolimus was initially developed as an intravenous formulation, but an oral formulation was subsequently produced [12,28]. Clinically, rapalogs are generally well tolerated, with the most common side effects including stomatitis, rashes, fatigue, hyperglycemia, hyperlipidemia, hypercholesterolemia, and myelosuppression [3,12,25].
ATP-competitive inhibitors
Different from rapalogs, ATP-competitive inhibitors do not require co-factors such as FKBP12 to bind to mTOR. By competingwith ATP for theATP-binding sites of mTOR, this class of mTOR inhibitors can inhibit the kinase activity of both mTORC1 and mTORC2. Although there is a concern that the simultaneous inhibition of mTORC1 and mTORC2 might result in greater toxicities in normal tissues, ATP-competitive mTOR inhibitors have been shown to display stronger anti-proliferative activity than rapalogs across a broad range of cancers includingovarian cancer [12,15].
Metformin
Metformin,the most commonly prescribed oral anti-diabetic agent, has been shown to reduce the incidence of malignancies in patients with diabetes. The activation of 5′ adenosine monophosphateactivated protein kinase (AMPK) by metformin plays an important role in mediating the drug’s effects. AMPK activation results in the phosphorylation and activation of TSC2, which exerts inhibitory effects on mTORC1. Metformin-induced AMPK activation also reduces AKT activity by inhibiting insulin receptor substrate 1 (IRS-1). Ultimately, AMPK activation results in the inhibition of the PI3K/AKT/mTOR signaling pathway, making metformin an effective treatment for cancer [28].
Pan-class I PI3K inhibitors Pan-class IPI3K inhibitors can inhibit the kinase activity ofall 4 isoforms of classI PI3K.The main advantage of pan-class IPI3K inhibitors is that most cancer cells express multiple PI3K isoforms with redundant oncogenic signaling functions. Early clinical trials have suggested that the most common toxicitiesof pan-class IPI3K inhibitors are hyperglycemia, skin toxicities, stomatitis, and gastrointestinal side effects. Of these, hyperglycemia is likely to be a mechanism-based toxicity given the well described role of PI3K in insulin receptor signaling [3,25].
Isoform-selective PI3K inhibitors
This class of agents target the specific PI3K p110 isoforms involved in particular types of cancer. The p110α isoform (which is encoded by the PIK3CA gene) is a frequent genetic driver (PIK3CA mutations) of ovarian cancer, whereas p110β activity is known to be essential in cancer cells lacking PTEN. As for the p110δ isoform, it plays a fundamental role in the survival of normal B cells and is implicated in malignancies affecting this lineage. Thus, the main theoretical advantage of these inhibitors is that they have the potential to completely block the relevant target whilst causing limited toxicities compared with pan-PI3K inhibitors. Consistent withthese findings, preclinical studies have detected significant activities of PI3Kα inhibitor in tumors exhibiting PIK3CA mutations, PI3Kβ inhibitors in tumors with PTEN loss, and PI3Kδ inhibitors in hematologic malignancies. In addition, PI3Kδ inhibitors have already shown very promising activity in patients with chronic lymphocytic leukemia [26].
Dual PI3K/mTOR inhibitors
Structural similarities between the ATP-binding domain of p110 and the catalytic domain of mTOR have led to the development of a class of agents that inhibit both class I PI3K and mTORC1/2. Theoretically, dual mTOR/PI3K inhibitors should lead to more complete suppression of the PI3K/AKT/mTOR pathway than targeting either component independently.In agreement with this, in preclinical studies of ovarian cancer dual PI3K/mTOR inhibitors were found to exhibit greater in vitro and in vivo anti-tumor activity than mTOR inhibitors alone [27]. The safety profile of these inhibitors is similar to that of pan-PI3K inhibitors, with common adverse events including nausea, diarrhea, fatigue, and vomiting [3,25].
11.1.11 Endogenous, hyperactive Rac3 controls proliferation of breast cancer cells by a p21-activated kinase-dependent pathway
Uncontrolled cell proliferation is a major feature of cancer. Experimental cellular models have implicated some members of the Rho GTPase family in this process. However, direct evidence for active Rho GTPases in tumors or cancer cell lines has never been provided. In this paper, we show that endogenous, hyperactive Rac3 is present in highly proliferative human breast cancer-derived cell lines and tumor tissues. Rac3 activity results from both its distinct subcellular localization at the membrane and altered regulatory factors affecting the guanine nucleotide state of Rac3. Associated with active Rac3 was deregulated, persistent kinase activity of two isoforms of the Rac effector p21-activated kinase (Pak) and of c-Jun N-terminal kinase (JNK). Introducing dominant-negative Rac3 and Pak1 fragments into a breast cancer cell line revealed that active Rac3 drives Pak and JNK kinase activities by two separate pathways. Only the Rac3-Pak pathway was critical for DNA synthesis, independently of JNK. These findings identify Rac3 as a consistently active Rho GTPase in human cancer cells and suggest an important role for Rac3 and Pak in tumor growth.
Uncontrolled cell proliferation is a major feature of cancer. Experimental cellular models have implicated some members of the Rho GTPase family in this process. However, direct evidence for active Rho GTPases in tumors or cancer cell lines has never been provided. In this paper, we show that endogenous, hyperactive Rac3 is present in highly proliferative human breast cancer-derived cell lines and tumor tissues. Rac3 activity results from both its distinct subcellular localization at the membrane and altered regulatory factors affecting the guanine nucleotide state of Rac3. Associated with active Rac3 was deregulated, persistent kinase activity of two isoforms of the Rac effector p21-activated kinase (Pak) and of c-Jun N-terminal kinase (JNK). Introducing dominant-negative Rac3 and Pak1 fragments into a breast cancer cell line revealed that active Rac3 drives Pak and JNK kinase activities by two separate pathways. Only the Rac3–Pak pathway was critical for DNA synthesis, independently of JNK. These findings identify Rac3 as a consistently active Rho GTPase in human cancer cells and suggest an important role for Rac3 and Pak in tumor growth.
Rac proteins are members of the Rho GTPase family and act as molecular switches in regulating a variety of biological response pathways, including cell motility, gene transcription, cell transformation, and cell-cycle progression (1). The Rac family includes Rac1, the myeloid-lineage-specific Rac2, and the recently cloned Rac3 proteins (2). Rac3 differs from Rac1 and Rac2 in two domains, the insert region and the C terminus, which influence transformation (3, 4), interaction with guanine nucleotide exchange factors (GEFs) (5, 6), and subcellular localization (7, 8). Small GTPases, including Rac, cycle between an inactive GDP-bound state and an active GTP-bound state. Two classes of regulatory factors, GTPase-activating proteins (GAPs) and GEFs, determine by their opposing effects the ratio of GDP versus GTP, which is bound to the GTPase (1). GAP proteins increase the intrinsic rate of GTP hydrolysis, rendering the GTPase inactive, whereas GEFs enhance the exchange of bound GDP for GTP, thereby activating the protein. Active Rac regulates distinct downstream signaling pathways by interacting with specific effector proteins, including a family of serine-threonine protein kinases termed Paks (p21-activated kinases) (9–11).
Apart from its well documented role in cytoskeletal rearrangements in growth factor-stimulated cells (12), Rac1 is required for Ras-induced malignant transformation and is involved in transcription and growth control (1, 13, 14). Recently, the importance of the Rac effector Pak in cell transformation has been highlighted by inhibiting RasV12- and Rac1V12-induced transformation of Rat-1 fibroblasts with a catalytically inactive form of Pak (15, 16). The involvement of Rac1 in driving cell-cycle progression through the G1 phase and stimulating DNA synthesis has been shown by introducing dominant-active and -negative Rac1 mutants into fibroblasts (17, 18). However, the signaling pathways used by Rac to control mitogenesis and proliferation still remain poorly understood. Overexpression of constitutively active Rac-effector-domain mutants in fibroblasts indicated that although Rac1 mediated cyclin D1 transcription by Pak in NIH 3T3 cells (19), Pak was not involved in the DNA synthesis of Swiss 3T3 cells (20). Accumulating evidence, however, suggests higher complexity where Pak-binding proteins, such as the GEF Pix, contribute to the Rac–Pak interaction in vivo and influence subsequent cellular functions (21–23).
All biological functions listed above have been attributed to Rac1 in experimental cell systems using overexpression or microinjection of mutant forms. Endogenously active Rho GTPases, including Rac, have not yet been observed. In this paper, we describe a consistently active Rac3 GTPase leading to hyperactivity of its effector protein kinase, Pak, in human breast cancer-derived epithelial cell lines. Analysis of growth properties and DNA synthesis revealed that both proteins are required to convey the highly proliferative phenotype displayed by these cells.
Highly Proliferating Cancer Cells Contain Hyperactive Rac3.
Comparison of growth rates among several breast cancer cell lines showed that three lines (MDA-MB 435, T47D, and MCF 7) grew faster under normal and low-serum conditions (Fig. (Fig.1).1). Interestingly, in contrast to MDA-MD 231 and Hs578T cells, these three highly proliferative cell lines do not possess mutated Ras (28, 29). To assess whether Rho GTPases drive this cellular phenotype, we determined whether these cell lines contained active GTP-bound Rac or Cdc42. We used a recently described assay, the PBD-pulldown assay (24), which is based on the specific binding of the GTP-bound forms of Rac and Cdc42 to the PBD of Pak (10). Neither active Rac1 (Fig. (Fig.22A) nor active Cdc42 (data not shown) could be detected in any of the cell lysates obtained from serum-starved cells. However, both proteins were detected if the PBD-pulldown assay was performed with in vitro guanosine 5′-[γ-thio]triphosphate (GTP[γS])-loaded cell lysates, confirming that Rac1 and Cdc42 were present in their inactive GDP-bound forms in these cells (Fig. (Fig.22A for Rac1). Next we wanted to determine whether active Rac3 was present in breast cancer cell lines. Because Rac3 effectors have not yet been characterized, we demonstrated by overlay binding and kinase assays that Rac3 bound to and activated Pak as efficiently as Rac1 (data not shown). We verified that the PBD-pulldown assay specifically detected the active GTP-bound form of Rac3 (GTP[γS]-loaded Rac3wt or Rac3V12, Fig. Fig.22B) and not the inactive form. To probe for Rac3 protein in breast cell lysates, a Rac3-specific antibody was used. GST-PBD-pulldown experiments from cell lysates revealed the presence of hyperactive Rac3 in highly proliferative cell lines (MDA-MB 435, T47D, and MCF 7), but not in normal breast cell lines or in less proliferative breast cancer cells (Fig. (Fig.22C). Additionally, as indicated by the virtual absence of Rac3 in the supernatant of the PBD pulldown, all the Rac3 protein present in these cell lines was active (Fig. (Fig.22C). To demonstrate that consistent Rac3 activation is not limited to cell lines, we performed an initial screening of human metastatic breast cancer tissues and found active Rac3 in one of three samples, underlining the potential clinical relevance of the cellular findings (Fig. (Fig.22D).
Differential growth rates of human breast cell lines. pq0104939001
Figure 1Differential growth rates of human breast cell lines. Human breast cell lines, including HMEC 184 (○), MDA-MB 231 (▵), Hs578T (□), MDA-MB 435 (●), T47D (▴), and MCF 7 (♦), were grown in 10% serum (A) or 0.5% serum (B) conditions. The cells were split in duplicate over 6-well plates at 5 × 105 cells per well and counted daily with a hemocytometer for 4 days. Data shown in A and B are representative of three independent experiments.
Figure 2 Active Rac3 is present in highly proliferative cell lines and in human breast cancer tissue. (A and C) Cell lysates from serum-starved breast cancer cell lines without (A and C) or after (+) GTP[γS] loading (A) were incubated with 10 μg of GST-PBD. Active Rac proteins (PBD pulldown) were detected by immunoblot with anti-Rac1 (A) or anti-Rac3 antibodies (C). Blotting of PBD supernatants revealed the GDP-bound form of Rac3 in lysates. Equal amounts of Rac3 protein were detected by immunoblot (IB) in all cell lines. (B) A PBD-pulldown assay of extracts from HeLa cells expressing Myc-Rac3wt or -Rac3 mutants, followed by an anti-Myc immunoblot, detected only active Rac3 (GTP[γS] loading or Rac3V12). (D) PBD pulldown of lysates obtained from three different human metastatic breast cancer tissues, followed by anti-Rac1 and anti-Rac3 immunoblots, revealed active Rac3 in tissue 1. (E) PBD pulldown of lysates derived from MDA-MB 435 and MDA-MB 231 cells expressing LacZ control or Myc-Rac3wt without or after in vitro GTP[γS] loading. Consistent activation of Myc-Rac3wt occurred only in MDA-MB 435 cells. (F) Subcellular localization of Rac1 and Rac3. Cytosol (c) and membranes (m) were obtained after nitrogen cavitation and fractionation of breast cancer cell lines and immunoblotted with anti-Rac1 and anti-Rac3 antibodies. All blots are representative of at least three experiments.
Subcellular Localization and GTPase-Regulatory Factors Influence Rac3 Activity.
Constitutive activation of Ras proteins in cancer cells is often caused by activating point mutations at the switch I or II regions (29). cDNA cloning and complete sequence analysis of full-length Rac3 did not reveal any mutations in the breast cell lines studied and did not explain the observed Rac3 activation. GTPase-regulatory proteins such as GEFs and GAPs, which are usually regulated by upstream stimuli, control cycling between the active and inactive forms of Rac. To confirm the presence of an altered regulatory mechanism involved in Rac3 activation, we used the PBD-pulldown assay to analyze the activation state of Myc-tagged Rac3wt transfected into either MDA-MB 231, a cell line harboring only GDP-Rac3, or MDA-MB 435, a cell line that contains endogenous, active GTP-Rac3. Fig. Fig.22E shows that activated Myc-Rac3 was detected only in the MDA-MB 435 cell line, confirming that the regulation of the GDP/GTP state of Rac3 was altered in these cells. We then investigated several upstream stimuli that have been shown to affect GTPase-regulatory proteins (28, 30–32). We excluded the possibility of an autocrine growth-stimulatory loop by culturing MDA-MB 231 cells with the conditioned medium from MDA-MB 435, which did not affect the Rac3 activation state (data not shown). Treatment of cell cultures with phosphatidylinositol 3-kinase or tyrosine kinase inhibitors, including wortmannin, LY294002, and genistein, did not decrease Rac3 activation (data not shown). At this point, we speculated that an oncogenic, Rac3-specific GEF is present in certain breast cancer cells. GEFs possess a pleckstrin homology domain that is essential for membrane localization and for their oncogenic properties (5, 33). Analysis of the subcellular localization of the Rac family members revealed that Rac3 is located in the membranes of breast epithelial cell lines, independently of its activation state (Fig. (Fig.22F). In contrast, endogenous Rac1 in its inactive GDP-bound state was essentially cytosolic (Fig. (Fig.22F). Thus, the distinct localization of Rac3 and Rac1 may contribute to their different activation states in certain breast cancer cell lines. It is conceivable that the highly proliferative cell lines (Fig. (Fig.1)1) express a constitutively active, membrane-bound Rho GEF that activates adjacent Rac3 protein. This hypothesis was further supported by using an hydroxymethylglutaryl-CoA reductase inhibitor, lovastatin, that interferes with isoprenoid synthesis and thereby with posttranslational processing of GTPases. Unprocessed Rac3 from lovastatin-treated MDA-MB 435 cells was predominantly cytosolic and inactive (GDP-Rac3) (data not shown). The requirement of membrane localization for consistent Rac3 activity was further supported by using a Rac3S189 mutant. Replacing cysteine-189 of the CAAX box with serine abolishes isoprenoid incorporation, rendering the GTPase cytosolic. This Rac3 mutant remained in its inactive GDP-bound state when transfected into MDA-MB 435 cells (data not shown).
Several Rho GTPase-regulating GEFs have been identified (5), including the Rac1-specific GEF Tiam-1, which has been linked to tumors such as invasive T-lymphomas (34). Although Tiam-1 is expressed in virtually all tissues, no evidence of oncogenic truncations or alternative splicing of Tiam-1 transcripts has been found (35). A variation of Tiam-1 transcript levels in certain cancer cell lines might lead to overexpression and possibly activation of Tiam-1 protein. However, the activation state of Rac3 protein in the cell lines used in this study does not seem to correlate with Tiam-1 expression levels as reported by Habets et al. (35). Hyperactivity of Rac3 in cancer cells could also result from an absent or dysfunctional Rac3-specific GAP protein. By accelerating the intrinsic GTP hydrolysis rate, GAPs render the GTPase inactive and act as tumor suppressors. Deletion or mutations in the RasGAP gene NF1 and the RhoGAP homologs bcr and DLC-1 have been reported in cancer cells (36, 37).
Active Rac3 Drives Epithelial Cell Proliferation.
To study whether active Rac3 could account for the high proliferation rate of certain breast cancer cells, we expressed a constitutively active Rac3 mutant (Rac3V12) in normal mammary epithelial cells (HMEC 184) that contain only GDP-Rac3 (Fig. (Fig.22C). Rac3V12 expression significantly increased the incorporation of BrdUrd into nascent DNA (Fig. (Fig.3),3), emphasizing that transfection of active Rac3 drives epithelial cell proliferation.
Rac3V12 induces DNA synthesis in human mammary epithelial cells pq0104939003
Rac3V12 induces DNA synthesis in human mammary epithelial cells
Figure 3Rac3V12 induces DNA synthesis in human mammary epithelial cells. HMEC 184 cells, infected with recombinant LacZ or Rac3V12 Semliki Forest virus, were allowed to express protein for 14 h in serum-free medium containing 10 μM BrdUrd. Cells were fixed and stained with anti-Myc antibody for Myc-Rac3V12 expression level (Upper) or with FITC-conjugated anti-BrdUrd antibody for BrdUrd incorporation (Lower). The presence of bright fluorescent nuclei indicates BrdUrd-positive cells. The percentage was calculated after counting 400 cells in each of three independent experiments.
Hyperactive Pak and c-Jun Kinases in Cancer Cells.
The signaling cascade utilized by Rac proteins to control cell proliferation still remains to be identified (1, 9), but might involve Paks. We analyzed Pak activity in cell lysates derived from serum-starved breast cancer cell lines by using in-gel kinase assays and by usingin vitro kinase assays after immunoprecipitation with Pak-specific antibodies. Pak activity was increased 4- to 6-fold in the three cell lines containing active Rac3 (Fig. (Fig.44A). This increased kinase activity was mainly associated with the Pak2 isoform, which can phosphorylate and positively regulate Raf-1 activity, another key component in cell proliferation (38–40).
Figure 4Rac3 activates Pak and JNK by two different pathways. (A) Breast cancer cell lysates from serum-starved cells were analyzed for Pak and JNK activities. Pak activities in cell lysates were analyzed by in-gel kinase assays. JNK activity was determined by …
Intracellular Rac-regulated signaling pathways impinge on distinct mitogen-activated protein kinase cascades. Constitutively active Rac has been shown to positively regulate the activity of the stress-activated kinases JNK and p38 (1). Moreover, ERK activity can be indirectly stimulated by Rac or mediated by crosstalk between the distinct mitogen-activated protein kinase cascades (1, 41). Determination of distinct mitogen-activated protein and stress-activated protein kinase activities in the breast cell lines studied here showed that consistent Rac3 and Pak kinase activities were associated with enhanced JNK activity (Fig. (Fig.44A). In contrast, no correlation existed between p38 or ERK kinase activities and active Rac3 or Pak (data not shown).
Rac3 Triggers Pak and JNK Activities by Separate Pathways.
To determine whether the highly proliferative phenotype of breast cancer cells depends directly on a consistently active Rac3-Pak-JNK cascade, we used virus-mediated protein expression in MDA-MB 435 cells to examine the ability of Rac3 and Paks to control JNK activation and cellular proliferation. The importance of Pak as an effector protein in Rac-mediated activation of JNK is still controversial and seems to be cell-type-dependent (42). Expression of the PBD domain, which controls the activity of both Rac and Pak (21), completely inhibited Pak and JNK stimulation (Fig. (Fig.44B). The mutation of leucine to phenylalanine at position 107 of the PBD domain suppresses the autoinhibitory function of the PBD (21). Thus, PBD F107 will act only to sequester active Rac3 and blocks its ability to bind and activate endogenous effectors. Expression of either dominant-negative Rac3N17 or PBD F107 almost completely blocked Pak and JNK activities, demonstrating that Rac3 is upstream of these proteins (Fig. (Fig.44B). Moreover, Pak kinase activity can be inhibited independently of Rac3 by overexpressing the kinase autoinhibitory domain, PID, which does not interact with Rac (21, 43). Transfection of PID into MDA-MB 435 cells dramatically inhibited Pak activity as expected, but did not decrease JNK activation (Fig. (Fig.44B). Our results indicate that in MDA-MB 435 cells, consistent stimulation of JNK by Rac3 is independent of PAK activity and that Rac3 initiates two different pathways involving Pak and JNK, respectively.
Rac3 and Pak Are Both Required for Breast Cancer Cell Proliferation.
We subsequently determined which of these two Rac3 pathways promoted the increased cell proliferation in breast cancer cell lines with hyperactive Rac3. We studied the consequence of expressing inhibitory Rac mutants or Pak fragments on DNA synthesis. LacZ-expressing MDA-MB 435 cells still proliferated in low-serum conditions and 35% incorporated BrdUrd (Fig. (Fig.5).5). This percentage increased to 50% when Rac3wt, which will be partially activated in these cells (Fig. (Fig.22E), is expressed (Fig. (Fig.55Bottom Right). In contrast, expression of inhibitory proteins, including Rac3N17 or the PBD that suppressed Pak and JNK activation (Fig. (Fig.44B), almost completely blocked S-phase entry, as indicated by the absence of BrdUrd incorporation (Fig. (Fig.5).5). Expression of the PID that inhibited Pak kinase activity without affecting JNK stimulation (Fig. (Fig.44B) also arrested proliferation in MDA-MB 435 cells (Fig. (Fig.5).5). These experiments emphasize the crucial role of active Rac3 for DNA synthesis in breast cancer cell lines and demonstrate that Pak kinase activity is necessary for Rac3-induced proliferation.
Rac3 mediates proliferation in MDA-MB 435 cells pq0104939005
Figure 5 Rac3 mediates proliferation in MDA-MB 435 cells by a Pak-dependent pathway. MDA-MB 435 cells growing in 0.5% FBS were infected with Semliki Forest virus encoding for LacZ, Rac3N17, Pak1-PBD, Pak1-PBD F107, Pak1-PID, or Rac3wt. After 12 to14 h of protein expression in serum-free medium, 20 μM BrdUrd was added for 20 min before the cells were fixed and stained with anti-Myc antibody and phalloidin for expression (Top) or with FITC-conjugated anti-BrdUrd antibody for BrdUrd incorporation (Lower five micrographs). The presence of bright fluorescent nuclei indicates BrdUrd-positive cells. The percentage was calculated after counting 400 cells in each of four independent experiments.
Our results establish the persistent activation of a small Rho GTPase, Rac3, and the effector kinase Pak in human breast cancer cells. In contrast to Rac1, endogenous Rac3 is localized at the plasma membrane in both guanine nucleotide states. It seems likely that a Rac3 regulatory protein is altered or deleted in highly proliferating cancer cells, and that its specificity toward Rac3 results from the adjacent location of both proteins at the membrane and/or from discrete Rac3 domains, which convey a specific interaction. The cytoskeletal phenotypes of serum-starved breast cancer cells, such as ruffles or lamellipodia typical of Rac1 protein activation, did not seem to correlate with the GDP versus GTP state of endogenous Rac3. This may suggest that Rac family members are specialized in certain cellular functions, as already reported for Rac2 in leukocyte phagocytosis (44) and now demonstrated by us for Rac3 in cancer cell proliferation. Our studies establish further that endogenous, active Rac3 is essential for breast cancer cell proliferation via a Pak-dependent pathway. Paks have been shown to directly phosphorylate Raf kinase, which binds to retinoblastoma protein and regulates its function (45), and to interact with cyclin-dependent kinases to up-regulate cyclin D1 expression (46). Initial screening of various human cancer-derived cell lines revealed the presence of hyperactive Rac3 and Pak kinase in other types of highly proliferating tumors (data not shown). Further investigations, primarily in animal models and clinical settings, will be necessary to assess whether loss of Rac3 and Pak regulation correlates with certain breast tumor stages and is accompanied by specific alterations in cell-cycle regulators. Approaches to inhibit Rac3 or Pak activity would then open a new avenue for cancer therapeutics.
11.1.12 Curcumin-could-reduce-the-monomer-of-ttr-with-tyr114cys-mutationvia autophagy in cell model of familial amyloid polyneuropathy.
Transthyretin (TTR) familial amyloid polyneuropathy (FAP) is an autosomal dominant inherited neurodegenerative disorder caused by various mutations in the transthyretin gene. We aimed to identify the mechanisms underlying TTR FAP with Tyr114Cys (Y114C) mutation. Our study showed that TTR Y114C mutation led to an increase in monomeric TTR and impaired autophagy. Treatment with curcumin resulted in a significant decrease of monomeric TTR by recovering autophagy. Our research suggests that impairment of autophagy might be involved in the pathogenesis of TTR FAP with Y114C mutation, and curcumin might be a potential therapeutic approach for TTR FAP.
Transthyretin (TTR) familial amyloid polyneuropathy (FAP) is an autosomal dominant inherited disease, characterized clinically by progressive sensory, motor, and autonomic impairment, which typically lead to death around a decade after diagnosis.1 Since the first identification of TTR with Val30Met mutation (TTR V30M), the most common gene mutation in FAP patients, more than 100 TTR mutations have been found to cause FAP.2 However, the detailed pathogenesis underlying TTR FAP remains undefined. Previous studies of the TTR V30M mutant have shown that misfolding and self-aggregation of TTR are implicated in the pathogenesis of TTR FAP involving abnormal endoplasmic reticulum (ER) stress.3
Corresponding to the various TTR gene mutations and a wide range of geographical distributions, FAP presents diverse characteristics in genotype-phenotype in different regions. We have recently published the first report of a TTR Tyr114Cys (TTR Y114C) mutation in a Chinese family with TTR FAP.4 Compared with TTR V30M, the TTR Y114C mutation showed different clinical manifestations, and was also observed in a Japanese family.5,6 This suggests that the pathogenesis of the TTR Y114C and TTR V30M mutations might be different. Studies focused on monomer generation and tetramer depolymerization have been performed.1,2 However, the mechanisms underlying the clearing of the abnormally increased monomer are unknown.
Autophagy is the major lysosomal pathway via which cells degrade intracytoplasmic protein. It is widely accepted that autophagy plays a key role in the process of amyloid deposition in certain neurodegenerative diseases, including alpha-synuclein, beta peptides, tau oligomers, and misfolded prion protein.7 Therefore, autophagy may be involved in degradation of the TTR monomer in TTR FAP.
Curcumin and its analogs have demonstrated a protective effect in many diseases involving antimicrobial, antitubercular,8 and anticancer mechanisms,9 and they can also modulate innate immunity.10 Of note, curcumin has been shown to promote autophagy.11 Therefore, we hypothesized that autophagy might be involved in the pathogenetic mechanism of the TTR Y114C mutation in TTR FAP and curcumin might have potential therapeutic role in this disease. In this study, we aimed to identify the role of autophagy in the pathogenetic mechanism of TTR FAP and to assess the therapeutic effect of curcumin in the disease.
TTR Y114C mutation led to increased monomeric TTR and impaired autophagy in vitro
To investigate the alteration of monomeric TTR with different mutations, we generated HEK293T cell lines with wild-type TTR, TTR Y114C, and stable overexpression of TTR V30M. Wild-type TTR represented the normal control and TTR V30M represented the positive control. Western blotting analysis of the TTR level in the cells when cultured for 24 hours showed that the monomer of TTR Y114C and TTR V30M was increased by approximately 2.3 times and 2.78 times, respectively, compared with wild-type TTR (Figure 1A and B). Mutation of TTR Y114C was related to the increase in monomeric TTR, as well as the mutation of TTR V30M.
Changes in autophagy and endoplasmic reticulum stress related to wild-type TTR, TTR V30M, and TTR Y114C dddt-8-2121Fig1
Changes in autophagy and endoplasmic reticulum stress related to wild-type TTR, TTR V30M, and TTR Y114C
Figure 1Changes in autophagy and endoplasmic reticulum stress related to wild-type TTR, TTR V30M, and TTR Y114C.
Next we investigated the activation of several markers associated with ER stress, including ER-resident chaperone BiP and p-eIF2α. Our results showed the levels of BiP and p-eIF2α is higher in TTR V30M than those in wild-type TTR. In contrast, BiP and p-eIF2α levels in TTR Y114C were similar to those in wild-type TTR (Figure 1A and C), indicating ER stress might not be the main pathogenetic mechanism for the TTR Y114C mutation. We then investigated whether autophagy plays a role in the mechanism of TTR Y114C mutation. LC3-II is well known to be a robust marker of autophagosomes, and immunofluorescent staining of LC3-II can be used to assay for autophagosome formation. A high ratio of LC3-II to LC3-I would indicate induction of autophagy. Our results revealed that the ratio of LC3-II/I was markedly decreased for TTR Y114C, but less suppressed for TTR V30M (Figure 1A and D). Likewise, a significant decrease in LC3-II immunoreactivity was detected in TTR Y114C (Figure 1E). The results of Western blotting and immunofluorescence indicated that autophagy in TTR Y114C was significantly downregulated. Therefore, impaired autophagy might be responsible for the pathogenesis of TTR Y114C mutation.
Curcumin decreased monomeric TTR by promoting autophagy
The effects of curcumin were investigated in TTR Y114C and wild-type TTR stable overexpressed HEK293T cells. Curcumin did not show toxic effects in the stable overexpressed cell lines at curcumin concentrations below 10 µM (Figure 2A and B). We chose 5 µM as the experimental concentration, because it is the minimal effective concentration of curcumin in these cell lines. Further, we wanted to determine whether curcumin could decrease monomeric TTR by promoting autophagy at the minimal effective concentration. Therefore, we used curcumin (2.5 µM and 5 µM) as a protective agent to assess whether it could decrease monomeric TTR with mutation by promoting autophagy. Quantification of LC3-II and LC3-I indicated markedly higher activation of LC3 in TTR Y114C treated with curcumin 5 µM for 24 hours (Figure 2D). In contrast, treatment with curcumin at different concentrations could not activate LC3 in wild-type TTR (Figure 2C, E). We next examined the ratio of monomers to tetramers in TTR Y114C, which was significantly decreased after 24 hours of treatment with 5 µM curcumin compared with no treatment with curcumin (Figure 2D and F). However, for wild-type TTR, the ratio of monomers to tetramers was unchanged after treatment with curcumin (Figure 2C and E). These results indicate that treatment with curcumin 5 µM for 24 hours was able to decrease the monomer in the TTR Y114C mutation by promoting autophagy.
Curcumin decreased monomeric TTR by promoting autophagy dddt-8-2121Fig2
Curcumin decreased monomeric TTR by promoting autophagy
Figure 2Curcumin decreased monomeric TTR by promoting autophagy.
Protective effect of curcumin on TTR Y114C could be partially blocked by 3-MA
To further validate whether the decrease in monomer by curcumin in our experiments was mediated by autophagy, 3-MA, an inhibitor of autophagosome formation, was implied to negatively regulate autophagy. 3-MA (1 mM) was added to the cell culture medium 2 hours before curcumin and incubated for 24 hours. Analysis of LC3, tetrameric TTR, and monomeric TTR from TTR Y114C revealed that 3-MA partly reversed the LC3 II activation induced by curcumin and increased the monomer of TTR Y114C (Figure 3). These results confirm that curcumin induced the decrease in the TTR Y114C monomer by promoting the autophagy pathway.
Figure 3 Protective effect of curcumin on TTR Y114C could be partially blocked by 3-MA.
Discussion
TTR FAP is a severe autosomal dominant inherited disease, for which the treatment options are limited. Liver transplantation performed early in the course of the disease is the only therapeutic strategy known to stabilize this neuropathy.1,13 More recently, tafamidis meglumine, a potent inhibitor of misfolding and deposition of mutated TTR, has completed an 18-month, placebo-controlled Phase II/III clinical trial for the treatment of FAP.14 However, in June 2012, the US Food and Drug Administration Peripheral and Central Nervous System Drugs Advisory Committee rejected this drug, stating a lack of convincing data supporting its efficacy.15 Hence, it is important to identify the pathogenetic mechanism of FAP to find an alternative effective treatment strategy.
Accumulating studies focused on the TTR mutation gene and protein have provided insights into the pathogenesis of TTR FAP, including decreased stability of TTR tetramers, conformational change in the crystal structure of variant TTR, altered kinetics of denaturation, and disturbing endoplasmic ER quality control system.1,16–18 Previous studies have demonstrated that increased levels of ER stress are correlated with extracellular TTR deposition. Two ER stress markers, BiP and p-eIF2α, have been observed to be present and upregulated in the salivary gland tissue of FAP patients.3 However, the precise molecular mechanisms underlying TTR FAP and its phenotypic heterogeneity are not yet fully understood.
Our current study investigated whether the two mutations, TTR Y114C and TTR V30M, share the same pathogenesis and evaluated the effect of pathogenic mutations on the clearance of the monomer. Our results show that the ratio of LC3-II/I was markedly decreased, while BiP and p-eIF2α levels remained constant in TTR Y114C when compared with wild-type TTR and TTR 30M. The results of our research indicate the impaired autophagy contributed to the TTR Y114C mutation, but not ER stress. This observation indicates that abnormal accumulation of TTR caused by a different mutation might be cleared by different pathways, and more studies are necessary to confirm whether this difference applies to other TTR mutations.
Curcumin is known to have neuroprotective properties through a variety of mechanisms.8–11 Our research indicates that curcumin decreased the monomeric TTR by promoting autophagy, and without toxic effects. Moreover, this protective effect of curcumin on TTR Y114C could be partially blocked by 3-MA. Pullakhandam et al showed that curcumin binds to wild-type TTR and prevents urea-induced perturbations in the tertiary structure of TTR in vitro.19 Recently, Ferreira et al reported that dietary curcumin modulated TTR amyloidogenicity.20 Therefore, curcumin might be an effective therapy for FAP involving multiple molecular pathways.
Overall, our findings show that abnormal accumulation of TTR caused by different mutations might be cleared in different ways, and curcumin might be an effective therapy for FAP by promoting autophagy. Further studies are necessary to determine whether this phenomenon exists in other TTR mutations.
Stephen Williams, PhD
For PI3K and related inhibitors of PI3K/AKT/mTOR i would refer you to two people who should be in the discussion of this signaling pathway and PI3K/AKT inhibitors used for chemotherapy. The first is Dr. Mien-Chie Hung and the second is Dr. Gordon Mills. They both had been at MD Anderson and developed some of the first inhibitors as well as the earliest discoveries of overactivity of PI3K/AKT in ovarian cancer.
Next the field had never progressed any inhibitors past Stage II as there has been some serious toxicities seen in preclinical phases (most long term tox studies are done after patients are enrolled in phase I).
These will show how inhibitors of certain isoforms of PI3K (namely delta) had to be developed to circumvent some of the severe toxicity seen with the earliest inhibitors (wortmanin and LY294002.
Phosphatidylinositol 3-kinase (PI3K) inhibitors as cancer therapeuticshttp://www.ncbi.nlm.nih.gov/pmc/articles/PMC3843585/ will explain about some of the toxicities and describes the one PI3K that has made it to phase II
Most of them have failed and I believe now are being thought as an adjuvant not front line therapy
Lab Director at Emergency County Hospital Targu Jiu
In experimental models, disrupting the MDM2–p53
interaction restored p53 function and sensitized tumors to
chemotherapy or radiotherapy. (Kojima et al., 2005). This
strategy could be particularly beneficial in treating
cancers that do not harbor TP53 mutations. For example
in hematologic malignancies, such as multiple myeloma,
chronic lymphocytic leukemia (CLL), acute lymphoblastic
leukemia (ALL), acute myeloid leukemia (AML), and
Hodgkin’s disease, the induction of p53 – using a small
MDM2-inhibitor molecule, nutlin-3 – can induce the
apoptosis of malignant cells. Nutlins are a group of cisimidazoline
analogs, first identified by Vassilev et al.
(2004), which have a high binding potency and selectivity
for MDM2. Crystallization data have shown that nutlin-3
mimics the three residues of the helical region of the
trans-activation domain of p53 (Phe19, Trp23 and
Leu26), which are conserved across species and critical
for binding to MDM2 (Wade et al., 2010). Nutlin-3
displaces p53 by competing for MDM2 binding. It has
also been found that nutlin-3 potently induces apoptosis
in cell lines derived from hematologic malignancies,
including AML, myeloma, ALL, and B-cell CLL (Secchiero
et al., 2010).
Stephen J Williams, PhD
Now as far as PKM2 you would want to look at a company called Synta Pharmaceuticals and their inhibitor Elesclomal. elesclomol binds copper ions causing a change in conformation that enables its uptake through membranes and into cells. Elesclomol binds copper in an oxidative, positively charged state called Cu(II). Once inside mitochondria, the elesclomol-Cu(II) complex interacts with the energy production mechanism of the cell, or the electron transport chain. This interaction reduces the copper from Cu(II) to Cu(I), resulting in a cascade of reduction-oxidation, or redox, reactions, that causes a rapid increase of oxidative stress, disruption of mitochondrial energy production, and ultimately, triggering of the mitochondrial apoptosis pathway.
The important part is that it seemed, to prefer tumors which had lower LDH activity, meaning that these tumor cells actually did have a more active electron transport chain than tumors with high LDH (Warburg) and therefore in clinical trials the tumors with lower LDH activity responded more favorably.
The Wnt/β-catenin pathway plays a key role in liver development, regeneration and tumorigenesis. Among human cancers tightly linked to abnormal Wnt/β-catenin signaling, hepatoblastoma (HB) presents with the highest rate (50-90%) of β-catenin mutations. HB is the most common malignant tumor of the liver in childhood. This embryonic tumor differs from hepatocellular carcinoma by the absence of viral etiology and underlying liver disease, and by distinctive morphological patterns evoking hepatoblasts, the bipotent precursors of hepatocytes and cholangiocytes. Recent studies of the molecular pathogenesis of hepatoblastoma have led to identify two major tumor subclasses resembling early and late phases of prenatal liver development and presenting distinctive chromosomal alterations. It has been shown that the molecular signature of Wnt/β-catenin signaling in hepatoblastoma is mainly imposed by liver context, but differs according to developmental stage. Finally, the differentiation stage of tumor cells strongly influences their invasive and metastatic properties, therefore affecting clinical behavior.
7.10.1.2 Targeting the Wnt/β-Catenin Signaling Pathway in Liver Cancer Stem Cells and Hepatocellular Carcinoma Cell Lines with FH535
Hepatocellular carcinoma (HCC), the most common liver cancer, is the fifth most common cancer and the third highest cause of cancer-related mortality worldwide [1]–[2]. The alarming rise in HCC incidence in Europe and North America in recent years is related mainly to hepatitis C virus infection, although other factors such as excessive alcohol consumption and obesity also contribute to this increase [3]. The etiology of HCC is complex and involves numerous genetic and epigenetic alterations and the disruption of various signaling pathways including the Wnt/β-catenin, Ras/Raf/MAPK, PI3K/AKT/mTOR, HGF/c-MET, IGF, VEGF and PDGF pathways. Among these, the Wnt/β-catenin pathway is considered among the most difficult to inhibit [4]. Currently, few chemical agents targeting the Wnt/β-catenin pathway are available or under investigation [5].
Activation of the canonical Wnt/β-catenin pathway involves the binding of Wnt proteins to cell surface Frizzled receptors and LRP5/6 co-receptors. In the absence of Wnt proteins, much of the cellular β-catenin is bound to E-cadherin on the cell membrane. Cytosolic β-catenin is constitutively phosphorylated at specific serine residues by an enzymatic complex that includes adenomatous polyposis coli (APC), Axin, and the kinases glycogen synthase kinase-3β (GSK-3β) and casein kinase I, marking it for ubiquitin-mediated proteolysis. Under these conditions, the TCF/LEF transcription factors are bound to their cognate DNA recognition elements along with members of the Groucho family of co-repressors, insuring the transcriptional silencing of β-catenin target genes. Engagement of Wnt proteins with the Frizzled receptor activates the Dishevelled protein, resulting in the dissociation of the cytosolic destructive complex and inhibition of GSK-3β. This leads to the stabilization and accumulation of cytoplasmic β-catenin, which then enters the nucleus, binds TCF/LEF proteins and leads to the subsequent dissociation of groucho co-repressors, recruitment of the coactivator p300 and activation of β-catenin target genes [6]–[9]. Many of the β-catenin targets, including Cyclin D1, c-myc and Survivin, promote cell cycle progression and inhibit apoptosis [10]–[12]. Consistent with this data, activation of the Wnt/β-catenin pathway is seen in a variety of cancers, including HCC. Aberrant activation of the Wnt/β-catenin pathway has been observed in at least 1/3 of HCC, and roughly 20% of HCCs have mutations in the β-catenin gene. More than 50% of HCC tumors display nuclear accumulation of β-catenin indicating that other factors may be involved such as aberrant methylation of the tumor suppressors APC and E-cadherin, inactivation of casein kinase and GSK-3β, or increased secretion of Wnt ligants [4]–[5].
There has been increasing interest in the role of liver cancer stem cells (LCSC) in tumorigenesis, tumor progression, invasion and metastases. The cancer stem cell theory suggests that a tumor is comprised of a heterogeneous population of cells that form a distinct cellular hierarchy. Recent studies have provided convincing evidence that these cells do exist in solid tumors of many types including, brain, breast, colorectal, liver, pancreas and prostate cancers. In 2006, two different groups isolated a CD133+ subpopulation from HCC cell lines and described higher proliferative and tumorigenic potential, consistent with stem cell properties. CD44 was also found as an important marker used in combination with other stem cell markers to better define the surface phenotype of LCSC. It has been demonstrated that CD133+ and CD90+ cells co-expressing CD44+ are more aggressive than those expressing CD133 or CD90 alone [13]–[14].
The chemical agents used to target Wnt-/β-catenin pathway are at the membrane, cytosol and transcription factor levels [5]. The small molecular agent FH535 is a dual inhibitor of peroxisome proliferator-activated receptor (PPAR) and β-catenin/TCF/LEF. FH535 has been shown to inhibit proliferation of HCC and hepatoblastoma cell lines and its specificity on inhibition of β-catenin/TCF/LEF activity was illustrated in hepatoblastoma cell line HepG2 [15].
The aim of this study was to determine if FH535 can inhibit the activation of β-catenin-regulated genes by endogenous and ectopically expressed β-catenin in the HCC cell lines Huh7, Hep3B and PLC and liver cancer stem cells (LCSC). The specificity of FH535 on inhibition of β-catenin via TCF/LEF activation was assayed in dual luciferase reporter transfected in LCSC and in HCC cells. Proliferation, cell cycle, and other targeted genes and proteins were assayed.
FH535 inhibits transcriptional activation mediated by wild-type and constitutively active β-catenin
FH535 has been shown to block signaling through endogenous β-catenin in several cell lines, including the hepatoblastoma cell line HepG2 [15]. To further explore this regulation and to test whether FH535 could block ectopic β-catenin, co-transfections with β-catenin expression vectors and the TCF4-dependent luciferase reporter vector TOPFlash were performed in the human HCC cell lines Huh7 and Hep3B (Fig. 1). In both cell lines, co-transfected wild-type β-catenin expression vector increased luciferase activity from TOPFlash nearly 15-fold compared to cells co-transfected with the empty vector (E.V.) control. This β-catenin-dependent increase was inhibited by FH535 in a dose-dependent manner. β-catenin is often mutated in various cancers, including HCC. One natural mutation changes the serine at position 37; this altered form of β-catenin is resistant to degradation by the APC complex and thus has higher stability. To test whether this form of activated form of β-catenin could also be blocked by FH535, an expression vector for βCatS37A, in which the serine at position 37 has been changed to an alanine, was co-transfected with TOPFlash. As expected, βCatS37A-mediated transactivation of TOPFlash was significantly higher than transactivation by wild-type β-catenin. However, in both cell lines, βCatS37A-mediated transactivation was significantly inhibited by FH535. As controls, cells were also co-transfected with FOPFlash, which is identical to TOPFlash except that the TCF4 sites have been mutated and therefore no longer responsive to β-catenin; FOPFlash was not activated by wild-type β-catenin or βCatS37A as shown in Figure 1.
Huh7 (Panel A) and Hep3B (Panel B) HCC cells were transfected with the luciferase reporter genes TOPFlash (left panels), which contains three TCF binding sites, or E3-pGL3 (right panels), which contains the AFP enhancer element E3 that has a highly conserved TCF site. Cells were additionally co-transfected with an expression vector that contained no insert (empty vector control, E.V.), wild-type β-catenin (β-catenin), or a constitutively active form of β-catenin (βcatS37A). Renilla luciferase was used to control for variations in transfection efficiency. Six hours after the addition of DNA, cells were treated with DMSO alone (no treatment) or increasing amounts of FH535. After 48 hours, luciferase levels were determined; firefly luciferase was normalized to renilla. In both cell lines, FH535 inhibited β-catenin-dependent activation of target genes. *P<0.05. The experiment was done twice with similar results.
doi:10.1371/journal.pone.0099272.g001
TOPFlash contains three consensus TCF4 binding motifs that confer responsiveness to β-catenin. To test whether FH535 could also block β-catenin-mediated transactivation of a TCF4 motif in the context of a natural regulatory region, co-transfections were performed with E3-pGL3. E3 is a ~340 bp fragment that contains alpha-fetoprotein (AFP) enhancer element E3, one of three enhancers that control hepatic expression of the mouse AFP gene. E3 contains binding sites for multiple factors, including Foxa/HNF6, C/EBP, orphan nuclear receptors, and TCF4 [26]–[27]. We recently showed that this enhancer is regulated by β-catenin in cells and transgenic mice [21]. E3-pGL3 was transactivated by β-catenin and to a greater extent by βCatS37A (Fig. 1). However, this transactivation by both wild-type and S37A forms of β-catenin was blocked by FH535 in a dose-dependent manner.
3.2 FH535 inhibits β-catenin-mediated transcriptional activation in LCSC
Previous studies have shown that β-catenin signaling is elevated in EpCAM positive cells with LCSC properties [28]. We previously described that CD133+, CD44+, CD24+ LCSC aggressively form tumors when small numbers of these cells are injected into nude mice [29]. To test the ability of FH535 to inhibit β-catenin in these LCSCs, transient transfections were performed with TOPFlash. As controls, TOPFlash was also transfected into the HCC cell lines Huh7 and PLC (Fig. 2). In all three populations, untreated cells exhibited low luciferase levels. When treated with the GSK-3β inhibitor LiCl, which leads to endogenous β-catenin activation[30], TOPFlash activity increased dramatically. FH535 effectively blocked LiCl-mediated activation of TOPFlash in a dose-dependent manner. Interestingly, this inhibition was more robust in LCSC than in either HCC cell line. As a control, transfections were also performed with FOPFlash, which is no longer responsive to β-catenin. As expected, luciferase activity in FOPFlash-transfected cells was neither increased by LiCl nor inhibited by FH535.
Figure 2. FH535 inhibits TOPFlash activation in LCSC and HCC cell lines.
LCSC (left panel), Huh7 (middle panel) and HPLC (right panel) cells were co-transfected with TOPFlash or FOPFlash luciferase reporter genes along with renilla luciferase. After 6 hours, cells were left untreated (no treatment) or treated with LiCl alone or LiCl with increasing amounts of FH535. LiCl is a known activator of β-catenin. After an additional 36 hours, cells were harvested and luciferase levels were determined; firefly luciferase was normalized to renilla. TOPFlash activity was highly induced in all three cell populations; this activation was inhibited by FH535. The negative control FOPFlash showed minimal response to LiCl or FH535. TOPFlash inhibition by FH535 was more robust in LCSC than in either HCC cell line. * P<0.003, # P<0.001. The experiment was done twice with similar results.
doi:10.1371/journal.pone.0099272.g002
3.3 FH535 inhibits proliferation of LCSC and HCC cell lines
Numerous studies have demonstrated that β-catenin plays an important role in proliferation during normal development and in cellular transformation in many tissues, including the liver. Liver development is impaired in the absence of β-catenin, and mutations that activate the β-catenin pathway are found in about 1/3 of HCC [4]–[5]. Furthermore, the growth of adult liver progenitor stem cells (oval cells) can be inhibited by blocking the β-catenin pathway. Since our data indicated that FH535 can block β-catenin-mediated transcriptional activation, we also tested whether proliferation of LCSC and HCC cell lines was affected by this compound. LCSC were cultured in the presence of 10% or 1% serum and with between 5 µM and 30 µM FH535 for 72 hours, and cell proliferation was monitored by 3H-thymidine incorporation (Figs. 3A and 3B, respectively). Proliferation decreased with increasing amounts of FH535, with a more dramatic reduction observed in cells grown in the presence of lower serum; the concentration of FH535 to cause a 50% inhibition of cell grown (IC50) was 13.8 µM for cells grown in 10% serum and 5.1 µM for cells grown in 1% serum. This inhibition was more potent than that seen with XAV939 (IC50 = 55 µM), which inhibits tankyrase, thus stabilizing axin and promoting β-catenin degradation (Fig. 3C) [31]. FH535 also blocked proliferation of HCC cells at concentrations that were similar to that seen with LCSC (IC50 of 10.9 µM, 9.25 µM and 6.6 µM for Huh7, PLC and Hep3B, respectively; Fig.3D). To confirm that FH535 indeed inhibited cell proliferation and did not lead to increased cell death, FH535 and 3H-thymidine were added simultaneously to Huh7 cells, which were then cultured for 18 h. In this scenario, we observed a significant inhibition of proliferation at 2.5, 5, 10 and 15 µM of FH535 treatment as compared to control (p<0.05, n = 6), with FH535 at 15 µM causing a 41% inhibition (Figure S3). This data indicates that FH535 is inhibiting cell proliferation rather than increasing cell death.
Figure 3. FH535 inhibits proliferation of LCSC and HCC cell lines.
Cells were seeded in 96-well plates in 0.2 ml of media as described below for 72 hours, followed by the addition of 3H-thymidine at 1 µCi/well for 4 hours. Incorporation of 3H-thymidine was determined by scintillation counting. In panels A, B and D, the final concentration of DMSO in each well was 0.05%; in panel C, the final DMSO concentration in each well was 0.1%. (A) LCSCs were plated at 1000 cells/well in DMEM with 10% FBS along with DMSO alone or with increasing amounts of FH535. (B). LCSCs were plated at 5000 cells/well in DMEM with 1% FBS with DMSO alone or with increasing concentrations of FH535. (C). LCSCs were plated in DMEM with 10% FBS at 1000 cells/well with DMSO alone or increasing concentrations of XAV939. (D). Huh7, Hep3B and PLC cells were plated in DMEM with 10% FBS at 1000, 2500, and 5000 cells/well, respectively, with DMSO alone or increasing concentrations of FH535. Pvalues are for all the three cell lines treated with FH535 are compared to controls. The experiment was done twice with similar results.
doi:10.1371/journal.pone.0099272.g003
3.4 FH535 induces cell cycle arrest in the HCC cell line Huh7 and in LCSC
The ability of FH535 to inhibit cell proliferation prompted us to investigate the cell cycle distribution following treatment. Huh7 cells were synchronized by growth in 0.1% FBS for 24 hours and then cultured in the presence of 10% FBS and with no FH535 or FH535 at 7.5 µM and 15 µM. After 24 hours, cells were harvested and DNA content was analyzed by propidium iodide staining. In the presence of FH535, there was a statistically significant increase in the number of cells in G0/G1 and a corresponding decreased in the percentage of cells in S phase compared to cells grown in the absence of FH535 (Fig. 4A). The number of cells in G2 was not significantly altered by FH535. In addition, there was no sub-G1 peak detected by flow cytometry, indicating that FH535 was not promoting apoptosis at the concentrations being use (see Figure S4). We also did cell cycle analysis in LCSC after FH535 treatment and found FH535 at 15 µM significantly caused G1 phase arrest in LCSC (P = 0.012). FH535 also significantly reduced G2/M phase in the LCSC after 24 h of 7.5 µM and 15 µM FH535 treatment (P = 0.038 and P<0.001 respectively), no significant S phase inhibition was observed in LCSC (p = 0.446) (Fig. 4B.). Our data are similar to previously published results and reflects β-catenin regulation of cell cycle is different in different cell types [32]–[33]. Cell cycle regulators (cyclins, CDKs and regulators) can vary in different cell types, which could lead to different responses after FH535 treatment. This may worth exploring in our future study.
Figure 4. FH535 alters cell cycle progression in Huh7 and LCSC cells.
A. Huh7 cells were cultured in DMEM +10%FBS for 24 h. The cells were washed with serum free DMEM 3 times, then cultured in DMEM +0.1% FBS for 24 h for cell synchronization. Cells were then cultured in DMEM+10% FBS along with different concentrations of FH535 for 24 h. The cells were harvested and stained with propidium iodide (PI) and analyzed by flow cytometry according to the GenScript protocol (Piscataway, NJ, USA). Treatment with FH535 increased the percentage of cells in G1 and decreased the percentage of cells in S phase. The experiment was done twice with similar results. B. LCSC cells were cultured in CelProgen complete LCSC culture medium for 24 h. Cells were then washed with serum free CelProgen medium 3 times and cultured in CelProgen Medium +0.1% FBS for 24 h for synchronization of the cells. The cells were then returned to CelProgen Complete Medium +10% FBS with different concentrations of FH535 for 24 h. Cell cycle was assayed as per Huh7 described above.
doi:10.1371/journal.pone.0099272.g004
3.5 Expression of β-catenin target genes cyclin D1 and Survivin is inhibited by FH535
β-catenin controls cell proliferation and survival by regulating the expression of numerous targets genes. Two well-established targets are the genes encoding Survivin (Birc5) and Cyclin D1 (CcnD1). Survivin is an anti-apoptotic protein that also regulates progression through mitosis [34]; Cyclin D1 controls proliferation by activating the G1 kinases cdk4 and cdk6 [35]. Survivin and Cyclin D1 transcription are regulated through TCF elements in their promoter regions [36]. To test whether FH535 inhibits expression of these two β-catenin target genes, real-time RT-PCR was performed with LCSC and HCC cells that were treated with increasing amounts of FH535. Cyclin D1 and Survivin mRNA levels were reduced by FH535 in all three cell populations in a dose-dependent manner (Fig. 5). To confirm that this reduction in mRNA levels also led to lower protein levels, western analysis was performed using whole cell extracts from Huh7 cells. Both Cyclin D1 and Survivin protein levels were reduced in a dose-dependent manner, with the greatest reduction seen in the presence of 10 µM FH535 (Fig. 6.). Densitometric analysis indicated that FH535 at 5 and 10 µM inhibited Cyclin D1 28% and 64% respectively; FH535 at 5 and 10 µM inhibited surviving 24% and 48% respectively (Fig. 6).
Figure 5. FH535 reduces cyclin D1 and survivin mRNA levels in LCSC and in HCC cell lines.
LCSCs, Huh7 and Hep3B cells were treated with DMSO alone or increasing concentrations of FH535 for 38-time PCR for expression of Cyclin D1 (Panel A) or Survivin (Panel B). In both cases, mRNA levels were plotted relative to β2-microglobulin. The experiment was done twice with similar results.
Figure 6. FH535 reduces cyclin D1 and Survivin protein levels in Huh7 cells.
Huh7 cells were treated with DMSO alone or increasing amounts of FH535 for 38-PAGE, and transferred for Western analysis with antibodies against Cyclin D1, Survivin, and β-actin. The top of shows the western blot image; the bottom graph shows densitometric analysis of the western data. This densitometric analysis indicated that FH535 at 5 and 10 µM inhibited Cyclin D1 protein levels 28% and 64% respectively; FH535 at 5 and 10 µM inhibited Survivin protein levels 24% and 48% respectively. The experiment was done twice with similar results.
doi:10.1371/journal.pone.0099272.g006
Discussion
In recent years, numerous signaling pathways have been implicated in hepatic carcinogenesis. The β-catenin pathway is essential in stem cells for self-renewal and maintenance of stem cell properties. Disruption of this balance results in both genetic and epigenetic changes, found in many cancers, including colon cancer and HCC [4]. In this study, we used FH535 as an inhibitor of the β-catenin pathway. This compound has been used previously to inhibit β-catenin expression in cells from colon and lung as well as in cells from hepatoblastoma and HCC [15]. In this report, the authors concluded that FH535 was toxic to a number of cell lines, including Huh7. However, their assays could not distinguish between toxicity and reduced cell proliferation. Our data indicates that FH535 does indeed inhibit cell proliferation; we did not directly measure toxicity.
FH535 inhibition of LCSC proliferation is of interest due to its potential therapeutic effect in chemo-resistant HCC. Our group and others have focused on strategies to inhibit the proliferation of LCSC and differences in resistance patterns with non-liver cancer stem cell lines in vitro and in vivo.
Despite numerous efforts, the etiology of HCC tumorigenesis, whether transformed cells originate from mature hepatocytes or stem/progenitor cells remains unclear. Stem cells are defined by their potential for self-renewal and by their ability to proliferate and differentiate into diverse cell types [37]. In recent years, studies have provided convincing evidence that these cells do exist in solid tumors of many types including, brain, breast, colorectal, liver, pancreas and prostate cancers [27]. In this study we have used LCSC that are 64.4%, 83.2%, 96.4% and 96.9% positive, respectively, for CD133, CD44, CD24 and Aldehyde A1 as determined by flow cytometry. These cells have been previously profiled not only by checking the LCSC markers but also by evaluating their tumorigenic potential using low cell numbers (using 2000 LCSCs instead of 100,000 HCC cells to generate tumors) and studying resistance to several drugs. We previously found that these LCSC have intermediate to high resistance to drugs compare to non- liver cancer stem cell lines using different inhibitors.
In this study, we found that FH535, LCSC inhibition of proliferation was affected by FBS concentration in the culture medium, suggesting that the PPAR pathway may be involved in LCSC proliferation as found in the human cancer cell line HCT116 [15]. This could be explained by a variety of fatty acids and their derivatives present in the FBS that are natural agonists to PPAR. It is possible the PPAR agonists suppress the inhibitory effects of FH535 in cell culture. Indeed, in HCT116 cells, FH535 inhibition of β-catein/TCF-dependent luciferase reporter genes was five times stronger in serum-free medium than in media containing 10% FBS. The ability of FH535 to inhibit tumor growth was dramatically increased when 10% FBS was replaced with 10% BSA [15]. Lysophosphatidic acid was found to be an effective PPAR agonist that could reverse FH535 induced inhibition of HCT116 growth [15]. However, the potential function of PPAR in LCSC is beyond the scope of this study and needs further investigation. Recently, FH535 was found to be the most potent drug among several other Wnt/β-catenin inhibitors on human biliary tract cancer cells cultured in serum-free medium [38]. Our study found that FH535 is much more potent than XAV939 in 10%FBS DMEM. This may be related to the PPAR inhibition potential of FH535. Our study found that FH535 inhibited HCC cell lines Huh7, Hep3B and PLC proliferation, indicating that Wnt/β-catenin signaling plays an important role not only in LCSC but also in HCC.
FH535 inhibition of LCSC and HCC proliferation was illustrated by its ability to inhibit β-catenin/TCF/LEF-dependent luciferase reporter activity. To our knowledge, this is the first report on the ability of FH535 to inhibit β-catenin/TCF/LEF activity in LCSC and in HCC cell lines. Previously, Handeli and Simon reported that FH535 inhibits β-catenin/TCF/LEF activity in the HepG2 cell line, which was mistakenly labeled as HCC by these authors [15]. For over thirty years this cell line was considered HCC by numerous investigators. Lopez et al., who initially isolated these cells, recently concluded that HepG2 cells should in fact be considered a hepatoblastoma cell line [39]. Further studies will be needed to investigate how FH535 inhibition of β-catenin influences LCSCs and HCCs. As shown here, cyclin D1 and Survivin expression are inhibited by FH535. Survivin is an anti-apoptotic protein that also regulates progression through mitosis [26], whereas Cyclin D1 controls proliferation by activating the G1 kinases [35]. Real-time RT-PCR and Western analysis confirmed that the expression of these target genes was evident at the mRNA and protein level. Our preliminary data indicate that FH535 treatment does not alter CD133, CD13 and EPCAM expression in LCSC and HCC cell lines (data not shown). Further analysis of these and other stem cell markers are warranted.
In conclusion, our data show that FH535 is a potent inhibitor of the Wnt/β-catenin pathway in LCSCs and HCC cell lines. Whether its ability to inhibit PPAR also affects the growth of LCSCs and HCC cells will require further investigation. Further studies will also be needed to investigate the in vivo efficacy and toxicity of FH535 on HCC xenografts in an animal model. The role of combination therapy using FH535 with other anti-HCC drugs and the possibility of finding cross-talk of Wnt/β-catenin pathway with other signaling pathways should be investigated.
7.10.1.3 Wnt signaling in hepatocellular carcinoma: analysis of mutation and expression of beta-catenin, T-cell factor-4 and glycogen synthase kinase 3-beta genes.
Hepatocellular carcinoma (HCC) is a common killer cancer in the world. Recently, abnormal activation of the Wnt pathway has been found to be involved in the carcinogenesis of several human cancers including HCC. The goal of the present study was to investigate the mechanism of inappropriate activation of the Wnt pathway in hepatocarcinogenesis. We analyzed the alterations of three key components of the Wnt pathway: beta-catenin, glycogen synthase kinase (GSK)-3beta and T-cell factor (Tcf)-4 in 34 HCC and paracancerous normal liver by immunohistochemistry, polymerase chain reaction (PCR)-single-strand conformation polymorphism (SSCP), direct sequencing, and quantitative real-time reverse transcription (RT)-PCR. We found that 61.8% (21/34) of all HCC examined showed an abnormal beta-catenin protein accumulation in the cytoplasm or nuclei. The RT-PCR-SSCP and direct sequencing showed that beta-catenin exon 3 mutations existed in 44.1% (15/34) of the HCC. No mutations of GSK-3beta or Tcf-4 were detected in HCC. Moreover, messenger RNA of beta-catenin and Tcf-4, but not GSK-3beta, was found to be overexpressed in HCC. On analyzing the relationship between alterations of beta-catenin or Tcf-4 and C-myc or Cyclin D1 expression, we found that mutations of beta-catenin, as well as overexpression of beta-catenin or the Tcf-4 gene were independently correlated with C-myc gene overexpression in HCC. Our present findings strongly suggest that mutations of beta-catenin, as well as overexpression of beta-catenin and the Tcf-4 gene, independently activate the Wnt pathway in HCC, with the target gene most likely to be C-myc.
The regulation of cell growth and survival can be subverted by a variety of genetic defects that alter transcriptional programs normally responsible for controlling cell number. High throughput analysis of these gene expression patterns should ultimately lead to the identification of minimal expression profiles that will serve as common denominators in assigning a cancer to a given category. In the course of defining the common denominators, though, we should not be too surprised to find that cancers within a single category may nevertheless exhibit seemingly disparate genetic defects. The wnt pathway has already provided an outstanding example of this. We now know of three regulatory genes in this pathway that are mutated in primary human cancers and several others that promote experimental cancers in rodents (Fig. 1). In all of these cases the common denominator is the activation of gene transcription by β-catenin. The resulting gene expression profile should provide us with a signature common to those cancers carrying defects in the wnt pathway. In this review, the wnt pathway will be covered from the perspective of cancer, with emphasis placed on molecular defects known to promote neoplastic transformation in humans and in animal models.
Oncogenes and tumor suppressors in the wnt signaling pathway. Lines ending with arrows or bars indicate activating or inhibitory effects, respectively. Green and red indicate proto-oncogenic and tumor suppressive activity, respectively, in human cancer or transgenic animals. Definition of the genes and the basis for their activities are described in the text.
The wnt signaling mechanism
The model illustrated in Figure 2 is a proposed mechanism for wnt signaling and is based on the following literature. Signaling is initiated by the secreted wnt proteins, which bind to a class of seven-pass transmembrane receptors encoded by the frizzled genes (Bhanot et al. 1996; Yang-Snyder et al. 1996; He et al. 1997). Activation of the receptor leads to the phosphorylation of the dishevelled protein which, through its association with axin, prevents glycogen synthase kinase 3β (GSK3β) from phosphorylating critical substrates (Itoh et al. 1998; Kishida et al. 1999; Lee et al. 1999; Peters et al. 1999; Smalley et al. 1999). In vertebrates, the inactivation of GSK3β might result from its interaction with Frat-1 (Thomas et al. 1999; Yost et al. 1998; Li et al. 1999a; Salic et al. 2000). The GSK3β substrates include the negative regulators axin and APC, as well as β-catenin itself (Rubinfeld et al. 1996; Yost et al. 1996; Yamamoto et al. 1999). Unphosphorylated β-catenin escapes recognition by β-TRCP, a component of an E3 ubiquitin ligase, and translocates to the nucleus where it engages transcription factors such as TCF and LEF (Behrens et al. 1996; Molenaar et al. 1996;Hart et al. 1999). Additional components in the pathway include casein kinases I and II, both of which have been proposed to phosphorylate dishevelled (Sakanaka et al. 1999; Willert et al. 1997; Peters et al. 1999). The serine/threonine phosphatase PP2A associates with axin and APC, although its functional role in the pathway remains obscure (Hsu et al. 1999; Seeling et al. 1999). Also obscure is the manner by which the wnt receptors communicate with dishevelled.
Proposed mechanism for the transmission of wnt signals. In the absence of wnt –wnt) GSK3β phosphorylates APC and axin, increasing their binding affinities for β-catenin, which too is phosphorylated by GSK3β, marking it for destruction. In the presence of wnt (+wnt) FRAT prevents GSK3β from phosphorylating its substrates, and β-catenin is stabilized. Casein kinase1ε (CK1ε) binds to and phosphorylates dishevelled (dvl) modulating the FRAT1/GSK3β interaction. RGS, PDZ, and DIX are protein interaction domains.
Receptors, ligands, and related proteins
The proto-oncogenic effects of wnt were discovered over 18 years ago inciting intense investigation into the role of wnt genes in human cancer (Nusse and Varmus 1982). The subsequent discovery of wingless, the fly homolog of wnt-1, paved the way for assembling a signaling pathway subsequently found to contain cancer causing genes (Cabrera et al. 1987; Rijsewijk et al. 1987). Although wnt was the prototypical oncogene in this pathway, no formal proof for its involvement in human cancer has ever been documented. There have been numerous reports on the overexpression, and sometimes underexpression, of wnt genes in human cancers, but mRNA expression levels are merely correlative. More compelling evidence, such as amplification, rearrangement, or mutation of genes encoding wnt ligands or receptors has not been forthcoming. In lieu of these sorts of findings, we are left to speculate on the consequences of epigenetic events implicating these genes in human cancer. In doing so we can use animal and cell culture models to guide our interpretation.
The wnt ligands, of which there are at least 16 members in vertebrates, are secreted glycoproteins that can be loosely categorized according to their ability to promote neoplastic transformation (for review, seeWodarz and Nusse 1998). For example, the activation of wnt-1, wnt-3, or wnt-10b by retroviral insertion in the mammary gland will promote tumor formation in mice (Lee et al. 1995; Nusse and Varmus 1982; Roelink et al. 1990). Oncogenic potential can also be assessed in cultured mammalian cells, such as C57MG and CH310T1/2, where expression of the proto-oncogenic wnts results in morphological transformation (Bradbury et al. 1994; Wong et al. 1994). These cells are transformed by wnt-1, wnt-2, wnt3a but not by wnt-4, wnt-5a, and wnt-6. The transforming wnt genes also promote the accumulation of β-catenin in some cultured mammalian cells (Shimizu et al. 1997). Some aspects of the wnt cancer pathway are also recapitulated inXenopusdevelopment, where injection of transforming wnts into early embryos results in duplication of the dorsal axis (Wodarz and Nusse 1998). A caveat here is that the lack of specific receptors for certain wnts might also explain their inactivity in some of these assays (He et al. 1997). Nevertheless, identifying those wnts capable of neoplastic transformation will aid the interpretation of epigenetic evidence implicating wnts in cancer. For example, expression of thewnt-16 gene is activated by the E2A–Pbx1 fusion product in acute lymphoblastoid leukemia (McWhirter et al. 1999), but the oncogenic potential of wnt-16 is unknown.
As might be expected from the plethora of wnt genes, there are also numerous wnt receptors. At least 11 vertebrate frizzled genes have been identified, but how they differ in function and ligand specificity is far from clear. The analysis of mere binding specificity may not be sufficient to sort out the appropriate combinations of functional receptor-ligand interactions. Wnt-3a and wnt-5a both bind to Human frizzled 1 (Hfz1), yet only wnt-3a mediates TCF-dependent transcription (Gazit et al. 1999). This suggests that the activation of TCF/LEF-dependent transcription is a good correlate to neoplastic transformation. Implementation of this assay, along with a second assay involving the translocation of PKC to the cell membrane, resulted in the categorization of murine wnt receptors into two exclusive groups (Sheldahl et al. 1999). Human FzE3 fell into the TCF/LEF activation group, consistent with previous work showing that its overexpression resulted in nuclear localization of β-catenin (Tanaka et al. 1998). This receptor was also expressed in numerous human esophageal cancers, but not in matched normal tissue (Tanaka et al. 1998).
In addition to the frizzled receptors, there exists a family of secreted proteins bearing homology to the extracellular cysteine-rich domain of frizzled. The so-called secreted frizzled-related proteins (sFRP) bind to the wnt ligands, thereby exerting antagonistic activity when overexpressed in wnt signaling assays (Leyns et al. 1997; Wang et al. 1997). The vertebrate sFRPs, like the frizzled proteins, exhibit functional specificity with respect to the various wnts. InXenopus assays, the prototypical frizzled related protein frzb, now known as sFRP-3, inhibited wnt-1 and wnt-8, but not wnt-5a (Leyns et al. 1997; Lin et al. 1997; Wang et al. 1997). Assays in mammalian cells showed that FrzA, now termed sFRP-1, inhibited wnt-1-induced accumulation of β-catenin (Dennis et al. 1999;Melkonyan et al. 1997). Again, binding specificity may not relate to functional specificity, as wnt-5a associated with sFRP-3 but was unable to inhibit its activity (Lin et al. 1997). Even the significance of specific functional interactions might be suspect based on recent titration experiments with purified soluble sFRP-1. At low concentrations sFRP-1 enhanced signaling activity by soluble wingless protein, whereas at higher concentrations it was inhibitory (Uren et al. 2000). The authors proposed high and low states of binding affinity that involved the carboxy-terminal heparin binding domain and the amino-terminal cysteine-rich domain of sFRP-1, respectively. Binding to the cysteine-rich domain might confer inhibition while binding to the carboxy-terminal region could facilitate presentation of active ligand to receptor. The potential for some sFRPs to activate wnt signaling is consistent with a previous study in which sFRP-2, then known as SARP-1, increased the intracellular concentration of β-catenin and conferred anti-apoptotic properties to cultured MCF-7 cells (Melkonyan et al. 1997). Functional studies are further complicated by the binding of a sFRP to the putative human receptor frizzled-6, underscoring additional possible modes of regulation (Bafico et al. 1999). The sFRPs have not been directly linked to cancer, but one could speculate that the anti-apoptotic activity observed with the SARP-1 could contribute to tumor progression. Alternatively, the identification of sFRP-2 as a target of the hedgehog signaling pathway might be relevant to human basal cell cancers (Lee et al. 2000). Additional structurally distinct secreted inhibitors of wnt signaling include the recently discovered dickopft-1 and wif-1 proteins (Fedi et al. 1999; Glinka et al. 1998;Hsieh et al. 1999).
GSK3β
The serine/threonine kinase GSK3β binds to and phosphorylates several proteins in the wnt pathway and is instrumental to the down regulation of β-catenin (Dominguez et al. 1995; He et al. 1995; Hedgepeth et al. 1999b; Ikeda et al. 1998;Itoh et al. 1998;Li et al. 1999a; Nakamura et al. 1998b; Rubinfeld et al. 1996;Yamamoto et al. 1999; Yost et al. 1996). As a negative regulator of wnt signaling, GSK3β would qualify as a potential tumor suppressor. However, mutations or deletions in the gene coding for GSK3β were not been detect ed in a survey of colorectal tumors (Sparks et al. 1998). Perhaps GSK3β can compensate for the loss of GSK3β and the biallelic inactivation of both these genes is unlikely in tumor progression. Alternatively, the utilization of GSK3β by pathways independent of wnt could make its overall ablation incompatible with cell viability. Nevertheless, inactivation of GSK3β can still be achieved by a means other than genetic ablation and can occur in a manner that uniquely affects wnt signaling. This mode of inactivation involves the association of GSK3β with Frat-1. Frat-1 was identified by insertional mutagenesis in a screen for genes that enhanced the progression of transplanted T-cell lymphomas in mice (Jonkers et al. 1997). Subsequent transgenic expression of Frat-1 alone did not induce spontaneous lymphomas, but greatly enhanced lymphomagenesis initiated either by leukemia virus M-MuLV or expression of the Pim1 oncogene (Jonkers et al. 1999). A connection to GSK3β was realized by the discovery of the Frat-1 Xenopushomolog GBP, a GSK3β binding protein inhibitory to wnt signaling when expressed in Xenopus embryos (Yost et al. 1998). Frat-1 is also antagonistic to wnt signaling in mammalian cells, presumably because it competes with axin for binding to GSK3β (Li et al. 1999a; Thomas et al. 1999). GBP also inhibited the phosphorylation and degradation of β-catenin in vitro when added to Xenopusextracts (Salic et al. 2000). Although Frat-1 contributes to cancer progression in a transgenic mouse model, its contribution to human cancer has not been documented.
Dishevelled
The genetic analysis of dishevelled in developmental systems has defined it as a positive mediator of wnt signaling positioned downstream of the receptor and upstream of β-catenin (Noordermeer et al. 1994). Overexpression or constitutive activation of dishevelled would be expected to promote neoplastic transformation, but its involvement in human cancers has not been reported. This might reflect the dual function of dishevelled, one that transduces wnt signals for the stabilization of β-catenin and a second that relays signals for the activation of jun kinases (Li et al. 1999b; Moriguchi et al. 1999). Although these two functions are housed in physically separable regions of the protein, dysregulation of one function, without impacting the other, could place severe constraints on selection for potential oncogenic mutations. A possible connection of dishevelled to cancer is through casein kinase II, which binds to and phosphorylates dishevelled and also promotes the formation of lymphomas when expressed in transgenic mice (Seldin and Leder 1995; Song et al. 2000; Willert et al. 1997).
β-catenin
Mutations in the β-catenin gene (CTNNb1) affecting the amino-terminal region of the protein make it refractory to regulation by APC (Morin et al. 1997; Rubinfeld et al. 1997). These mutations affect specific serine and threonine residues, and amino acids adjacent to them, that are essential for the targeted degradation of β-catenin (for review, see Polakis 1999). The mutations abrogate the phosphorylation dependent interaction of β-catenin with β-TRCP, a component of an E3 ubiquitin ligase that makes direct contact with amino terminal sequence in β-catenin (Hart et al. 1999). This regulatory sequence in β-catenin is mutated in a wide variety of human cancers as well as in chemically and genetically induced animal tumors. Importantly, β-catenin mutations in tumors are exclusive to those that inactivate APC. This is particularly apparent in colorectal cancer where the vast majority of these tumors contain APC mutations and the overall frequency of β-catenin mutations is quite low (Samowitz et al. 1999; Sparks et al. 1998;Kitaeva et al. 1997) (Table 1). When colorectal tumors lacking APC mutations were analyzed separately, the likelihood of finding a CTNNb1 mutation was greatly increased (Iwao et al. 1998; Sparks et al. 1998). The exclusivity of CTNNb1 and APC mutations in colorectal cancer was also evident from the analysis of replication error-positive tumors identified by microsatellite instability. Both the hereditary and sporadic forms of replication error-positive colorectal cancers had a relatively high frequency of β-catenin mutations, whereas APC mutations were relatively rare (Mirabelli-Primdahl et al. 1999; Miyaki et al. 1999) (Table 1). Interestingly, this correlation between microsatellite instability andCTNNb1 mutations was not apparent in endometrial cancers (Mirabelli-Primdahl et al. 1999).
Aggressive fibromatosis, otherwise known as desmoid tumor, is a locally invasive fibrocytic growth that occurs with increased incidence in patients with familial adenomatous polyposis coli (FAP). FAP individuals carry APC mutations in their germline and present with multiple intestinal adenomas at an early age. Desmoids also occur sporadically and, with the exception of colorectal cancer, represent a rare example of biallelic inactivation of APC in individuals without a pre-existing germline mutation in APC (Alman et al. 1997). Not surprisingly, mutations inCTNNb1 have also been detected in sporadic desmoid tumors (Shitoh et al. 1999;Tejpar et al. 1999). The β-catenin mutations were found in over half of the 42 desmoids analyzed, while inactivating mutations in APC were detected in nine and, again, there was no overlap between APC and β-catenin mutations (Tejpar et al. 1999). The β-catenin mutations were all of the missense variety and were confined to codons 41 and 45. Some of the desmoids lacked mutations in either β-catenin or APC, but all displayed increased expression of β-catenin, implying that yet unidentified defects in β-catenin regulation exist in some of these tumors.
There appears to be a low probability of accruing biallelic inactivating mutations in APC in most sporadic cancers, despite increased cancer incidence at numerous extracolonic sites in FAP patients. This suggests that the stabilization of β-catenin can promote cancer in many tissue types, but the biallelic inactivation of APC is an unlikely means to this end. Components in the wnt pathway other than APC, such as β-catenin, might make easier targets for oncogenic mutations. Indeed, several mutations in CTNNb1 were recently identified in gastric cancers, which occur with increased incidence in FAP patients (Park et al. 1999). In this study, 27% of intestinal type gastric cancers harbored mutations in β-catenin. Hepatoblastoma also occurs with increased incidence in FAP individuals (Hughes and Michels 1992;Giardiello et al. 1996; Cetta et al. 1997), but biallelic inactivation of APC is uncommon in the sporadic forms of these tumors. In three separate studies, mutations in β-catenin were identified at high frequency in hepatoblastoma, while no APC mutations were found (Koch et al. 1999; Jeng et al. 2000; Wei et al. 2000). Hepatoblastoma is also associated with Beckwidth–Wiedemann syndrome (BWS), however, a direct link between wnt signaling and the genetic defects underlying BWS are unlikely as a tumor from one of these patients also contained a somatic mutation in β-catenin (Wei et al. 2000). By contrast, a subset of patients with Turcot’s syndrome harbor germline mutations in APC and are at increased risk of medulloblastoma (Hamilton et al. 1995; Lasser et al. 1994). Although inactivating mutations in APC have not been detected in the sporadic forms of medulloblastoma, CTNNb1mutations were found in a small percentage (Zurawel et al. 1998).
Hepatocellular carcinoma (HCC) has become one of the most common tumors harboring mutations in the wnt pathway. Based on five separate studies, the frequency of CTNNb1 mutations in hepatocellular carcinoma (HCC) was ∼20% overall and perhaps higher still for HCCs associated with hepatitis C virus (de La Coste et al. 1998; Miyoshi et al. 1998;Huang et al. 1999; Legoix et al. 1999; Van Nhieu et al. 1999) (Table1). Preliminary data indicated a poorer prognosis associated with nuclear accumulation of β-catenin in HCC and histological data indicated enhanced nuclear staining in the invasive and intravascular compartments of the tumors (Huang et al. 1999; Van Nhieu et al. 1999). In one of these studies an inverse correlation between β-catenin mutations and loss of heterozygosity in the genome was noted (Legoix et al. 1999). This suggests that chromosomal instability and mutations inCTNNb1 represent alternative modes of tumor progression in HCC.
It is noteworthy that c-myc and cyclin D genes are amplified in a subset of HCCs and both these genes are downstream targets of β-catenin (He et al. 1998; Nishida et al. 1994; Peng et al. 1993;Shtutman et al. 1999; Tetsu and McCormick 1999). It would be of interest to determine whether any overlap exists between their amplification and CTNNb1mutations in HCC. Animal models of HCC have provided some clues toward understanding the relationship between these genes in cancer. HCCs induced by transgenic expression of SV40 T antigen in murine liver did not contain mutations in CTNNb1 (Umeda 2000). As T antigen activates cyclin D kinase by sequestration of Rb, the activation of the cyclin D gene by mutant β-catenin may no longer be required. By contrast, activating mutations inCTNNb1 were identified in half of the HCCs generated by transgenic expression of c-myc in murine liver (de La Coste et al. 1998). This animal model suggests that β-catenin mutations occur as a second “hit” in HCC tumor progression in cooperation with a distinct cancer pathway initiated by c-myc. That CTNNb1mutations can occur subsequent to other oncogenic defects is also evident from their occurrence in Wilm’s tumor. Mutations in β-catenin were detected in 15% of these pediatric kidney cancers and in two of these cases they were concomitant with mutations in the Wilm’s tumor gene WT1 (Koesters et al. 1999). One of these cases was associated with Denys-Drash syndrome, a familial disorder attributable to germline mutations in WT1.
It makes sense that extracolonic tumors associated with FAP, such as desmoids, medulloblastoma, and HCC, would contain CTNNb1mutations in their sporadic forms. Thyroid cancers also occur with increased incidence in FAP and, not surprisingly, a high frequency ofCTNNb1 mutations was recently reported for anaplastic thyroid cancers (Cetta et al. 2000; Garcia-Rostan et al. 1999). Although many of these mutations affected amino acids known to influence the regulation of β-catenin, many of them affected residues for which the consequence of their mutation is unknown (Garcia-Rostan et al. 1999). In particular, the substitution K49R was detected nine times. This mutation was frequently detected in the context of independentCTNNb1 mutations in the same thyroid tumor, and up to four independent CTNNb1 mutations were found in some tumors. The occurrence of multiple independent CTNNb1 mutations was also noted in some HCCs and might reflect the multifocal origin of some cancers (Huang et al. 1999; Legoix et al. 1999; Van Nhieu et al. 1999). In one HCC study, examination of different tumor areas from the same patient revealed distinct CTTNb1 mutations in two independent cases (Huang et al. 1999).
Some cancers, such as endometrial ovarian tumors, do not occur with increased incidence in patients with FAP, yet they contain activating mutations in CTNNb1(Palacios and Gamallo 1998; Gamallo et al. 1999; Wright et al. 1999). Perhaps inactivation of the remaining wild-type APC allele in FAP individuals is unlikely in this tissue, or the expression of an alternative APC gene compensates for its loss. The CTNNb1 mutations associated with ovarian cancer appeared to be confined to the endometrioid subtype. In this tissue, cancers with activated β-catenin signaling were reported to be less aggressive than their nonactivated counterparts. In one report, a more favorable prognosis was associated with cancers exhibiting enhanced nuclear staining of β-catenin and another indicated higher frequency ofCTNNb1 mutations in lower grade tumors (Palacios and Gamallo 1998; Wright et al. 1999). A similar inverse correlation between tumor grade and occurrence ofCTNNb1 mutations was also reported for uterine endometrial cancers (Fukuchi et al. 1998). The overlap between mutations in CTNNb1 and other gene defects in ovarian cancers has not been explored in detail, although one study noted coexisting mutations in the PTEN tumor suppressor andCTNNb1 in endometrioid tumors (Wright et al. 1999).
Additional types of cancers with CTNNb1 mutations, albeit at low frequency, include melanoma and prostate. Although only one of sixty-five melanomas contained detectable mutations, nuclear localization of the protein was seen in one-third (Rimm et al. 1999). Thus, additional mechanisms for β-catenin activation likely occur in these tumors. Possibly the highest percentage ofCTNNb1mutations occurs in a common skin tumor known as pilomatricomas (Chan et al. 1999). That these tumors might contain CTNNb1 mutations was surmised from the genesis of similar tumors in transgenic mice expressing mutant β-catenin in the skin (Gat et al. 1998). The tumors appeared to originate from the hair follicle, which is consistent with the lack of hair in mice homozygous for mutations in LEF, a transcription factor responsive to β-catenin (van Genderen et al. 1994).
Axin
Axin was originally identified as an inhibitor of wnt signaling inXenopus embryos and was subsequently shown to bind directly to APC, β-catenin, GSK3β and dishevelled (for review, see Peifer and Polakis 2000). A plethora of in vitro and in vivo studies inXenopus, Drosophila, and cultured mammalian cells has demonstrated that axin is central to the down regulation of β-catenin (Zeng et al. 1997; Behrens et al. 1998; Hart et al. 1998;Ikeda et al. 1998; Nakamura et al. 1998a; Sakanaka et al. 1998; Fagotto et al. 1999; Hedgepeth et al. 1999a; Li et al. 1999a; Willert et al. 1999a; Farr et al. 2000). It is not entirely clear how axin functions, but it has been proposed to facilitate the phosphorylation of β-catenin and APC by GSK3β (Hart et al. 1998; Ikeda et al. 1998). Thus axin would be viewed as a tumor suppressor based on its ability to downregulate signaling, and this has now been verified by documentation of its biallelic inactivation in human hepatocellular cancers and cell lines (Satoh et al. 2000). Importantly, these mutations were identified in those HCCs that lacked activating mutations inCTNNb1. All of the mutations were predicted to truncate the axin protein in a manner that eliminated the β-catenin binding sites. Axin, which should now be regarded as a tumor suppressor, constitutes the third genetic defect in the wnt pathway that contributes to human cancer. There also exists a close homolog of axin termed conductin, which exhibits of all the binding and regulatory functions of axin (Behrens et al. 1998). That this apparent redundancy did not suppress axin mutations in HCC suggests conductin is either not functionally equivalent to axin or not expressed at levels sufficient to compensate for its loss in HCCs.
PP2A
The dependence upon serine/threonine kinases for the regulation of β-catenin implies that phosphatases are also involved. Indeed, the rapid dephosphorylation of the axin protein is a consequence of wnt signaling and has been proposed to both destabilize axin and reduce its affinity for β-catenin (Willert et al. 1999b;Yamamoto et al. 1999). Although axin binds directly to the PP2A catalytic subunit, the phosphatase affecting axin in response to wnt signaling has not been identified (Hsu et al. 1999). If PP2A is this phosphatase, it would be viewed as proto-oncogenic because it downregulates the tumor suppressor axin. On the contrary, expression of the PP2A regulatory subunit B56 in human colon cancer cells results in the downregulation of β-catenin, consistent with a tumor suppressive function in the wnt pathway (Seeling et al. 1999). Moreover, the beta isoform of the PP2A A subunit is deleted in some human colon tumors, again implying tumor suppression (Wang et al. 1998). Also, disruption of twins, aDrosophila gene coding for a PP2A subunit, complemented the overexpression and underexpression of the β-catenin homolog armadillo, in a manner consistent with negative regulation of wnt signaling (Greaves et al. 1999). By all accounts, PP2A plays a role in wnt signaling, but its potential role as proto-oncogene or tumor suppressor might be dependent upon the precise nature of the defect.
APC
Genetic analysis of FAP families led to the identification of theAPC gene, and subsequent studies demonstrating an interaction with β-catenin placed it tentatively in the wnt pathway (Groden et al. 1991; Kinzler et al. 1991; Munemitsu et al. 1995; Rubinfeld et al. 1993; Su et al. 1993). Experiments in Drosophilaultimately revealed that genetic ablation of APC indeed resulted in upregulation of β-catenin signaling (Ahmed et al. 1998). In some systems, such as Xenopus andCaenorhabditis elegans, a positive role for APC in the wnt pathway has been proposed, but the former study suffers from potential overexpression artifacts and the latter from a lack of relatedness to the vertebrate APC protein (Rocheleau et al. 1997; Vleminckx et al. 1997). In any case, APC is a tumor suppressor in human cancers and its mutation relates strongly to the regulation of β-catenin. The spectrum of APC mutations, which typically truncate the protein, suggest selection against β-catenin regulatory domains, albeit not necessarily against β-catenin binding (for review, see Polakis 1999). The selective pressure might be directed against the presence of Axin binding sites, of which there are three, dispersed across the central region of the APC protein (Behrens et al. 1998). The presence of axin binding sites are critical to APC in the regulation of β-catenin levels and signaling in cultured cells (Kawahara et al. 2000). Moreover, mice lacking wild-type APC but expressing a truncated mutant APC retaining a single axin binding site are viable and do not develop intestinal neoplasia (Smits et al. 1999). This has not been the case for mice with more extensive truncations in APC (Oshima et al. 1995a; Su et al. 1992). Also, milder forms of colorectal polyposis, as well as familial infiltrative fibromatosis (desmoid tumors), have been associated with germline mutations in the 3′ region of the APC open reading frame. These mutations predict truncated proteins that retain only one or two of the three axin binding sites in APC (Walon et al. 1997; Kartheuser et al. 1999; Scott et al. 1996;van der Luijt et al. 1996). A recent study has also demonstrated that the expression of just the central region of APC, which contains all of the axin and β-catenin binding sites, was sufficient to elicit cellular growth suppression (Shih et al. 2000). This effect is consistent with previous work showing that a like fragment of APC was sufficient to downregulate β-catenin levels in cancer cells (Munemitsu et al. 1995).
Although both copies of the APC gene are typically inactivated in colorectal cancers, it remains possible that a mutant truncated APC could contribute to cancer progression. This was tested by transgenic expression of two different APC mutants in a wild-type intestinal background (Oshima et al. 1995b). This did not result in cancer-prone mice, despite high levels of expression of mutant proteins and, therefore, argues against a dominant negative effect by these particular mutants. However, it does not rule out an additive contribution to tumor progression by mutant APC protein in a background lacking wildtype APC. In fact, genetic evidence argues in favor of selection for a somewhat specific mutant APC protein. The mutation cluster region (MCR) in APC, roughly defined by codons 1250–1500, is not only consistent with selection against specific sequence, but also retention of an APC molecule that extends into the MCR (Fig.3.). A correlation between the presence of a germline mutation in the MCR and the severity of polyposis has been noted (Ficari et al. 2000; Nagase et al. 1992; Wu et al. 1998). The enhanced severity of polyposis suggests there should also be selective pressure for somatic mutations in the MCR, which indeed appears to be the case (Miyoshi et al. 1992). Selective pressure for an MCR mutant has also been proposed based on the occurrence of somatic mutations in the MCR relative to the position of the germline mutation in FAP (Lamlum et al. 1999). Tumors from FAP patients with a germline MCR mutation exhibited frequent inactivation of the remaining APC allele by LOH, while those without a germline MCR mutation had frequent somatic mutations in the MCR (Fig. 3). Therefore, a germline mutation in the MCR could relieve the constraint for a subsequent somatic MCR mutation, thereby increasing the likelihood of polyposis. This implies that a truncated MCR APC mutant has an interfering or gain of function property that enhances tumor progression beyond simple loss of APC function. An interfering function would probably not involve interaction with wild-type APC, as recently suggested, because the MCR mutant is still selected for in the absence of a wild-type APC gene copy (Dihlmann et al. 1999). Finally, some of the germline mutations in APC do not disrupt the open reading frame yet correlate with increased risk of colorectal cancer (Frayling et al. 1998; Gryfe et al. 1999; Laken et al. 1997). These mutations have been proposed to increase the occurrence of subsequent truncating mutations by enhancing the mutational susceptibility of the affected nucleotide tract.
Mutations in APC. A compilation of germline and somatic mutations in APC illustrates selection for mutations in the mutation cluster region (MCR). MCR mutations result in truncated proteins retaining β-catenin binding but not regulatory activity. Somatic MCR mutations are more frequently selected for in FAP patients with germline mutations outside of the MCR.
Transcription factors
Prior to discussing the potential role for LEF/TCF transcription factors in cancer, it is important to outline the mechanism by which they have been proposed to operate. Although LEF/TCFs bind directly to DNA through their HMG domains, they are incapable of independently activating gene transcription (Eastman and Grosschedl 1999; Roose and Clevers 1999). This has best been illustrated for LEF, which through its binding to the cofactor ALY, makes indirect contacts with a second transcription factor AML (Bruhn et al. 1997). The TCFs do not contain the ALY binding site, but like LEF they cannot activate test genes comprised of multimerized TCF/LEF binding sites and a minimal promotor sequence. However, these reporter genes are activated on coexpression of TCF with β-catenin, suggesting that β-catenin supplies additional cofactors required for transcriptional activation (Molenaar et al. 1996). This activity was localized to the carboxy-terminal region of the Drosophila β-catenin armadillo, which when fused directly to TCF resulted in β-catenin independent transcriptional activation (van de Wetering et al. 1997).
The simple interpretation is that the TCF/LEF-β-catenin complex comprises a bipartite positive acting transcription factor in the wnt pathway. This interpretation agrees well with developmental studies in which the manipulation of LEF/TCF function results in phenotypes consistent with the genetic manipulation of wnt/β-catenin signaling (Behrens et al. 1996; Brunner et al. 1997; Huber et al. 1996; van de Wetering et al. 1997). For example, a zygotic homozygous null mutation inDrosophila LEF results in a loss of naked cuticle in the larval epidermis, a phenotype typical of loss of function wingless mutations (Brunner et al. 1997). Moreover, the formation of excess naked cuticle by ectopic expression of armadillo in wild-type embryos does not occur in the LEF null mutants. Exactly how β-catenin contributes to transcriptional activation is unclear, but might involve additional proteins that bridge the TCF/β-catenin complex to the basal transcriptional machinery. The bridging function might be fulfilled by Pontin 52, a TATA-binding protein that was reported to bind to β-catenin (Bauer et al. 1998). Also, a mutant form of β-catenin incapable of binding LEF squelched LEF-dependent reporter gene activation, presumably by titration of an essential cofactor (Prieve and Waterman 1999). Finally, the carboxy-terminal region of armadillo binds to the Zinc finger protein teashirt, a homeotic gene essential for a subset of wingless signaling outputs in Drosophila (Gallet et al. 1999).
The simple model of positive transcriptional activation by the TCF-β-catenin complex is not in accord with all experiments. Mutation of the TCF/LEF binding sites in the promotors of the wingless responsive gene ultrabithorax and the Wnt-responsive Xenopus gene Siamois enhanced their activities under conditions where the wingless/β-catenin signal input was weak (Brannon et al. 1999; Riese et al. 1997). The mammalian cyclin D gene was recently identified as a wnt target and, again, mutation of the corresponding TCF binding sites enhanced its basal activity (Tetsu and McCormick 1999). These results suggest TCF represses transcription of its target genes in unstimulated cells and the binding of β-catenin promotes derepression. Derepression cannot fully account for signaling activity, however, as mutations in the TCF binding sites compromise target gene activation under conditions of active wnt signaling (Brannon et al. 1999; Riese et al. 1997). Repression of gene expression by TCF is consistent with its direct physical interaction with at least three different gene products, the Groucho/TLE and CtBP corepressors, and the CREB binding protein CBP (Brannon et al. 1999;Cavallo et al. 1998; Levanon et al. 1998; Roose et al. 1998; Waltzer and Bienz 1998).
The groucho/TLE proteins bind to the central region of TCF/LEF at a site distinct from that of β-catenin binding and inhibit gene activation of TCF target genes (Levanon et al. 1998; Roose et al. 1998). By contrast, CtBP binds to two independent sites in the carboxy-terminal region of Xtcf-3, which when mutated abrogated the repressor function of this region of Xtcf-3 (Brannon et al. 1999). The binding sites for CtBP are not present in LEF, which might explain the ability of LEF, but not Xtcf-3, to induce axis duplication in Xenopus embryos. Finally, the Drosophila CREB binding protein CBP was reported to bind to the HMG domain of dTCF (Waltzer and Bienz 1998). Loss-of-function CBP mutants displayed some features of wingless over expression and also suppressed phenotypes resulting from loss of the β-catenin homolog armadillo. The genetics imply suppression of wingless by CBP, which is somewhat paradoxical when considering the role of CBP acetyltransferase activity in chromatin remodeling and gene activation. However, it was shown that CBP acetylates a lysine proximal to the armadillo binding site in TCF, thereby reducing its affinity for armadillo. Repression of β-catenin/TCF signaling by CBP does not occur in all settings, though, as two recent studies demonstrated activation ofXenopus TCF target genes by CBP (Hecht et al. 2000;Takemaru and Moon 2000). CBP directly associated with carboxy-terminal sequence in β-catenin and overexpression of E1A, which also directly binds CBP, reduced β-catenin dependent transactivation.
Does the activation of TCF/LEF target genes by β-catenin cause cancer? Good evidence to this effect was provided by the expression of a chimeric protein consisting of the LEF DNA binding sequence fused to the transcriptional activation domain of either VP16 or the estrogen receptor (Aoki et al. 1999). Expression of these constructs in chicken embryo fibroblasts resulted in their neoplastic transformation. The proliferative potential of TCF was also apparent from the phenotype resulting from homozygous disruption of TCF-4 in the germline of mice. These animals were incapable of maintaining a proliferative stem cell compartment in the small intestine and died shortly after birth (Korinek et al. 1998). Whether the TCF/LEF genes are directly activated by mutations in cancer is unclear, but mutations in TCF-4 have been detected in a subset of colorectal tumors (Duval et al. 1999). The mutations all occur as single base deletions in an (A)9 nucleotide repeat within the 3′ coding region of the gene. These deletions generate frame shifts predicted to effect the proportion of the long and short forms of TCF that normally result from alternative mRNA splicing. The mutations were also found in cancer cell lines, all of which possessed mutations in either APC or β-catenin. This indicates that the TCF mutations do not substitute for APC/β-catenin mutations but could act in an additive manner.
An additional mechanism by which TCFs could contribute to cancer was gleaned from the phenotype of mice homozygous for mutations in TCF-1 (Roose et al. 1999). Fifteen percent of these animals developed adenomatous intestinal polyps by one year of age, implicating TCF-1 as a tumor suppressor. The major isoforms of TCF-1 do not contain a β-catenin binding site and could therefore act in a dominant negative manner in wnt signaling. Crossing TCF-1 null mice with cancer-prone ApcMin/+ mice resulted in offspring with ten times the number of intestinal polyps relative to ApcMin/+ littermates. This experimental model suggests that the genetic ablation of TCF-1 could modify, or even independently contribute to cancer progression in humans. Additional potential mechanisms for cancer would include the inactivation of corepressors such as CtBP and TLE/groucho.
Cross talk
Defects leading to activation of the wnt pathway could also occur in signaling systems that are seemingly unrelated to wnt signaling. One potential mode of cross talk includes the kinase TAK1, which can substitute for MAPK kinase kinase in the yeast pheromone pathway. TAK1 (TGF-β activatedkinase) is activated by TGF-β in mammalian cells and has also been implicated in interleukin-1 activation of NFκB (Ninomiya-Tsuji et al. 1999; Yamaguchi et al. 1995). In c. elegans, the TAK1 homolog MOM-4 negatively regulates the TCF homolog POP-1 by activating another kinase LIT-1, which then phosphorylates POP-1 (Meneghini et al. 1999;Shin et al. 1999). LIT-1 is thought to gain access to POP-1 through its direct binding to the β-catenin homolog WRM-1 (Shin et al. 1999). Parallel interactions have been demonstrated for the mammalian counterparts of these proteins where the phosphorylation of TCF, by the LIT-1 homolog NLK, reduces its DNA binding affinity (Ishitani et al. 1999). Thus a MAPK-like signaling system might downregulate the wnt-1 pathway. A second opportunity for cross talk with wnt signaling was realized by a physical interaction between the β-catenin-TCF complex and SMAD4, a mediator of TGF-β signaling (Nishita et al. 2000). This interaction was proposed to be synergistic with respect to the activation of theXenopus wnt target gene twin. How this relates to cancer is somewhat puzzling when considering that TGF-β signaling is typically compromised by genetic and epigenetic defects during tumor progression.
An additional mode of cross regulation was recently revealed by the discovery that retinoids inhibit β-catenin dependent gene transcription (Easwaran et al. 1999). β-catenin associated with a retinoic acid receptor (RAR) and cooperated with retinoids to enhance activation of a retinoic acid responsive promotor. Moreover, the identification of RAR-γ as a target of wnt signaling inXenopus also points to an interaction between these signaling systems (McGrew et al. 1999). Signaling by β-catenin was also reported to be repressed by expression of sox3 and sox17 transcription factors, which associated directly with β-catenin (Zorn et al. 1999). Although inhibition of β-catenin signaling was clearly demonstrated, it is also possible that β-catenin drives gene activation independent of LEF/TCF, through its association with the sox proteins. Finally, the activation of the WISP genes by β-catenin is highly dependent upon the presence of a CREB binding site in the WISP promotor (Xu et al. 2000). This implies that cAMP-dependent protein kinase A feeds into wnt signaling and might cooperate with the activation of some wnt target genes. The binding of CBP to β-catenin is particularly relevant with respect to this proposal (Hecht et al. 2000; Takemaru and Moon 2000).
Conclusion
It is apparent that wnt signaling causes cancer and that tumor promotion by this pathway can proceed through a number of different genetic defects. Additional mechanisms by which defects in the regulation of wnt signaling contribute to tumor progression probably remain undiscovered. The manifestation of cancer by aberrant wnt signaling most likely results from inappropriate gene activation mediated by stabilized β-catenin. Target genes need not contain TCF/LEF binding sites in their promotors, though, as new potential modes of gene activation by β-catenin are becoming apparent. Several target genes of β-catenin signaling have now been identified and some of their functions are consistent with control of cellular growth, differentiation, and survival. For an excellent summary of wnt target genes, and a wealth of information on wnt signaling in general, I refer the reader to the Wnt Home Page posted by the Nusse lab (http://www.stanford.edu/rnusse/wntwindow.html).
7.10.2 The Wnt.β-catenin pathway in ovarian cancer: a review.
Objective: Ovarian cancer is the deadliest gynecologic malignancy and the fifth leading cause of death from cancer in women in the U.S. Since overall survival remains poor, there is a need for new therapeutic paradigms. This paper will review the Wnt/β-catenin pathway as it relates to epithelial ovarian cancer, specifically its role in chemoresistance and its potential role as a target for chemosensitization. Methods: A PubMed search was performed for articles published pertaining to Wnt/β-catenin pathway specific to ovarian cancer. Wnt/β-catenin signaling pathways play an active role in cancer stem cells (CSCs) and carcinogenesis of all ovarian cancer subtypes. Studies also have shown that ovarian CSCs are involved in chemoresistance, metastasis, and tumor recurrence. Results: Wnt/β-catenin target genes regulate cell proliferation and apoptosis, thereby mediating cancer initiation and progression. The Wnt/β-catenin pathway is one of the major signaling pathways thought to be involved in epithelial-to-mesenchymal transition (EMT). Alterations affecting Wnt pathway proteins on the cell membrane, in the cytoplasm, and in the nucleus have been shown to play important roles in the tumorigenesis of ovarian cancer. Conclusions: Wnt signaling is activated in epithelial ovarian cancer. Given the role of the Wnt/β-catenin pathway in carcinogenesis, more pre-clinical studies are warranted to further investigate other Wnt inhibitors in ovarian cancer. The Wnt pathway should also be investigated as a potential target in the development of new drugs for ovarian cancer as a single agent and in combination with chemotherapy or other targeted agents.
Introduction
Ovarian cancer is the deadliest gynecologic malignancy and the fifth leading cause of death from cancer in women in the U.S. In 2013, there will be an estimated 22,240 newly diagnosed cases of ovarian cancer and an estimated 14,030 deaths in the United States [1].A major contributor to the high mortality rate is the fact that 70% of women with ovarian cancer initially present with metastases throughout the peritoneal cavity. Over the last two decades, advances in chemotherapy have improved the overall survival in patients with advanced stage ovarian cancer [2]. These advances include the introduction of taxane/platinum-based chemotherapy, intraperitoneal delivery of chemotherapy,dose-dense chemotherapy, and the availability of novel agents such as bevacizumab [3,4].Since overall survival remains poor, there is a need for new therapeutic paradigms. Further research is needed to understand how molecular pathways contribute to the development of metastasis, recurrence, and resistance of ovarian cancer to chemotherapeutic agents. Studies have shown that ovarian cancer stem cells (CSCs) are also involved in chemoresistance, metastasis, and tumor recurrence [5]. CSCs area subpopulation of cancer cells that possess characteristics associated with normal stem cells and are able to generate tumors through the stem cell processes of self-renewal and differentiation.These cells are proposed to persist in tumors as a distinct population that cause recurrence and metastasis by giving rise to new tumors. Recently, chemoresistance has been reported to be associated with acquiring epithelial to mesenchymal transition (EMT) in ovarian cancer cells [6].CancercellsundergoingEMT are unique in that they have stem-like properties that enable cancer cell dissemination and metastasis formation [7]. Major signaling pathways involved in EMT include TGF-β, Wnt/ β-catenin, Notch, Hedgehog, and others [8]. Endometrioid ovarian carcinomas often harbor mutations in the β-catenin gene, but mutations in the Wnt/β-catenin pathway are rare in serous, clear cell, and mucinous ovarian carcinomas [9]. There is emerging data that suggests that despite not having mutations, the Wnt/β-catenin pathway plays a role in carcinogenesis of all ovarian cancer subtypes [10–12]. It has been suggested that the Wnt/β-catenin target genes can be divided into two groups: a “stemness/proliferation group” that is active early in tumor progression and an “EMT/ dissemination group” that is expressed in late stage tumors. The Wnt/ β-catenin pathway has been shown to be a therapeutic molecular target for CSCs[13].Wnt/β-catenin target genes regulate cell proliferation and apoptosis,thereby mediating cancer initiation and progression [14,15]. Given the role of the Wnt/β-catenin pathway in carcinogenesis, we will review the Wnt/β-catenin pathway as it relates to epithelial ovarian cancer, specifically its role in chemoresistance and its potential role as a target for chemosensitization.
Historical perspective of Wnt signaling in the ovary
In the late 1990s, the importance of the Wnt pathways in the embryonic development of the ovary was established. Wnt-4, a Wnt ligand, demonstrated a critical role in embryonic ovarian development [16]. Wnt-7a was shown to affect sex-specific differentiation of the reproductive tract [17]. In 2002, Ricken et al. reported that components of the Wnt signaling pathways are expressed in the immature rat ovary, and that their expression is localized to specific ovarian compartments [18]. This study reported the expression of three different Wnt transcripts (Wnt-2b, Wnt-5a, Wnt-11) that were common to five ovarian cancer cell lines derived from histologically varied human ovarian carcinomas.These results raised the possibility that aberrant Wnt expression may be involved in ovarian tumorigenesis in humans. Prior to this study, alterations in Wnt expression had been described in a variety of female human tumors, including breast and endometrial cancer, but this was the first study to suggest its involvement in ovarian cancer. When β-catenin gene mutations were initially discovered in ovarian cancer, they were thought to be limited to the endometrioid subtype [19]. A study by Wu et al. carried out a comprehensive molecular analysis of 45 tumor specimens of primary ovarian endometrioid adenocarcinomas and two ovarian endometrioid adenocarcinomaderived cell lines. They found Wnt/β-catenin pathway defects in both the cell lines and in nearly half of the primary ovarian endometrioid adenocarcinomas analyzed. β-catenin deregulation was most often attributable to a mutation of the β-catenin gene (CTNNB1) itself, although less frequently β-catenin deregulation may have resulted from inactive mutations in the APC, AXIN1, orAXIN2 genes [20]. Depending on the study, a wide range (3–59%) of serous ovarian cancers have also been reported to stain positive for cytoplasmic and nuclear β-catenin by immunohistochemistry even in the absence of mutations in APC, Axin or β-catenin, which are more common in the endometrioid subtype [21–23]. More recent data have shown that although gene mutations in the Wnt/β-catenin pathway are relatively uncommon in ovarian cancer in general, especially in serous ovarian cancer,components of the pathway are still important in the molecular events that lead to ovarian cancer development [12]. There are three main Wnt signaling pathways: the canonical Wnt/β-catenin pathway, the non-canonical planar cell polarity pathway, and the non-canonical Wnt–Ca2+ pathway. These pathways belong to one of two categories: canonical or non-canonical. The difference between these two categories is the presence or absence of β-catenin. The canonical Wnt/β-catenin pathway involves this protein and the non-canonical pathway operates independently of it.
Components of the Wnt signaling pathway
Non-canonical Wnt signaling pathways
Wnt proteins, which serve as ligands for the Wnt pathway, consist of 19 cysteine-rich glycoproteins. They bind to the Frizzled (Fzd) transmembrane receptor, one of the main receptors of the Wnt pathways [24]. When Wnt binds to Fzd, it can activate one of the three distinct intracellular signaling pathways. While the canonical Wnt/β-catenin signaling pathway leads to the accumulation and stabilization of cytosolic, unphosphorylated (“free”) β-catenin, the non-canonical pathways promote an increase in intracellular calcium or mediate cell polarity. In all three pathways, a Wnt ligand binds to Fzd receptor and promotes recruitment of Dishevelled (Dsh) protein (Figs. 1 and 2). In the planar cell polarity non-canonical pathway, this complex binds to the Dsh-associated activator of morphogenesis (Daam1). This cascade of events leads to the activation of Rac and RhoA GTPases which mediate cell polarity (Fig. 1). In the Wnt-Ca2+ noncanonical pathway, the Wnt/Fzd/Dsh complex binds with a G protein (Ror 1/2) as shown in Fig. 1, which leads to activation of calmodulindependent kinase II, protein kinase C and the phosphatase calcineurin. This binding promotes the increase in intracellular calcium levels which then mediates other signaling pathways. The Wnt pathways are critical to embryonic development of a variety of organs including the ovaries. Activation of Wnt signaling occurs via the canonical Wnt/β-catenin pathway and the non-canonical cell polarity pathway and the Wnt/ Ca2+ pathway; however, as it relates to oncology research the Wnt/β-catenin canonical pathwayis the mostrelevant [25].
Canonical (Wnt/β-catenin) signaling pathway
In the canonicalWnt/β-catenin pathway, the pathway is “off” when either there is no Wnt ligand, no receptor, or the receptor is being blocked (Fig. 2A). Dikkopf family (DKK1–4) binds directly to one of the transmembrane receptors (Fzd, LRP5/6) and blocks Wnt from binding. Wnt-inhibitory factor (WIF-1) and the family of secreted Fzd receptor proteins (SFRP1-5) bind to Wnt itself and prevent them from binding to the receptors. If the Wnt ligand does not bind to the receptors, β-catenin is degraded by a destruction complex that is comprised of Axin, adenomatous polyposis coli (APC), and glycogen synthase kinase 3β (GSK3β). β-Catenin is phosphorylated by the kinases casein kinase 1 (CK1) and GSK3β, followed by ubiquitination and proteasomal degradation by the 26S proteasome. Low cytoplasmic levels of β-catenin allow for the recruitment of the corepressor Groucho to lymphoid enhanced factor–T-cell factor (TCF/LEF) transcription factors,which blocks the target genes from being activated and ensures transcriptional repression (Fig. 2A). Activation of the canonical Wnt pathway involves the stabilization of β-catenin through the binding of Wn tligands to cell surface receptors including Fzd family receptors and low-density lipoprotein receptor (LDLR)-related proteins: LRP5 and LRP6. When the Wnt pathway is “on”, cytosolic β-catenin is stabilized (Fig. 2B). LRP6/LRP5 is phosphorylated by the kinases CK1 and GSK3β. Dsh molecules are recruited to the plasma membrane to interact with Fzd. Interaction of Axin with phosphorylated LRP6/LRP5 and Dsh leads to inactivation of the destruction complex and degradation of β-catenin is inhibited. βCatenin accumulates in the cytoplasm and enters the nucleus and activates Wnt target genes by binding to the transcriptional factors of the TCF/LEF family by displacing Groucho and interacting with coactivators such as B-cell lymphoma 9/Legles (BCL9/LGS) and Pygopus (Pygo) to promote transcription of target genes [26]. TCF/LEF, BCL9/ LGS, and Pygo all bind with β-catenin in the nucleus to form a transcriptional activation complex (Fig. 2B). β-Catenin promotes transcription of genes related to proliferation and survival. Some of the key downstream proteins and genes that are activated with the binding of β-catenin to the transcriptional factors of the canonical pathway include c-MYC (MYC), Cyclin D1 (CCND1), Survivin (BIRC5), Axin2 (AXIN2), and matrix metalloproteinases (MMPs). There have been over 100 target genes identified as regulated by the Wnt pathway and 23 of them have been shown to be overexpressed in ovarian cancer [27].
Regulation of the Wnt pathway
The remainder of the review will focus on the canonical Wnt/ β-catenin pathway, because the Wnt/β-catenin pathway has been the most well described in the literature as it relates to cancer research and specifically ovarian cancer. It is regulated at multiple levels: gene mutations, extracellular inhibitors, and intranuclear transcription cofactors. These all contribute to the diverse mechanisms that are involved in the Wnt pathway.When there is no Wnt ligand, a destruction complex regulates β-catenin levels. Specifically, CK1 and unphosphorylated GSK3β phosphorylate β-catenin and target the protein for ubiquitination and proteasomal degradation. Phosphorylation of GSK3β by protein kinases (A, B, and C), Akt/PI3K, and MAPK inhibits its ability to phosphorylate and target β-catenin for degradation. The majority of ovarian cancers have an activation of PI3K (phosphoinositide 3-kinase) by gene amplification, which can potentially phosphorylate GSK3β, impeding the phosphorylation of β-catenin and resulting in cellular differentiation, division, and survival [28,29].
Alterations of the Wnt pathway in ovarian cancer
Membranous factors
The first event in the activation of the Wnt pathway is the binding of a Wnt ligand to Fzd and LRP6/LRP5. Two subtypes of the Fzd receptor are increased in epithelial ovarian cancer, Fzd1 and Fzd5. A higher number of malignant ovarian specimens stained positive for both receptors than normal ovary and the Fzd5-positive tumors had a worse 6-year survival than those that were Fzd5-negative [30]. During metastatic spread of epithelial ovarian cancer, there is adhesion of cancer cells to submesothelial interstitial collagens. When β1 integrin mediated anchoring to the mesothelium and submesothelial matrix occurs, it facilitates the formation of metastatic tumor sites on other peritoneal organs. The engagement of collagen-binding β1 integrins have been shown to upregulate LRP6, WNT5A, MMP9, PTGS2 (COX2), PLAUR (uPAR), VIM (vimentin), SNAII (Snail) at the mRNA level [31]. This suggests tha tmetastatic spread of ovarian cancer is likely facilitated by the upregulation of LRP6 and targeting LRP6 may be an effective strategy for treating ovarian cancer.
There are several proteins that act as antagonists to the Wnt pathway. These proteins include: the Dikkopf family (DKK1–4), Wnt inhibitory factor (WIF-1) and the family of secreted Fzd receptor proteins (SFRP1-5)(Fig.2A). SFRPs bind directly to the Wnt ligand or Fzd receptor and inhibit Wnt from binding to Fzd and activating the pathway. Loss of SFRP4 expression correlates with a more aggressive ovarian cancer phenotype and the level of SFRP4 is directly related to prognosis [32]. Investigators have studied the re-expression of SFRP4 in epithelial ovarian cancer cell lines, and found that re-expression inhibited the Wnt/ β-catenin signaling pathway, thereby inhibiting cell migration and EMT. These proteins provide important potential therapeutic targets by either re-expression, if their expression is lost,or potentially upregulated.
Cytoplasmic and nuclear factors
Endometrioid ovarian carcinomas often have mutations in the βcatenin gene. Table 1 summarizes the studies that show β-catenin mutations in human ovarian cancer, from 16% to 54% in endometrioid cancers and 14% in mucinous cancers. Despite no reported mutations in the CTNNB1 gene in serous and clear cell cancers, nuclear β-catenin has been observed in serous and clear cell ovarian cancer [21]. Lee et al. showed a statistically significant correlation between nuclear β-catenin expression and high-grade serous ovarian cancer [23]. The protooncogene, frequently rearranged in advanced T-cell lymphomas-1 (FRAT1), which inhibits phosphorylation of β-catenin, was found to be overexpressed in serous ovarian cancer and was strongly correlated with the accumulation of cytoplasmic β-catenin, leading to an increase in nuclear β-catenin [21]. Pygo, oneof the co-activators that binds to β-catenin is a necessary component of tumor cell growth and is widely expressed in ovarian cancer, both in cell lines and in primary tumor tissue [33]. RNA expression of BCL9/LGS, also a co-activator,is common in both epithelial ovarian cancer and normal ovaries. Upregulation of these co-activators is further evidence that the Wnt pathway plays a pivotal role in the tumorigenesis of ovarian cancer.
Intercellular interactions
Cells undergoing EMT are known to lose E-cadherin and gain vimentin expression, resulting in tumor cell invasion and metastasis [34]. Epithelial ovarian cancer cells also undergo a mesenchymal to epithelial transition (MET) because the normal ovarian surface epithelium is mesenchymally derived. This dynamic process has been termed EMP (epithelial to mesenchymal plasticity). It is thought that both transitions are equally important for metastasis formation and that the “metastable” state is actually when the cells transition between the two states [34]. Metastatic epithelial ovarian cancer cells adhere to the interstitial collagen of the peritoneal cavity via integrins. Cell–matrix and cell– cell adhesions are paramount to this process and are mediated by integrins and E-cadherins. Integrin engagement has been linked to increased internalization of E-cadherin [31]. In epithelial cancer, the MET component dominates, unlike other epithelial cell-derived cancers where the EMT component dominates; therefore, E-cadherin expression is increased with malignant transformation in ovarian cancer [31]. E-cadherin-based adherens junctions are stabilized by β-catenin, and the loss of stability in the junctions may cause an increase in cytoplasmic and/or nuclear β-catenin. Integrins have also been suggested to inhibit GSK3β, elevate levels of nuclear β-catenin, and increase β-catenin-regulated promoter activation. Burkhalter et al.
showed that an inhibitor of β-catenin and TCF-4, a member of the TCF/LEF transcription factor family, reduced cellular invasion [31]. Most of the regulation of the Wnt pathway ultimately leads to an accumulation or depletion of β-catenin in the nucleus, or affects the binding of nuclear β-catenin to TCF/LEF, which determines whether apoptosis can occur. It is important to note that the transcriptional regulatory activity of β-catenin is also controlled by factors other than Wnt signaling. One example of Wnt-independent regulation of β-catenin is through E-cadherin expression, which selectively depletes the transcriptionally active pool of β-catenin [35]. This is especially significant as epithelial ovarian cancer cells are known to undergo MET which causes an increase in E-cadherin.
Extracellular factors
Not only have membranous and intercellular components of theWnt pathway been found to be upregulated in epithelial ovarian cancer, but extracellular activators also are upregulated. These factors specifically include Wnt-1,Wnt-2b,Wnt-5a, and Wnt-11 [30]. Ricken et al. reported the possibility that Wnt-5a could be involved in ovarian carcinogenesis [18]. This study used RT-PCR on RNA from five ovarian cancer cell lines and confirmed the expression of transcripts for Wnt-2b, Wnt-5a and Wnt-11. Filho et al. showed that upregulation of Wnt-1 and Wnt-5a, detected by immunohistochemistry in patient samples, portended a significantly lower survival than ovarian cancer patient samples that did not have an upregulation of Wnt-1 and Wnt-5a [30].
Gene expression
Kumar et al. analyzed 1500 miRNAs to identify which ones were potentially different between A2780 (parental ovarian cancer cell line) and A2780.cp70 (cisplatin resistant cell line) and found changes in 11 miRNAs [36]. The microRNA data was validated by quantitative realtime PCR for these 11 miRNAs. Ingenuity Pathway Analysis (IPA) and Kyoto Encyclopedia of Genes and Genomes (KEGG) analysis were performed for the 11 miRNAs and their targets to identify the pathways involved in cisplatin resistance. Not only was Wnt signaling one of the pathways identified, but so were MAPK and mTOR signaling pathways which both cross-talk with the Wnt pathway by causing the phosphorylation of GSK3β, blocking its ability to phosphorylate βcatenin to allow it to be ubiquitinated. Four gene expression datasets: Moffitt Cancer Center (MCC), Total Cancer Care (TCC), the Cancer Genome Atlas(TCGA),andMDAnderson (MDA) were analyzed, and only four pathways were noted to be differentially expressed between normal ovarian surface epithelium and ovarian cancer. One of these pathways is the “Cytoskeleton remodeling/TGF–Wnt pathway” [37]. The“Cytoskeleton remodeling/TGF/WNT” pathway was previously described as a common pathway created by the crosstalk between the TGF-β pathway and the Wnt pathway that is involved in cytoskeleton remodeling: cell–cell adhesion and cell–matrix adhesion [38]. This pathway has been associated with metastasis in various cancer types and is critical for cancer cell migration and invasion. The same group at H. Lee Mof fitt Cancer Center found that six common molecular signaling pathways were associated with chemoresistance and survival in ovarian cancer that included the TGF– Wnt pathway and specifically Wnt pathway activated by Wnt-2, one of the 19 Wnt ligands [39]. In addition, this group also used the same novel computer analysis technique to identify genes and molecular signaling pathways associated with cancer cell proliferation. Genes and pathways associated with cancer cell proliferation and survival were analyzed against the NCI 60 cell line-drug screening database to identify agents predicted to have pathway- and gene-specific activity. They identified 81 existing agents that could potentially be repurposed to target the TGF-Wnt pathway that are currently the focus of in vitro functional analyses [40].
Non-canonical pathways
Fig. 1. Non-canonical Wnt signaling pathways. In the planar cell polarity pathway Wnt–Frizzled complex binds to the Dsh-associated activator of morphogenesis (Daam1). This cascade of events leads to the activation of Rac and RhoA GTPases which mediate cell polarity. In the Wnt–Ca2+ pathway, the Wnt/Fzd/Dsh complex binds with a G protein, which leads to activation of calmodulin-dependent kinase II (CaMKII), protein kinase C (PKC), and the phosphatase calcineurin. This binding promotes the increase in intracellular calcium levels which stimulates other signaling pathways.
Fig.2.The canonical Wnt signaling pathway. (A)In the absence of Wnt ligand, β-catenin is degraded through interactions with Axin, APC and GSK3β “destruction complex”. β-Cateninis phosphorylated by the kinases CK1 (casein kinase 1) and GSK3β (glycogen synthase kinase 3β), followed by ubiquitylation and proteasomal degradation. Low cytoplasmic levels of βcatenin allow for the recruitment of the corepressor Groucho to LEF (lymphoid enhanced factor)–TCF (T-cell factor) transcription factors which ensures transcriptional repression. Dikkopf (DKK) family proteins, the Wnt-inhibitory factor (WIF), and the family of secreted Frizzled receptor proteins (SFRPs) all act as antagonists to the Wnt pathway. SFRP binds directly to the Wnt ligand or th eFrizzled receptor to inhibit Wnt binding to Frizzled. (B) In the presence of Wnt ligands, Wnt proteins bind to Frizzled/LRP6/LRP5 receptor complex at the cell surface. LRP6/LRP5 is phosphorylated by the kinases casein kinase 1 (CK1) and glycogen synthase kinase 3β (GSK3β). Dishevelled (Dsh) molecules are recruited to the plasma membrane to interact with Frizzled. Interaction of Axin with phosphorylated LRP6/LRP5 and Dsh leads to inactivation of the destruction complex. Degradation of β-catenin is inhibited. β-Catenin accumulates inthe cytoplasm and nucleus. β-Catenin forms a transcriptionally active complex with TCF/LEF by displacing Groucho and interacting with co-activators suchasBCL9/LGS (B-cell lymphoma 9/Legless) and Pygo (Pygopus) to promote transcription of target genes (Axin, CyclinD1, Survivin). β-Catenin is also a coactivator of CREB binding protein (CBP) which is the binding protein of the cAMP response element-binding protein (CREB). β-Catenin/CBP binds to Wnt-responsive element (WRE) and activates transcription. This leads to cell proliferation, survival, and self-renewal.
Potential therapeutic targets of the Wnt pathway in ovarian cancer
Identification of the specific membranous, intracellular, and extracellular components of the Wnt pathway gives insight to potential targets for therapy. There currently are several small molecules that have recently entered into phase I clinical trials that target the Wnt pathway (Table 2). In order for the Wnt protein to be secreted by the cell to act as a ligand it must first undergo fatty acyl modification. Once it undergoes palmiteolyation it is shepherded through the secretory pathway by Wntless chaperone protein. PORCN is the founding member of a 16-gene family with acyltransferase activity and Porcupine (Porcn) is the acyltransferase enzyme that adds the fatty acid to Wnt which is a crucial step in the secretion of the Wnt ligand. Without Porcn to catalyze this modification, the Wnt protein remains trapped inside the cell. Currently being studied in a phase 1 trial is the small molecule, LGK974 (Novartis Pharmaceuticals) that inhibits Porcn(NCT01352203) [41]. Drugs that specifically target the Wnt signaling pathway in the nucleus include the small molecule inhibitor, PRI-724, which specifically blocks the recruitment of β-catenin with its coactivator CBP which is the binding protein of the cAMP response element-binding protein CREB. βCatenin/CBP binds to Wnt-responsive element (WRE) and activates transcription; therefore, PRI-724 prevents activated transcription by aberrant Wnt signaling. This drug is being studied in solid tumors and myeloid malignancies (NCT01606579) [41]. Other pathways may cross-talk with the Wnt pathway. In Wnt signaling, Axin is a key scaffolding protein of the destruction complex of β-catenin, and Poly (ADP ribose) polymerases (PARPs) promote the ribosylation of Axin, thereby causing it to become degraded and no longer facilitate β-catenin destruction. If PARP is inhibited, Axin is stabilized, which allows it to degrade β-catenin [42]. There are several PARP inhibitors that are currently being used in clinical trials for ovarian cancer. In addition, preclinical studies have been carried out with XAV939, which is a small-molecule PARP inhibitor that targets tankyrases, a specific type of PARP. Huang et al. used a chemical genetic screen to identify the small molecule, XAV939, which selectively inhibits β-catenin mediated transcription. XAV939 was shown to stimulate β-catenin degradation by stabilizing Axin. They used a quantitative chemical proteomic approach to show that XAV939 stabilizes Axin by inhibiting tankyrase1 and tankyrase2.They showed that both tankyrase isoforms 1 and 2 stimulate Axin degradation through the ubiquitin–proteasome pathway [43]. JW55 (Tocris Bioscience) is a selective tankyrase 1 and 2 inhibitor which has been shown to inhibit the growth of cancer. JW55 inhibits the canonical Wnt signaling pathway in colon carcinoma cells that contained mutations either in the APC locus or in anallele of β-catenin [44]. Frizzled, oneof themembrane receptors that activates thepathway upon Wnt ligand binding, has been reported to be overexpressed in ovarian cancer. There are two drugs that specifically target the Fzd receptorthatarebeingevaluatedinclinicaltrials.OMP-18R5(OncoMed Pharmaceuticals/Bayer) is one of the Wnt-targeted compounds that is in clinical development (NCT01345201) [41]. It is a monoclonal antibody that targets Fzd receptors and blocks their association with Wnt ligands. This drug is being used in combination with the standard chemotherapy for breast, lung, pancreas, and colon cancer. Another drug, OMP-54F28, binds to and sequesters the Wnt ligand and is a fusion protein of the Fzd8 ligand-binding domain with the Fc region of a human immunoglobulin (OncoMed Pharmaceuticals/Bayer) (NCT01352203) [41]. There has been a growing trend in oncology to evaluate“repurposed” drugs which are drugs that have been used in the past for other purposes and are now being screened for their function as anticancer drugs. Several drugs have been shown to work through the Wnt pathway including the FDA-approved anti-helminth compound, niclosamide, non-steroidalanti-inflammatory drugs(NSAIDS), and two antipsychotic drugs: lithium and valproic acid. NSAIDS have been shown to cause degradation of TCF and inhibit Wnt target genes such as COX2. Although they do not target the Wnt pathway directly, they could be a potential anti-Wnt agent. Niclosamide inhibitsWnt/β-catenin pathway activation. In colorectal cancer, it was shown to downregulate Dvl2, a member of the Dsh protein family, which in turn decreased downstreamβ-catenin signaling [45]. Recently, niclosamide has been reported to target not only Wnt/β-catenin but also other signaling pathways involved in CSC maintenance such as NF-κB, Notch, ROS, mTORC1, and Stat3 [46,47]. Niclosamide has also been reported to inhibit Wnt/β-catenin signaling by inducing degradation of the Wnt surface receptor, LRP6 [48]. Our laboratory has seen an increase expression of LRP6 in ovarian cancer patients. Yo et al. identified a subset of chemoresistant ovarian tumor cells that fulfilled the definition of CSCs and subjected these cells to high-throughput drug screening using more than 1200 clinically approved drugs. Sixty-one potential compounds were identified on preliminary screening and after more stringent screening, niclosamide was found to be the best drug to selectively target ovarian CSCs both in vitro and in vivo [49].
Wnt/β-catenin pathway and CSC
TheWnt/β-catenin pathway is an important pathway in cell survival and has been implicated in the mechanism of chemoresistance of ovarian CSCs. CSCs are a subpopulation of tumor cells that possess characteristics associated with normal stem cells and have the ability to self-renew and differentiate. Wnt/β-catenin signaling plays an important role in the transcription of multidrug resistance genes such as ABCB1/MDR-1 [50]. Chemoresistance, which can be a result of the inhibition of apoptosis, has been reported to be associated with acquiring EMT in ovarian cancer cells [51,52]. Ovarian cancer cells undergoing EMT have stem-like properties that enable cancer cell dissemination and metastasis formation. A recent study done at Georgia Institute of Technology confirmed that metastasizing ovarian cancer cells taken from patients have a different molecular structure from primary tumor cells and display genetic signatures consistent with EMT [53]. The Wnt/ β-catenin pathway is one of the major signaling pathways thought to be involved in EMT and thus has been shown to play an integral role in metastasis.
Conclusions
Alterations affectingWnt pathway proteins on the cel lmembrane, in the cytoplasm, and in the nucleus have been shown to play important roles in the tumorigenesis of ovarian cancer. Pre-clinical studies have shown an upregulation of 5 of the 19 known Wnt ligands in ovarian cancer, which leads to increased activity of the Wnt pathway. Fzd is one of the membrane receptors that activates the pathway upon Wnt ligand binding. It has been reported to be overexpressed in ovarian cancer. Our laboratory has also seen an upregulation of LRP6 detected by immunohistochemistry (unpublished data). In ovarian cancer, an increase in nuclear β-catenin has been shown to be the result of an upregulation in the β-catenin gene itself and also mutations in the proteins necessary to degrade cytoplasmic β-catenin such as Axin2 and APC. The β-catenin destruction complex consists of Axin2, APC, and GSK3β, which must not be phosphorylated in order to cause βcatenin degradation. GSK3β is frequently phosphorylated in ovarian cancer through other pathways, such as PI3K, inhibiting its ability to degrade β-catenin. Upregulation of co-activators of β-catenin also contributes to the increase in transcription of the target genes. As many as 23 different target genes that lead to cell proliferation and survival, which is a result of nuclear β-catenin build-up, have been shown to be overexpressed in ovariancancer. Wntsignalingis activated in epithelial ovarian cancer, both directly through ligand activated upregulation of the pathway and through a ligand independent increase in nuclear β-catenin through cross-talk with other pathways. Recently, Yo et al. reported that niclosamide, which has been shown to have anti-Wnt activity inhibits growth in ovariantumor-initiatingcells[49].Morepre-clinicalstudies,specifically animal studies and mechanistic studies, are warranted to further investigate other Wnt inhibitors in ovarian cancer. The Wnt pathway is very complex, and further studies with targeted agents need to be done to see if inhibition of a single component of the pathway will be clinically useful. This paper supports the fact that the Wnt pathway shows promise as an effective target for anti-cancer therapy in ovarian cancer. As more efficacy data is collected from the phase 1 studies with Wnt inhibitors LGK974, OMP-54F28, OMP-18R5, and PRI724: NCT01352203, NCT01608867, NCT01345201, and NCT01606579 (www.clinicaltrials.gov), they should be considered as potential agents in the treatment of ovarian cancer. Given the fact that the Wnt pathway is involved in so many biological pathways, results from these studies will be important to determine if effective Wnt pathway inhibition will be excessively toxic to patients. Future directions for investigating the Wnt pathway in ovarian cancer should include genetic sequencing of ovarian cancer patients with the aim of targeting those patients who specifically have upregulation of Wnt pathway target genes. More quantitative data is needed to specifically look at the mechanisms of these drugs in patients by performing qPCR on tissue obtained before and after treatment. The Wnt pathway should be investigated as a potential target in the development of new drugs for ovarian cancer as a single agent and in combination with chemotherapy or other targeted agents.
References
[1] Morton BA, Beatty BG, Mison AP, Wanek PM, Beatty JD. Role of bone marrow transplantation in 90Y antibody therapy of colon cancer xenografts in nude mice. Cancer Res Suppl 1990;50:1008s–10s. [2] OzolsRF,BundyBN,GreerBE,FowlerJM,Clarke-PearsonD,BurgerRA,etal.PhaseIII trialofcarboplatinand paclitaxelcompared with cisplatinand paclitaxel inpatients with optimally resected stage III ovarian cancer: a Gynecologic Oncology Group study. J Clin Oncol 2003;21:3194–200. [3] Garcia AA, Hirte H,Fleming G,YangD, Tsao-WeiDD, RomanL,etal. Phase IIclinical trialofbevacizumabandlow-dosemetronomicoralcyclophosphamideinrecurrent ovarian cancer: a trial of the California, Chicago, and Princess Margaret Hospital phase II consortia. J Clin Oncol 2008;26:76–82. [4] Penson RT, Dizon DS, Cannistra SA, Roche MR, Krasner CN, Berlin ST, et al. Phase II study of carboplatin,paclitaxel, andbevacizumab with maintenance bevacizumab as first-line chemotherapy for advanced mulleriantumors.J ClinOncol2010;28:154–9. [5] Steg AD, Bevis KS, Katre AA, Ziebarth A, Dobbin ZC, Alvarez RD, et al. Stem cell pathways contribute to clinical chemoresistance in ovarian cancer. Clin Cancer Res 2012;18:869–81. [6] RicciF,BernasconiS,PeregoP,GanzinelliM,RussoG,BonoF,etal.Ovariancarcinoma tumor-initiatingcellshaveamesenchymalphenotype.CellCycle2012;11:1966–76. [7] Raimondi C, Gianni W, Cortesi E, Gazzaniga P. Cancer stem cells and epithelial– mesenchymal transition: revisiting minimal residual disease. Curr Cancer Drug Targets 2010;10:496–508.
[8] Talbot LJ, Bhattacharya SD, Kuo PC. Epithelial–mesenchymal transition, the tumor microenvironment,andmetastaticbehaviorofepithelialmalignancies.IntJBiochem Mol Biol 2012;3:117–36. [9] DubeauL.Thecelloforiginofovarianepithelialtumours.LancetOncol2008;9:1191–7. [10] Boyer A, Goff AK, Boerboom D. WNT signaling in ovarian follicle biology and tumorigenesis. Trends Endocrinol Metab 2010;21:25–32. [11] Rask K, Nilsson A, Brannstrom M, Carlsson P, Hellberg P, Janson PO, et al. Wntsignalling pathway in ovarian epithelial tumours: increased expression of betacatenin and GSK3beta. Br J Cancer 2003;89:1298–304. [12] Gatcliffe TA, Monk BJ, Planutis K, Holcombe RF. Wnt signaling in ovarian tumorigenesis. Int J Gynecol Cancer 2008;18:954–62. [13] Takebe N, Harris PJ, Warren RQ, Ivy SP. Targeting cancer stem cells by inhibiting Wnt, Notch, and Hedgehog pathways. Nat Rev Clin Oncol 2011;8:97–106. [14] Verras M, Sun Z. Roles and regulation ofWnt signaling and beta-catenin inprostate cancer. Cancer Lett 2006;237:22–32. [15] Wend P, Holland JD, Ziebold U, Birchmeier W. Wnt signaling in stem and cancer stem cells. Semin Cell Dev Biol 2010;21:855–63. [16] Vainio S, Heikkila M, Kispert A, Chin N, McMahon AP. Female development in mammals is regulated by Wnt-4 signalling. Nature 1999;397:405–9. [17] Parr BA, McMahon AP. Sexually dimorphic development of the mammalian reproductive tract requires Wnt-7a. Nature 1998;395:707–10. [18] Ricken A, Lochhead P, Kontogiannea M, Farookhi R. Wnt signaling in the ovary: identification and compartmentalized expression of wnt-2, wnt-2b, and frizzled-4 mRNAs. Endocrinology 2002;143:2741–9. [19] Saegusa M, Okayasu I. Frequent nuclear beta-catenin accumulation and associated mutations in endometrioid-type endometrial and ovarian carcinomas with squamous differentiation. J Pathol 2001;194:59–67. [20] Wu R, Zhai Y, Fearon ER, Cho KR. Diverse mechanisms of beta-catenin deregulation in ovarian endometrioid adenocarcinomas. Cancer Res 2001;61:8247–55. [21] WangY,HewittSM,LiuS,ZhouX,ZhuH,ZhouC,etal.Tissuemicroarrayanalysisof human FRAT1 expression and its correlation with the subcellular localisation of beta-catenin in ovarian tumours. Br J Cancer 2006;94:686–91. [22] Karbova E, Davidson B, Metodiev K, Trope CG, Nesland JM. Adenomatous polyposis coli (APC) protein expression in primary and metastatic serous ovarian carcinoma. Int J Surg Pathol 2002;10:175–80. [23] LeeCM,ShvartsmanH,DeaversMT,WangSC,XiaW,SchmandtR,etal.Beta-catenin nuclear localization is associated with grade in ovarian serous carcinoma. Gynecol Oncol 2003;88:363–8. [24] Gonzalez-SanchoJM,BrennanKR,Castelo-SoccioLA,BrownAM.Wntproteinsinduce dishevelled phosphorylation via an LRP5/6-independent mechanism, irrespective of theirabilityto stabilize beta-catenin. MolCellBiol 2004;24:4757–68. [25] HarrisPJ,SperanzaG,DanskyUllmannC.Targetingembryonicsignalingpathwaysin cancer therapy. Expert Opin Ther Targets 2012;16:131–45. [26] Kikuchi A, Yamamoto H, Kishida S. Multiplicity of the interactions of Wnt proteins and their receptors. Cell Signal 2007;19:659–71. [27] Barbolina MV, Burkhalter RJ, Stack MS. Diverse mechanisms for activation of Wnt signalling in the ovarian tumour microenvironment. Biochem J 2011;437:1–12. [28] Bader AG, Kang S, Zhao L, Vogt PK. Oncogenic PI3K deregulates transcription and translation. Nat Rev Cancer 2005;5:921–9. [29] Bast Jr RC, Hennessy B, Mills GB. The biology of ovarian cancer: new opportunities for translation. Nat Rev Cancer 2009;9:415–28. [30] Badiglian Filho L, Oshima CT, De Oliveira Lima F, De Oliveira Costa H, De Sousa DamiaoR,GomesTS,etal.CanonicalandnoncanonicalWntpathway:acomparison among normal ovary, benign ovarian tumor and ovarian cancer. Oncol Rep 2009;21:313–20. [31] Burkhalter RJ, Symowicz J, Hudson LG, Gottardi CJ, Stack MS. Integrin regulation of beta-catenin signaling in ovarian carcinoma. J Biol Chem 2011;286:23467–75. [32] Jacob F, Ukegjini K, Nixdorf S, Ford CE, Olivier J, Caduff R, et al. Loss of secreted frizzled-related protein 4 correlates with an aggressive phenotype and predicts poor outcome in ovarian cancer patients. PLoS One 2012;7:e31885. [33] Popadiuk CM, Xiong J, Wells MG, Andrews PG, Dankwa K, Hirasawa K, et al. Antisense suppression of pygopus2 results in growth arrest of epithelial ovarian cancer. Clin Cancer Res 2006;12:2216–23. [34] Christiansen JJ, Rajasekaran AK. Reassessing epithelial to mesenchymal transition as a prerequisite for carcinoma invasion and metastasis. Cancer Res 2006;66:8319–26. [35] Gottardi CJ, Wong E, Gumbiner BM. E-cadherin suppresses cellular transformation by inhibiting beta-catenin signaling in an adhesion-independent manner. J Cell Biol 2001;153:1049–60. [36] Kumar S, Kumar A, Shah PP, Rai SN, Panguluri SK, Kakar SS. MicroRNA signature of cis-platin resistant vs. cis-platin sensitive ovarian cancer cell lines. J Ovarian Res 2011;4:17. [37] Al Sawah DM E, Xiong Y, Ramirez-Diaz I, Abbasi F, Bou Zgheib N, Stickles X, et al. A novel strategy to identify ovarian cancer molecular signaling pathways and drug repurposing candidates. Abstract supplement from the Society of Gynecologic Oncology 44th Annual Meeting on Women’s Cancer; 2013. p. S7. [38] FloresRJ,LiY,YuA,ShenJ,RaoPH,LauSS,etal.Asystemsbiologyapproachreveals common metastatic pathways in osteosarcoma. BMC Syst Biol 2012;6:50. [39] Marchion IR-D D, Xiong Y, Al Sawah E, Abbasi F, Bou Zgheib N, Stickles X, et al. An innovativeinsilicomethodtoidentifyagentsthattargetpathwaysofhumancancer chemoresistance. Abstract supplement from the Society of Gynecologic Oncology 44th Annual Meeting on Women’s Cancer; 2013. p. S151. [40] RamirezDMI,XiongY,AlSawahE,BouZgheibN,BoacB,SticklesX,etal.Identifying drug repurposing opportunities to target genes and molecular pathways associated with cancer cell proliferation. Abstract supplement from the Society of Gynecologic Oncology 44th Annual Meeting on Women’s Cancer; 2013. p. S152.
7.10.3 Wnt Signaling in the Niche Enforces Hematopoietic Stem Cell Quiescence and Is Necessary to Preserve Self-Renewal In Vivo
Wingless (Wnt) is a potent morphogen demonstrated in multiple cell lineages to promote the expansion and maintenance of stem and progenitor cell populations. Pharmacologic modification of Wnt signaling has been shown to increase hematopoietic stem cells (HSC). We explored the impact of Wnt signaling in vivo, specifically within the context of the HSC niche. Using an osteoblast-specific promoter to drive the expression of a pan-inhibitor of canonical Wnt signaling, Dickkopf1 (Dkk1), we noted changes in trabecular bone and in HSC. Wnt signaling was inhibited in HSC and the cells exhibited reduced p21Cip1 expression, increased cell cycling and a progressive decline in regenerative function after transplantation. This effect was microenvironment-determined, but irreversible if the cells were transferred to a normal host. Wnt pathway activation in the niche is required to preserve the reconstituting function of endogenous hematopoietic stem cells.
The regulation of hematopoietic stem cell function is a complex and balanced process that requires coordinated input from inherent HSC programs and moderating signals provided by the surrounding microenvironment. Together, these signals permit the maintenance of the stem cell pool for the life of the organism, while also allowing for sufficient steady-state and injury-responsive blood cell production. These somewhat dichotomous aspects of HSC function require mechanisms that both preserve a quiescent population of stem cells and also promote their activation, expansion, differentiation and circulation under appropriate conditions (Akala and Clarke, 2006; Scadden, 2006). The morphogen family of signaling molecules has been identified as a prominent player in the function of numerous stem cell types, including the hematopoietic lineage. The wingless (Wnt) pathway has been studied extensively in the context of hematopoiesis, and the combined impact of multiple family members binding to a range of receptors leads to activation of canonical and non-canonical signaling pathways (Nemeth and Bodine, 2007). Canonical signals are mediated by TCF/LEF transcription factor activity (Daniels and Weis, 2005), and are considered to be largely dependent on the accumulation of nuclear β- (and/or γ-) catenin (Nemeth and Bodine, 2007).Wnt signals have been implicated in mammalian hematopoiesis by studies not intended to assess normal physiology in which Wnt activation had a strong expansive effect on reconstituting HSCs and multipotent progenitors (Baba et al, 2006; Murdoch et al, 2003; Reya et al, 2003; Trowbridge et al, 2006). With enforced, persistent Wnt activation, however, engineered mice developed hematopoietic failure with impaired differentiation of HSC (Kirstetter et al, 2006; Scheller et al, 2006). In contrast, deletion of members of the Wnt / β-catenin cascade under homeostatic conditions had little to no effect on blood cell production by HSCs (Cobas et al, 2004; Jeannet et al, 2007; Koch et al, 2007), raising the question of what physiological role, if any, Wnt signaling has on this cell type. Some of the variation observed may reflect differing influences exerted by canonical versus non-canonical Wnt signals, particularly given a recent report indicating that Wnt5a can modulate canonical signals mediated by Wnt3a (Nemeth et al, 2007). Wnt signals are also regulated by a host of soluble inhibitors that may interact directly with Wnt ligands, such as the frizzled-related proteins (sFRP) or by preventing Wnt binding to its receptors (Kawano and Kypta, 2003). The Dickkopf (Dkk) family of Wnt inhibitors falls into this latter category, by binding the Wnt coreceptor LRP5/6 in combination with a Kremen receptor, and leading to internalization of the complex (Mao et al, 2001; Mao et al, 2002). In order to specifically examine the impact of Wnt activation in an in vivomicroenvironment that has been shown to regulate HSC number and function, we utilized mice engineered to overexpress the Wnt inhibitor, Dkk1, under control of the osteoblast specific 2.3kb fraction of the collagen1α promoter. This promoter has been previously shown to direct transgene expression to osteoblastic cells, resulting in changes in the number and function of HSCs (Calvi et al, 2001; Calvi et al, 2003)
We noted very little overt phenotype in the hematopoietic compartment of the Dkk1 tg mice at steady-state, and confirmed that transgene expression did not extend to the primitive hematopoietic fraction itself. Clear alterations of bone morphology were observed, however, including a 20% decrease in trabecular bone (manuscript in preparation). Despite the absence of a steady-state hematopoetic phenotype, TCF/LEF activity was specifically reduced within the HSC-containing fraction of Dkk1 transgenic mice, and stem cell function was altered under specific conditions. For example, a highly significant defect in the maintenance of reconstitution potential of HSC was observed, either in settings of serial transplant, or following secondary transplantation of wildtype donor cells previously used to reconstitute Dkk1 tg hosts. In agreement with the functional data, HSC populations had a marked reduction of cells within the G0 fraction of the cell cycle, and displayed enhanced sensitivity to 5-fluorouracil treatment. Wnt signals therefore appear to participate in mediating HSC quiescence in vivo, a result that was largely unpredicted from previous studies, although recent analysis of Hmgb3 mutant mice also supports this conclusion (Nemeth et al., 2006). Our results highlight the importance of studying the impact of a signaling pathway over long-term experiments, and in a physiologic context when seeking to resolve the effects of manipulations on HSC function. In that context, Wnt signaling plays an unanticipated role in maintaining HSC quiescence, which may underlie its requirement in preserving the self-renewing capability of HSC.
Osteoblast expression of Dkk1 does not affect blood or marrow primitive hematopoietic cell populations at steady state
The Wnt inhibitor, Dkk1, has been shown to play an important role in bone formation during development (Niehrs, 2006), and is normally expressed by osteoblasts (Grotewold et al., 1999; MacDonald et al., 2004), hence may have regulatory roles as part of the endosteal HSC niche. To examine the impact of Wnt inhibition on hematopoietic stem cells localized to the periendosteal region, Dkk1 was overexpressed within osteoblastic lineage cells under the control of the truncated 2.3kb collagen 1α promoter (manuscript in preparation). Resulting Col1α2.3-Dkk1 transgenic (Dkk1 tg) mice were backcrossed for at least 5 generations to the C57Bl/6 background and examined for bone and blood phenotypic alterations. No significant differences in peripheral white or red blood cell counts were observed (figure S1a). Bone marrow (BM) and spleen cellularity were also unchanged when Dkk1 tg mice and their littermates were compared, although a slight but not significant trend towards reduced body weight and BM cellularity was apparent in transgenic mice (figure S1b and data not shown). In contrast, significant alterations in bone morphology were observed, as is reported elsewhere (manuscript in preparation, and (Li et al, 2006)) Of note, trabecular bone volume was reduced by approximately 20%, whereas cortical bone was unaffected in Dkk1 tg mice (data not shown). Trabecular bone has been shown by us and others to affect HSC number and function (Adams et al, 2007; Calvi et al, 2003; Jung et al, 2007; Zhang et al, 2003). A panel of antibodies using 7 different flurochromes was used for multiparametric analysis of primitive precursors within the BM of Dkk1 tg mice and their littermates, including populations of LT-HSC, ST-HSC, CMP, GMP, MEP and CLP (figure 1a,c). Subpopulations containing primitive HSCs were not significantly altered at steady-state (figure 1b). However, additional cell surface markers revealed a slight but significant increase in the population containing phenotypically-defined common lymphoid progenitors (figure 1d). The calculated absolute cell numbers based on these frequencies indicated a similar pattern of results (figure S2). Despite the elevation of early lymphoid progenitors in the BM of Dkk1 tg mice, no significant changes were observed in the relative proportion of early B lineage progenitor subsets in the BM (data not shown).
Seven color FACS analysis of primitive populations in wt and Dkk1-tg BM nihms-240191-f0001
Seven color FACS analysis of primitive populations in wt and Dkk1-tg BM
Figure 1Seven color FACS analysis of primitive populations in wt and Dkk1-tg BM
BM from Dkk1 tg and littermates was assayed by multiparameter FACS for relative proportion of primitive HSC populations. BM was stained with antibodies against Lineage markers, cKit, Sca-1, CD34, Flk2, CD16/32 and CD127 and gated as shown in panels (A) and (C). At least 10 mice per genotype were compared, over at least three separate experiments. The proportion of BM corresponding to the HSC-containing LK+S+ fraction (A, blue gate) is shown in (B, left axis), and is sub-sectioned according to CD34 and Flk2 expression to yield phenotypic assessments of LT-HSC and ST-HSC fractions (B, right axis). More differentiated progenitors gated in the LK+S− population (A, left, green gate) were sub-sectioned based on CD16/32 and CD34 expression to compare CMP, GMP and MEP progenitors as shown in (C, left panel). Frequencies of each population, from the same samples quantified for HSC frequency in (B) are shown in (D, left axis). The CLP fraction, gated on LKloSlo in (A, red gate), and gated further on CD127+ cells in (C, right panel) are quantified in (D, right axis). Significance was determined by a Student’s 2-tailed T-test. Error bars indicate the SE of the mean.
Dkk1-tg HSCs exhibit impaired Wnt signaling in a non-cell autonomous manner
To confirm that the transgenic expression of Dkk1 leads to the inhibition of Wnt/βcatenin signaling in the Dkk1 tg mice, HSC-containing populations were isolated from Dkk1 transgenic mice that had been intercrossed with the Topgal reporter strain. In these Topgal mice (DasGupta and Fuchs, 1999), multiple TCF/LEF binding sites have been inserted to control the expression of the reporter gene, β-galactosidase. Reporter activity using this construct has been shown to correlate with canonical Wnt signaling. Of note, TCF/LEF transcription has recently been shown to proceed even with the combined loss of β-catenin and γ-catenin, suggesting that canonical Wnt signals can be transduced by alternate intermediates (Jeannet et al, 2007). Reporter activity was examined within the LK+S+ (Lineage-cKit+Sca1+), HSC-containing population, and the LK+S− population which is devoid of LT-HSC potential. When the Wnt reporter activity detected in each of these populations was compared, a dramatic reduction (>100 fold reduction) in β-catenin activation was observed in the HSC-containing LK+S+ population isolated from Dkk1 tg mice (figure 2a). A more modest reduction (<5 fold reduction) was observed in the less-actively signaling LK+S− fraction. This finding indicates that despite the unchanged frequency of phenotypically-defined HSC-containing populations in unmanipulated Dkk1 tg animals, there is evidence that these cells are molecularly altered by osteoblast expression of the Wnt inhibitor. These data provide evidence for direct inhibition of Wnt signaling in the HSC population in addition to any effects that might be mediated by decreased trabecular bone mass. Wnt signaling is regulated, in part, via a negative feedback loop by TCF/LEF-dependent transcription of endogenous Dkk1 (Niida et al, 2004). Consistent with the decrease in Topgal reporter activity, expression of endogenous Dkk1 was also inhibited in the LK+S+ population of Dkk1 tg mice (figure 2b). Using primers specific for the Dkk1 tg, and in comparison its expression in wt and Dkk1 tg tibea, sorted LK+S+ cells do not express the Dkk1 transgene (figure 2c). Together, these results confirm that Dkk1 tg mice inhibit Wnt signaling specifically within the HSC compartment in a non-cell autonomous manner.
Assessment of canonical Wnt signal activity in HSC-containing populations of Dkk1-tg mice nihms-240191-f0002
Assessment of canonical Wnt signal activity in HSC-containing populations of Dkk1-tg mice
Figure 2Assessment of canonical Wnt signal activity in HSC-containing populations of Dkk1-tg mice
Functional impact of the Dkk1 transgene on BM reconstitution
Analysis of stem/progenitor activity cannot rely exclusively on the quantitation of precursors according to phenotypically-defined parameters. Using functional measures, we detected a consistent defect in multilineage and myeloid colony formation on a per cell basis in BM isolated from Dkk1 transgenic mice (figure 3a). This result was despite the absence of significant alteration of myeloid and more primitive progenitors by immunophenotype, possibly reflecting the elevated lymphoid fraction, whose progeny are not read out under these culture conditions. In vitro methods such as the CFU assay offer an entry-level analysis of hematopoietic activity, however functional reconstitution in vivo more accurately examines true HSC function (Purton and Scadden, 2007). Therefore, in order to better assess the functional capacity of HSCs isolated from the Dkk1 transgenic environment, BM was transplanted from wt or Dkk1 tg littermates with an equivalent dose of competing marrow from congenic donor mice into lethally irradiated recipients. Donor marrow was isolated from a single wt or transgenic mouse to assess any individual-to-individual variation. Following six months of engraftment, no significant changes in reconstitution were observed across the groups of recipients receiving BM isolated from individual wt or Dkk1 tg environments, although a range of reconstitution capacity was apparent in both groups (figure S3a). Using a limiting dilution assay to determine the frequency of repopulating cells present in BM isolated from individual Dkk1-expressing animals revealed a two-fold elevation in the number of functional reconstituting HSCs (Figure 3b). These transplant results indicate that cells isolated from the Dkk1-epressing niche are capable of reconstituting irradiated recipients, and appear to be present at a higher frequency when Wnt has been inhibited in this location. An important additional parameter to test when investigating HSC function is their longevity, or ability to respond to repeated rounds of expansion stress. To assay the longevity of HSCs isolated from Dkk1 tg mice, noncompetitive serial transplants were performed. As expected from the previous transplant experiments, Dkk1 tg BM was able to completely reconstitute wt irradiated recipients (data not shown).
Functional assessment of HSCs isolated from Dkk1-tg mice
Figure 3Functional assessment of HSCs isolated from Dkk1-tg mice
(A) BM from 8 pairs of wt and Dkk1-tg mice was plated in methylcellulose with growth factors (SCF, IL-3, IL-6, Epo) and scored for CFU-C (combined scoring for BFU-E, CFU-GM and CFU-GEMM colonies) after 12 days. All live colonies of more than 30 cells were counted for each of three wells plated per sample. Data are shown as mean colonies per well for each of 8 mice studied over three individual experiments. Significance was determined using a two-tailed Student’s T test. (B) Limiting dilution experiments were performed using three doses of test marrow (CD45.1) transplanted with 5×105 competing cells (CD45.2) into groups of at least 9 recipients (CD45.2) per dose. Test marrow was isolated from two wt and two Dkk1-tg mice, and the Dkk1-tg donors shown here were transplanted into separate groups of irradiated recipients. Data points are plotted as the percent of recipients per group that did not exhibit at least 1% multi-lineage PB engraftment at 6 months (percent unreconstituted). LT-HSC frequency and significance were determined using Poisson statistics: wt, 1 in 63,00 (circles) vs tg, 1 in 31,500 or 1: 37,000 (squares); p<0.02. Similar results were obtained in an independent assessment of two Dkk1-tg donors. (C) Non-competitive serial transplants were initiated by transplanting 1×106 whole BM pooled from three wt or Dkk1-tg donors (CD45.1) into each of 10 irradiated recipients (CD45.2). Secondary and tertiary transplants were performed after 14 weeks of engraftment by pooling BM from 3-4 reconstituted recipients to transplant 1×106 whole BM into new groups of 10 irradiated CD45.2 recipients. The Kaplan-Meier survival graph depicts the survival of tertiary recipients, mice receiving BM from Dkk1-tg mice (solid line) or wt controls (dashed line). Similar results were obtained in an independent assessment of 2 wt and 2 Dkk1-tg mice. (D) Prior to transplant into tertiary recipients, BM from 5 secondary recipients of both genotypes was assayed by FACS for the frequency of LT-HSCs (LK+S+CD34loFlk2−). Error bars indicated SD of the mean, and significance was determined by a two-tailed T test
Effect of temporary exposure to endosteal Dkk1 on HSC function
Figure 6Gene expression by quantitative PCR of sorted primitive populations
Understanding the role of specific signals in the varied regulatory functions of HSC activities is crucial for designing and developing therapeutic interventions involving these cells. The impact of the Wnt family on the expansion and regulation of hematopoietic cells has been examined in a variety of studies. However, the physiologic effects of this pathway remain somewhat ill-defined with often contradicting results. Some have demonstrated that Wnt cascade activation promotes the proliferation of HSCs and their progeny while maintaining at least short-term functional activity (Baba et al, 2006; Murdoch et al, 2003; Reya et al, 2003;Trowbridge et al, 2006). Others, employing persistent genetic activation of the pathway, have also demonstrated an increase in proliferation of cells with an HSC immunophenotype, but with marked impairment of HSC differentiation resulting in animal death (Kirstetter et al, 2006; Scheller et al, 2006). However, induced deletion of β-catenin, the primary downstream mediator of the Wnt cascade resulted in no apparent impact on HSC activity, even in a reconstitution assay that required expansion of β-catenin null transplanted HSCs (Cobas et al, 2004). Furthermore, recent combined deletions of both β-catenin and its homologue, γ-catenin, also maintain HSC function under steady-state and primary reconstitution conditions (Jeannet et al, 2007; Koch et al, 2007).All of these studies have either assayed Wnt activity in broad over- or under-stimulation settings, and the manipulations have been performed on the HSCs themselves, or broadly applied to recipient animals. The context in which morphogens are present is highly relevant to their effect and not previously studied for Wnt effects on hematopoiesis (Trowbridge et al., 2006). Indeed, Wnt ligands can modulate signaling initiated by other Wnt family members, underscoring the concept that context, and different signaling intermediates may have a strong impact on functional outcome (Nemeth et al, 2007).
In the present study, we have established a system that permits the analysis of localized Wnt inhibition, offering the opportunity to assay the impact of chronic or temporary exposure to this inhibited environment. In particular, we have directed expression of the Wnt inhibitor, Dkk1, to a cell population that has been previously demonstrated to exert a regulatory function over HSC activity, and which normally express Dkk-1, albeit at lower levels (Grotewold et al,1999; MacDonald et al, 2004). It should be noted that while an increasing number of reports suggest that phenotypically-identified HSCs inhabit additional physical locations within the bone marrow environment (Hooper et al, 2007; Scadden, 2006), the promoter used in our study has proven to functionally impact the number and activity of HSCs when used to direct modifying signal expression to a population of osteoblastic cells. Given that expression of Dkk1 also results in alterations to bone morphology itself, there is likely to be a dual effect of Dkk1: one altering the niche architecture and the other affecting Wnt signaling in stem/progenitor cells. Our studies demonstrated an effect of Dkk1 overexpression by non-HSCs on Wnt signaling in hematopoietic stem/progenitors, suggesting that this is at least a contributing factor to the phenotype observed. This observation that TCF/LEF reporter activity is reduced, as is expression of endogenous Dkk1, itself a Wnt signaling target (Niida et al, 2004) in BM cells of the transgenic mice indicates altered canonical Wnt signaling. It does not rule out that Dkk1 may exert additional Wnt-independent functions. The results presented here also indicate that the reduced longevity of HSCs does not require constant exposure to exogenous Dkk1, given that we were unable to detect Dkk1 tg expression within populations of primitive hematopoietic cells, and therefore the functional impact on transplanted cells is observed in a Dkk tg-free environment. It is important to note that transplantation of whole BM populations is generally not effective at engrafting non-hematopoietic cells (Koc et al, 1999).
Wnt mediates HSC quiescence and maintains reconstitution function in vivo
The results presented here establish a role for Wnt, in the maintenance of a quiescent fraction of functional HSCs in BM. This was associated with evidence of increased stem cells on limit dilution transplant analysis. However the ability of the same cells to function after serial rounds of transplantation was drastically reduced. The ability of stem cells to persist under the stress conditions of transplantation requires self-renewal capability that is compromised after Dkk1 exposure
The studies of inducible deletion of β- and γ-catenin noted that they were dispensable for HSC function, however did not include sequential transplants out to the extent where we observed our most dramatic phenotype (Cobas et al, 2004; Jeannet et al, 2007; Koch et al, 2007) Alternatively, it is possible that Dkk1 interferes with HSC function through a process that does not depend on β- or γ-catenin signaling (Jeannet et al, 2007; Niehrs, 2006).
Our results emphasize the importance of studying pathways within the context of other signals present in the natural microenvironment, and underscore the potential for unanticipated functional roles. It is clear that different combinations of signals may have a range of effects depending on the context in which they are received. Indeed, we observed an impact of Wnt-inhibition on the activation of the Notch target, Hes-1, raising the possibility that Notch and Wnt coordinate in vivo to maintain quiescence of HSCs, rather than participating in expansive and/or self-renewal functions (Duncan et al, 2005). Notably, elevated Hes-1 and p21 expression have recently been shown to correlate with the maintenance of quiescence and repopulating function of primitive HSCs (Yu et al, 2006). We noted a highly specific impact of the Dkk1 tg on the stem cell enriched LK+S+ fraction in Wnt-dependent pathway activation and inhibition and the Notch target, Hes-1, or the cell cycle regulator, p21 expression.
The effects of Dkk1 on cell cycling were unanticipated given previous reports of constitutively active β-catenin inducing increased stem/progenitor cell proliferation (Kirstetter et al, 2006; Scheller et al, 2006). However, others found that with deletion of the chromatin binding protein, Hmgb3, Wnt signaling was increased, yet stem cells more readily returned to quiescence after 5-FU challenge than controls. (Nemeth et al, 2006) Both increased and decreased activation of the pathway may therefore alter HSC cycling kinetics. This may again be due to the context differences observed with a microenvironmentally-provided signal in the current study contrasted with cell autonomous activation of the pathway in the prior reports. Alternatively, it may be an example of the complex effects of morphogens, which have dose-dependent actions (Delaney et al, 2005; Kielman et al, 2002; MacDonald et al, 2004). It may be that there is a bi-phasic response of cell cycling to the Wnt pathway and that proper control of stem cell quiescence requires a fine-tuned modulation of intermediate Wnt signaling intensity. This has implications for the potential use of Wnts as mediators of stem cell expansion ex vivo and for interruption of this pathway as an anti-leukemic intervention.
In sum, niche related expression of Dkk1 reveals a role for Wnt signaling in the physiologic regulation of the hematopoietic compartment, altering stem cell cycling and longevity following repeated expansion, or self-renewal. The phenotype observed was sufficiently distinct from what cell-autonomous modifications of the pathway would have predicted to argue for niche specific modeling of exogenous factors’ effects on stem cells. This may be particularly true for members of the locally acting morphogen group of cell modifiers.
7.10.4 Wnt.β-Catenin Signaling in Development and Disease
A remarkable interdisciplinary effort has unraveled the WNT (Wingless and INT-1) signal transduction cascade over the last two decades. Wnt genes encode small secreted proteins that are found in all animal genomes. Wnt signaling is involved in virtually every aspect of embryonic development and also controls homeostatic self-renewal in a number of adult tissues. Germline mutations in the Wnt pathway cause several hereditary diseases, and somatic mutations are associated with cancer of the intestine and a variety of other tissues.
The mouse wnt1 gene, originally named Int-1, was identified in 1982 by Nusse and Varmus as a preferential integration site for the Mouse Mammary Tumor Virus in virally induced breast tumors ( Nusse and Varmus, 1982). When sequenced, the Wnt1 proto-oncogene was seen to encode a secreted protein that is cysteine rich. Subsequently, Drosophila wingless (wg), which controls segment polarity during larval development ( Nüsslein-Volhard and Wieschaus, 1980), was shown to be a fly homolog of Wnt1 ( Rijsewijk et al., 1987). Segmentation of the epidermis of wg mutant fly embryos is severely impaired as evidenced by abnormalities in the overlying ventral cuticle. In contrast to the wild-type cuticle, which exhibits alternating denticle and naked belts, the wg cuticle is completely covered with denticles. Fly embryos carrying mutations in the porcupine, dishevelled, and armadillo genes display similar cuticle abnormalities to wgmutant embryos, whereas mutations in shaggy/zeste-white 3 cause the opposite phenotype, a naked cuticle. Epistatic analysis of cuticle structure in double mutants indicated that these genes constituted the core of a new signal transduction cascade ( Siegfried et al., 1992, Noordermeer et al., 1994 and Peifer et al., 1994).
In 1989, McMahon and Moon (McMahon and Moon, 1989) observed a duplication of the body axis inXenopus following injection of mouse Wnt1 mRNA into ventral blastomeres of embryos at the 4-cell stage. This observation supported the notion that Wnt signaling was shared between vertebrates and invertebrates and, moreover, provided a rapid and convenient assay to study components of the Wnt pathway in vertebrates. Axis duplication was also induced by Dishevelled (Dsh), β-catenin (the vertebrate homolog of armadillo), and a dominant-negative version of glycogen synthase kinase 3 (GSK3), the vertebrate homolog of shaggy/zeste-white 3 ( Dominguez et al., 1995, Guger and Gumbiner, 1995 and He et al., 1995). Although long elusive, the specific Wnt signal that triggers axis induction in Xenopus was identified as Wnt11 by Heasman and colleagues last year ( Tao et al., 2005).
The combined observations made in Drosophila and Xenopus delineated a highly conserved signaling pathway, activated by secreted Wnt proteins. Independent of these studies, the adenomatous polyposis coli (APC) gene was discovered in a hereditary cancer syndrome termed familial adenomatous polyposis (FAP) ( Kinzler et al., 1991 and Nishisho et al., 1991). Soon after, the large cytoplasmic APC protein was found to interact with β-catenin ( Rubinfeld et al., 1993 and Su et al., 1993). This observation provided the first connection between the Wnt pathway and human cancer.
Genome sequencing has since revealed that mammalian species have roughly 20 secreted Wnt proteins, which can be divided into 12 conserved Wnt subfamilies. Of these, only 6 subfamilies have counterparts in ecdysozoan animals such as Drosophila and Caenorhabditis. In contrast, at least 11 of the Wnt subfamilies occur in the genome of a cnidarian (the sea anemone Nematostella vectensis). This finding suggests that some Wnt subfamilies were lost during the evolution of the ecdysozoan lineage but more importantly reveals that a complex inventory of Wnt factors was present in multicellular animals well before the Cambrian explosion (550 million years ago). Thus, comparative genomic analysis underscores the crucial role that Wnt genes play in organismal patterning throughout the animal kingdom ( Kusserow et al., 2005).
Currently, three different pathways are believed to be activated upon Wnt receptor activation: the canonical Wnt/β-catenin cascade, the noncanonical planar cell polarity (PCP) pathway, and the Wnt/Ca2+ pathway. Of these three, the canonical pathway is best understood and is the primary subject of this review. For recent comprehensive overviews on the other Wnt signaling pathways, the reader is referred to Katoh (2005) and Kohn and Moon (2005). This review discusses how Wnt proteins are produced and secreted and how they activate the canonical Wnt signaling pathway in recipient cells. Further, the review examines the roles of the canonical Wnt pathway in development, tissue self-renewal, and cancer.
Wnt Protein Secretion
Wnt proteins are characterized by a high number of conserved cysteine residues. Although Wnt proteins carry an N-terminal signal peptide and are secreted, they are relatively insoluble. This insolubility has been attributed to a particular protein modification, cysteine palmitoylation, which is essential for Wnt function (Willert et al., 2003). Hofmann (2000) reported that a Drosophila gene required in the Wnt-secreting cell, termed porcupine, displays homology to acyl-transferases, enzymes that acylate a variety of substrates in the endoplasmic reticulum. Thus, porcupine and its worm homolog mom-1 are believed to encode the enzyme that is responsible for Wnt palmitoylation ( Zhai et al., 2004).
Recently, Banziger et al. (2006) and Bartscherer et al. (2006) uncovered in Drosophila another conserved gene that is essential for Wnt secretion, named wntless (wls) and evenness interrupted (evi), respectively. The gene encodes a seven-pass transmembrane protein that is conserved from worms (mom-3) to man (hWLS). In the absence of Wls/evi, Wnts are retained inside the cell that produces them. The Wntless protein resides primarily in the Golgi apparatus, where it colocalizes and physically interacts with Wnts. A genetic screen in C. elegans revealed that the retromer, a multiprotein complex involved in intracellular trafficking and conserved from yeast to man, is also essential for Wnt secretion and for the generation of a Wnt gradient ( Coudreuse et al., 2006). An attractive hypothesis is that the retromer complex is involved in recycling a Wnt cargo receptor (such as Wntless) between the default secretory pathway and a compartment dedicated to Wnt secretion (see Figure 1).
To be secreted, Wnt proteins in the endoplasmic reticulum (ER) need to be palmitoylated by the action of Porcupine. Wnt proteins also require Wntless (Wls/Evi) in order to be routed to the outside of the cell. Loading onto lipoprotein particles may occur in a dedicated endo/exocytic compartment. The retromer complex may shuttle Wls between the Golgi and the endo/exocytic compartment.
Wnt is thought to act as a morphogen (that is, a long-range signal whose activity is concentration dependent) (reviewed in Logan and Nusse, 2004). However, it is unclear how these long-range gradients are generated. It is conceivable that the palmitoyl moiety constrains movement away from membranes or lipid particles. Thus, Wnts may be tethered to intercellular transport vesicles or lipoprotein particles (Panakova et al., 2005). Alternatively, Wnts may be transported by cytonemes, which are long, thin filopodial processes. Additionally, studies in Drosophila suggest a role for extracellular heparan sulfate proteoglycans (HSPG) in the transport or stabilization of Wnt proteins. For instance, flies carrying mutations in Dally, a GPI-anchored HSPG, or in genes encoding enzymes that modify HSPGs resemblewingless mutants (reviewed in Lin, 2004).
Receptors, Agonists, and Antagonists for Wnt
Wnts bind Frizzled (Fz) proteins, which are seven-pass transmembrane receptors with an extracellular N-terminal cysteine-rich domain (CRD) (Bhanot et al., 1996). The Wnt-Fz interaction appears promiscuous, in that a single Wnt can bind multiple Frizzled proteins (e.g., Bhanot et al., 1996) and vice versa. In binding Wnt, Fzs cooperate with a single-pass transmembrane molecule of the LRP family known as Arrow inDrosophila ( Wehrli et al., 2000) and LRP5 and -6 in vertebrates ( Pinson et al., 2000 and Tamai et al., 2000). The transport of Arrow/LRP5/6 to the cell surface is dependent on a chaperone called Boca inDrosophila and Mesd in mice ( Culi and Mann, 2003 and Hsieh et al., 2003). And consistent with a role of the Boca/Mesd chaperone in the transport of Arrow/LRP5/6 transport, mutations in Boca and Mesdresemble loss of Arrow/LRP5/6. Although it has not been formally demonstrated that Wnt molecules form trimeric complexes with LRP5/6 and Frizzled, surface expression of both receptors is required to initiate the Wnt signal.
Derailed, a transmembrane tyrosine kinase receptor from the RYK subfamily, is an unusual Wnt receptor.Drosophila Wnt5 controls axon guidance in the central nervous system. Embryos lacking Dwnt-5 resemble those lacking Derailed, that is, they generate aberrant neuronal projections across the midline ( Yoshikawa et al., 2003). Derailed binds DWnt-5 through its extracellular WIF (Wnt inhibitory factor) domain. Signaling events downstream of this alternative Wnt receptor remain unclear. Somewhat unexpectedly, the Derailed kinase domain may be dispensable for signaling. Lu et al. (2004) propose that, unlike the Drosophila Ryk homolog Derailed, mammalian Ryk functions as a coreceptor along with Fz. Mammalian Ryk binds Dishevelled to activate the canonical Wnt/β-catenin signaling pathway. Another tyrosine kinase receptor, Ror2, harbors a Wnt binding CRD motif. Wnt5a can engage Ror2 to inhibit the canonical Wnt signaling pathway, although paradoxically Wnt5a can also activate the canonical pathway by directly engaging Fz4 (Mikels and Nusse, 2006) and Fz5 ( He et al., 1997).
At least two types of proteins that are unrelated to Wnt factors activate the Frizzled/LRP receptors. One of these factors is the cysteine-knot protein Norrin, which is mutated in Norrie disease, a developmental disorder characterized by vascular abnormalities in the eye and blindness. Norrin binds with high affinity to Frizzled-4 and activates the canonical signaling pathway in an LRP5/6-dependent fashion (Xu et al., 2004). Other factors that activate the canonical Wnt signaling pathway are R-spondins, which are thrombospondin domain-containing proteins. In Xenopus, R-spondin-2 is a Wnt agonist that synergizes with Wnts to activate β-catenin ( Kazanskaya et al., 2004). Human R-spondin-1 has been found to strongly promote the proliferation of intestinal crypt cells, a process which involves the stabilization of β-catenin (Kim et al., 2005). Indeed, studies in cultured cells demonstrate that R-spondins can physically interact with the extracellular domains of LRP6 and Fzd8 and activate Wnt reporter genes ( Nam et al., 2006).
The secreted Dickkopf (Dkk) proteins inhibit Wnt signaling by direct binding to LRP5/6 (Glinka et al., 1998). Through this interaction, Dkk1 crosslinks LRP6 to another class of transmembrane molecules, the Kremens (Mao et al., 2002), thus promoting the internalization and inactivation of LRP6. An unrelated secreted Wnt inhibitor, Wise, also acts by binding to LRP (Itasaki et al., 2003), as does the WISE family member SOST (Li et al., 2005 and Semenov et al., 2005).
Soluble Frizzled-Related Proteins (SFRPs) resemble the ligand-binding CRD domain of the Frizzled family of Wnt receptors (Hoang et al., 1996). WIF proteins are secreted molecules with similarity to the extracellular portion of the Derailed/RYK class of transmembrane Wnt receptors (Hsieh et al., 1999). SFRPs and WIFs are believed to function as extracellular Wnt inhibitors (reviewed in Logan and Nusse, 2004) but, depending on context, may also promote signaling by Wnt stabilization or by facilitating Wnt secretion or transport.
Canonical Wnt Signaling
Once bound by their cognate ligands, the Fz/LRP coreceptor complex activates the canonical signaling pathway (Figure 2). Fz can physically interact with Dsh, a cytoplasmic protein that functions upstream of β-catenin and the kinase GSK-3. Wnt signaling controls phosphorylation of Dsh (reviewed in Wallingford and Habas, 2005). However, it remains unclear whether the binding of Wnt to Fz regulates a direct Fz-Dsh interaction, nor is it known how Dsh phosphorylation is controlled or how phosphorylated Dsh functions in Wnt signal transduction.
canonical-wnt-signaling
Figure 2. Canonical Wnt Signaling
(Left panel) When Wnt receptor complexes are not bound by ligand, the serine/threonine kinases, CK1 and GSK3α/β, phosphorylate β-catenin. Phosphorylated β-catenin is recognized by the F box/WD repeat protein β-TrCP, a component of a dedicated E3 ubiquitin ligase complex. Following ubiquitination, β-catenin is targeted for rapid destruction by the proteasome. In the nucleus, the binding of Groucho to TCF (T cell factor) inhibits the transcription of Wnt target genes. (Right panel) Once bound by Wnt, the Frizzled(Fz)/LRP coreceptor complex activates the canonical signaling pathway. Fz interacts with Dsh, a cytoplasmic protein that functions upstream of β-catenin and the kinase GSK3β. Wnt signaling controls phosphorylation of Dishevelled (Dsh). Wnts are thought to induce the phosphorylation of LRP by GSK3β and casein kinase I-γ (CK1γ), thus regulating the docking of Axin. The recruitment of Axin away from the destruction complex leads to the stabilization of β-catenin. In the nucleus, β-catenin displaces Groucho from Tcf/Lef to promote the transcription of Wnt target genes.
Recent studies have indicated that the coreceptor LRP5/6 interacts with Axin through five phosphorylated PPP(S/T)P repeats in the cytoplasmic tail of LRP (Davidson et al., 2005 and Zeng et al., 2005). Wnts are thought to induce the phosphorylation of the cytoplasmic tail of LRP, thus regulating the docking of Axin. GSK3 phosphorylates the PPP(S/T)P motif, whereas caseine kinase I-γ (CK1γ) phosphorylates multiple motifs close to the GSK3 sites. CK1γ is unique within the CK1 family in that it is anchored in the membrane through C-terminal palmitoylation. Both kinases are essential for signal initiation. It remains presently debated whether Wnt controls GSK3-mediated phosphorylation of LRP5/6 (Zeng et al., 2005) or whether CK1γ is the kinase regulated by Wnt (Davidson et al., 2005). When bound to their respective membrane receptors, Dsh and Axin may cooperatively mediate downstream activation events by heterodimerization through their respective DIX (Dishevelled-Axin) domains.
The Cytoplasmic Destruction Complex
The central player in the canonical Wnt cascade is β-catenin, a cytoplasmic protein whose stability is regulated by the destruction complex. The tumor suppressor protein Axin acts as the scaffold of this complex as it directly interacts with all other components—β-catenin, the tumor suppressor protein APC, and the two kinase families (CK1α, -δ, -ɛ and GSK3α and -β [reviewed in Price, 2006]). When WNT receptor complexes are not engaged, CK1 and GSK3α/β sequentially phosphorylate β-catenin at a series of highly conserved Ser/Thr residues near its N terminus (Figure 2). Phosphorylated β-catenin is then recognized by the F box/WD repeat protein β-TrCP, a component of a dedicated E3 ubiquitin ligase complex. As a consequence, β-catenin is ubiquitinated and targeted for rapid destruction by the proteasome (Aberle et al., 1997). Note that the CK1 and GSK3 kinases perform paradoxical roles in the Wnt pathway. At the level of the LRP coreceptor they act as agonists, whereas in the destruction complex they act as antagonists
Although genetic observations imply an essential role for APC in the destruction complex, there is no consensus on its specific molecular activity. APC has a series of 15 and 20 amino acid repeats with which it interacts with β-catenin. Three Axin-binding motifs are interspersed between these β-catenin-binding motifs. Increasing the expression of Axin in cancer cells that lack APC restores the activity of the destruction complex, implying that APC is only essential when Axin levels are limiting. Quantitatively, Axin indeed appears to be the limiting factor (Lee et al., 2003) and may be the key scaffolding molecule that promotes the rapid assembly and disassembly of the destruction complex.
Given that CK1, Dsh, β-TrCP, and GSK3 participate in other signaling pathways, low levels of Axin may insulate the Wnt pathway from changes in the abundance or activity of these signaling components. It has been proposed that APC is required for efficient shuttling and loading/unloading of β-catenin onto the cytoplasmic destruction complex. Both APC and Axin can themselves be phosphorylated by their associated kinases, which changes their affinity for other components of the destruction complex. Our understanding of the relevance of these phosphorylation events in the regulation of Wnt signaling remains incomplete. For a comprehensive discussion of the kinases in the Wnt pathway, the reader is referred to a recent review (Price, 2006)
β-catenin plays a second role in simple epithelia, that is, as a component of adherens junctions. It is an essential binding partner for the cytoplasmic tail of various cadherins, such as E-cadherin (Peifer et al., 1992). Unlike the signaling pool of β-catenin, the pool that is bound to the adherens junction is highly stable. It is currently unclear whether the adhesive and signaling properties of β-catenin are interconnected. In a likely scenario, newly synthesized β-catenin first saturates the pool that is part of the adhesion junction, which never becomes available for signaling. “Excess,” free cytoplasmic β-catenin protein is then efficiently degraded by the APC complex. It is only this second, highly unstable pool that is subject to regulation by Wnt signals. In support of this model, these two functions of β-catenin are separately performed by two different β-catenin homologs in C. elegans ( Korswagen et al., 2000).
Upon receptor activation by WNT ligands, the intrinsic kinase activity of the APC complex for β-catenin is inhibited. It is unclear how this occurs, but it likely involves the Wnt-induced recruitment of Axin to the phosphorylated tail of LRP and/or to Fz-bound Dsh. As a consequence, stable, nonphosphorylated β-catenin accumulates and translocates into the nucleus, where it binds to the N terminus of LEF/TCF (lymphoid enhancer factor/T cell factor) transcription factors (Behrens et al., 1996, Molenaar et al., 1996 and van de Wetering et al., 1997).
It has been suggested that protein phosphatases may regulate β-catenin stability as antagonists of the serine kinases (reviewed in Price, 2006). For example, heterotrimeric PP2A is required for the elevation of β-catenin levels that is dependent on Wnt. Moroever, PP2A can bind Axin and APC, suggesting that it might function to dephosphorylate GSK3 substrates. If and how PP2A activity is regulated by Wnt signals remains to be resolved.
Crystallographic studies are starting to provide insights into the structure of the destruction complex. The central region of β-catenin (to which most partners bind) was the first component of the pathway to be crystallized. It consists of 12 armadillo repeats, which adopt a superhelical shape with a basic groove running along its length. Subsequently, structural interactions of Axin, APC, E-cadherin, and TCF with β-catenin have been visualized (Choi et al., 2006, and references therein). APC, E-cadherin, and TCF bind the central part of the basic groove in a mutually exclusive fashion. Despite very limited conservation of primary sequence in the respective interaction domains, the modes of binding are structurally very similar. Axin utilizes a helix that occupies the groove formed by the third and fourth armadillo repeats of β-catenin. Axin binding precludes the simultaneous interaction with other β-catenin partners in this region. Based on this observation, it is suggested that a key function of APC is to remove phosphorylated β-catenin from the active site of the complex (Xing et al., 2003). In a further study, the structure of Axin bound to APC (Spink et al., 2000) was solved. These studies form stepping stones to a better understanding of the dynamics of the destruction complex. Unfortunately, biochemical studies of the destruction complex in its different activation states are sorely lacking.
Nuclear Events
Upon stabilization by Wnt signals, β-catenin enters the nucleus to reprogram the responding cell (Figure 3). There is no consensus on the mechanism by which β-catenin travels between the cytoplasm and the nucleus. In many cases, cells that undergo Wnt signaling may actually display an overall rise in β-catenin protein without a clear nuclear preference. β-catenin’s nuclear import is independent of the Nuclear Localization Signal/importin machinery. β-catenin itself is a close relative of importin/karyopherins and directly interacts with nuclear pore components. Two proteins, Tcf and Pygopus are proposed to anchor β-catenin in the nucleus, although β-catenin can still localize to the nucleus in the absence of either of the two (reviewed in Staedeli et al., 2006). β-catenin can also be actively transported back to the cytoplasm, by either an intrinsic export signal or as cargo of Axin (Cong and Varmus, 2004) or APC (Rosin-Arbesfeld et al., 2000) that shuttle between cytoplasm and nucleus.
The β-catenin/Tcf complex interacts with a variety of chromatin-remodeling complexes to activate transcription of Wnt target genes. The recruitment of β-catenin to Tcf target genes affects local chromatin in several ways. Bcl9 acts as a bridge between Pygopus and the N terminus of β-catenin. Evidence suggests that this trimeric complex is involved in nuclear import/retention of β-catenin (Townsley et al., 2004), but it may also be involved in the ability of β-catenin to activate transcription (Hoffmans et al., 2005). The C terminus of β-catenin also binds to coactivators such as the histone acetylase CBP, Hyrax, and Brg-1 (a component of the SWI/SNF chromatin-remodeling complex).
Whereas the fly and worm genomes both encode a single Tcf protein, the vertebrate genome harbors fourTcf/Lef genes. Tcf factors bind their cognate motif in an unusual fashion, i.e., in the minor groove of the DNA helix, while inducing a dramatic bend of over 90°. Tcf target sites are highly conserved between the four vertebrate Tcf/Lef proteins and Drosophila Tcf. These sites resemble AGATCAAAGG ( van de Wetering et al., 1997). Wnt/TCF reporter plasmids such as pTOPflash ( Korinek et al., 1997), widely used to measure Wnt pathway activation, consist of concatamers of 3–10 of these binding motifs cloned upstream of a minimal promoter. The four vertebrate TCF/LEF differ dramatically in their embryonic and adult expression domains, yet they are highly similar biochemically, explaining the extensive redundancy unveiled in double knockout experiments (as in Galceran et al., 1999).
In the absence of Wnt signals, Tcf acts as a transcriptional repressor by forming a complex with Groucho/Grg/TLE proteins (Cavallo et al., 1998 and Roose et al., 1998). The interaction of β-catenin with the N terminus of Tcf (Behrens et al., 1996, Molenaar et al., 1996 and van de Wetering et al., 1997) transiently converts it into an activator, translating the Wnt signal into the transient transcription of Tcf target genes. To accomplish this, β-catenin physically displaces Groucho from Tcf/Lef (Daniels and Weis, 2005). The recruitment of β-catenin to Tcf target genes affects local chromatin in several ways. Its C terminus is a potent transcriptional activator in transient reporter gene assays (van de Wetering et al., 1997). It binds coactivators such as the histone acetylase CBP and Brg-1, a component of the SWI/SNF chromatin remodeling complex (reviewed in Staedeli et al., 2006). A recent study implies that the human and fly homologs of yeast Cdc37 (Parafibromin and Hyrax, respectively) also interact with the C-terminal transactivation domain of β-catenin to activate target gene transcription (Mosimann et al., 2006). Cdc37 is a component of the PAF complex. In yeast the PAF complex directly interacts with RNA polymerase II to regulate transcription initiation and elongation.
Two dedicated, nuclear partners of the TCF/β-catenin complex, Legless/Bcl9 and Pygopus, were recently found in genetic screens in Drosophila ( Kramps et al., 2002, Parker et al., 2002 and Thompson et al., 2002). Mutations in these genes result in phenotypes similar to wingless, and overexpression of both genes promotes TCF/β-catenin activity in mammalian cells ( Thompson et al., 2002). Bcl9 bridges Pygopus to the N terminus of β-catenin. The formation of this trimeric complex has been implicated in nuclear import/retention of β-catenin ( Townsley et al., 2004) but may also directly contribute to the ability of β-catenin to transactivate transcription ( Hoffmans et al., 2005). Although most if not all Wnt signaling events in Drosophila appear to be dependent on Bcl9 and Pygopus, it is currently unclear if this holds true in vertebrate development.
Tcf itself can be regulated by phosphorylation. The MAP kinase-related protein kinase NLK/Nemo (Ishitani et al., 1999) phosphorylates Tcf, thereby decreasing the DNA-binding affinity of the β-catenin/Tcf complex and inhibiting transcriptional regulation of Wnt target genes. In C. elegans, LIT-1/NLK-dependent phosphorylation results in PAR-5/14-3-3- and CRM-1-dependent nuclear export of POP-1/Tcf ( Meneghini et al., 1999 and Lo et al., 2004). And lastly, a recent study utilizing chromatin immunoprecipitations suggests that APC, independent of its role in the cytoplasmic destruction complex, acts on chromatin to facilitate CtBP-mediated repression of Wnt target genes in normal, but not in colorectal cancer cells ( Sierra et al., 2006).
Wnt Target Genes
Loss of components of the Wnt pathway can produce dramatic phenotypes that affect a wide variety of organs and tissues. A popular view equates Wnt signaling with maintenance or activation of stem cells (Reya and Clevers, 2005). It should be realized, however, that Wnt signals ultimately activate transcriptional programs and that there is no intrinsic restriction in the type of biological event that may be controlled by these programs. Thus, Wnt signals may promote cell proliferation and tissue expansion but also control fate determination or terminal differentiation of postmitotic cells. Sometimes, these disparate events, proliferation and terminal differentiation, can be activated by Wnt in different cell types within the same structure, such as the hair follicle or the intestinal crypt (Reya and Clevers, 2005).
Numerous Tcf target genes have been identified in diverse biological systems. These studies tend to focus on target genes involved in cancer, as exemplified by the wide interest in the Wnt target genes cMyc and Cyclin D1. For a comprehensive, updated overview of Tcf target genes, the reader is referred to the Wnt homepage (http://www.stanford.edu/∼rnusse/wntwindow.html). The Wnt pathway has distinct transcriptional outputs, which are determined by the developmental identity of the responding cell, rather than by the nature of the signal. In other words, the majority of Wnt target genes appear to be cell type specific. It is not clear whether “universal” Wnt/Tcf target genes exist. The best current candidates in vertebrates are Axin2/conductin (Jho et al., 2002) and SP5 (Weidinger et al., 2005). As noted (Logan and Nusse, 2004), Wnt signaling is autoregulated at many levels. The expression of a variety of positive and negative regulators of the pathway, such as Frizzleds, LRP and HSPG, Axin2, and TCF/Lef are all controlled by the β-catenin/TCF complex.
Wnt Signaling in Self-Renewing Tissues in Adult Mammals
Wnt signaling not only features in many developmental processes; in some self-renewing tissues in mammals it remains essential throughout life. It is this aspect of Wnt signaling that is intricately connected to the development of disease. The examples discussed below illustrate how the Wnt pathway is involved in adult tissue self-renewal. Mutations in the Wnt pathway tip the homoeostatic balance in these tissues to cause pathological conditions such as disturbances in skeletal bone mass or cancer.
Gut
Figure 4. Self-Renewing Tissues in the Adult Mammal
Current evidence indicates that the Wnt cascade is the dominant force in controlling cell fate along the crypt-villus axis. In neonatal mice lacking Tcf4, the differentiated villus epithelium appears unaffected, but the crypt progenitor compartment is entirely absent (Korinek et al., 1998). This implies that physiological Wnt signaling is required for the establishment of this progenitor compartment.
Hair Follicle
Multipotent epidermal stem cells reside in the bulge region of the hair follicle (Figure 4). Bulge stem cells can generate all hair lineages but also the sebocytes and even the stem cells of the interfollicular epidermis (Alonso and Fuchs, 2003). To form a hair, cells migrate downward from the bulge through the outer root sheath. At the base of the hair, the cells enter a transit-amplifying compartment termed the germinative matrix where they undergo terminal differentiation in the precortex compartment of the hair.
Hematopoietic System
Hematopoietic stem cells (HSCs) are the best studied stem cells in mammals. A number of studies have implicated the Wnt signaling pathway as an important regulator of hematopoietic stem and progenitor cells. HSCs themselves as well as the bone marrow microenvironment can produce Wnt proteins. Indeed, Tcf reporters are active in HSCs in their native microenvironment.
Bone
In postnatal and adult life, osteoblasts produce bone matrix, whereas osteoclasts resorb the matrix. Bone density is determined by the relative activities of these two cell types. Gain-of-function mutations in the human LRP5 gene occur in bone diseases, indicating that canonical Wnt signaling may regulate bone mass. This observation has motivated genetic studies in mouse models, which generally confirm the importance of this signaling pathway in bone homeostasis, primarily as a positive regulator of the osteoblast lineage. Similar to humans carrying the gain-of-function LRP5G171V mutation, transgenic mice expressing this allele in osteoblasts display increased bone density and elevated numbers of active osteoblasts (reviewed in Hartmann, 2006).
Wnt Signaling in Cancer
Colon Cancer
The APC gene was among the first tumor suppressors to be cloned. A germline APC mutation is the genetic cause of a hereditary cancer syndrome termed Familiar Adenomatous Polyposis (FAP) (Kinzler et al., 1991 and Nishisho et al., 1991). FAP patients inherit one defective APC allele and as a consequence develop large numbers of colon adenomas, or polyps, in early adulthood. Polyps are benign, clonal outgrowths of epithelial cells in which the second APC allele is inactivated. Inevitably, some of these polyps progress into malignant adenocarcinoma. Loss of both APC alleles occurs in the large majority of sporadic colorectal cancers (Kinzler and Vogelstein, 1996). Mutational inactivation of APC leads to the inappropriate stabilization of β-catenin (Rubinfeld et al., 1996; Figure 4). Indeed, Tcf reporter constructs, normally transcribed only upon Wnt signaling, are inappropriately transcribed in APC mutant cancer cells through the action of constitutive complexes between β-catenin and the intestinal TCF family member Tcf4 (Korinek et al., 1997). In rare cases of colorectal cancer where APC is not mutated, Axin2 is mutant (Liu et al., 2000), or activating (oncogenic) point mutations in β-catenin remove its N-terminal Ser/Thr destruction motif (Morin et al., 1997). Of note, patients with hereditary Axin2 mutations display a predisposition to colon cancer (Lammi et al., 2004).
In intestinal epithelial cells in which APC is mutated, the constitutive β-catenin/Tcf4 complex activates a genetic program in crypt stem/progenitor cells (van de Wetering et al., 2002). In the crypt, the Wnt signaling gradient drives expression of this genetic program to maintain progenitor cell proliferation. The Wnt gradient also controls expression of the EphB/EphrinB sorting receptors and ligands (Battle et al., 2002). The resulting EphB/EphrinB countergradients establish crypt-villus boundaries as well as position the Paneth cells at the bottom of the crypt. Several EphB genes are initially upregulated as Wnt/Tcf4 target genes in early adenomas, but their expression is lost upon cancer progression (Batlle et al., 2005) apparently as the result of a selection process. Activating Wnt pathway mutations are not restricted to cancer of the intestine. Loss-of-function mutations in Axin have also been found in hepatocellular carcinomas, whereas oncogenic β-catenin mutations occur in a wide variety of solid tumors (reviewed inReya and Clevers, 2005).
Several animal models exist for FAP. Dove and colleagues first described the multiple intestinal neoplasia(min) mouse, which carries a stop codon in APC (Apcmin). Unlike FAP patients, Apcmin mice develop adenomas predominantly in the small intestine ( Su et al., 1992). Several additional Apc knockout models have been generated in mice. Invariably, these mice develop neoplastic lesions but they may differ in tumor incidence and tissue type in which tumors first appear. In a recent elegant study, the Wnt cascade was mutationally activated in adult mice by conditional deletion of Apc ( Sansom et al., 2004). Within days, villi were entirely populated by crypt-like cells, demonstrating the direct link between active Wnt signaling and the proliferation of crypt progenitors, which when unrestrained results in cancer. Zebrafish that are mutant in Apc resemble the mouse models in that heterozygous mutants develop adenomas in organs of endodermal origin including the intestine. These fish may prove useful for genetic screens for genes that modify cancer risk ( Haramis et al., 2006).
Hair Follicle Tumors
Leukemia
Drawing from the parallels between self-renewal and cancer in the gut and hair follicle, the effects of Wnt pathway components on hematopoietic progenitors predict that Wnt deregulation may contribute to hematological malignancies. Indeed, a recent report suggests that leukemic growth of both myeloid and lymphoid lineages is dependent on Wnt signaling. Granulocyte-macrophage progenitors from Chronic Myelogenous Leukemia patients and blast crisis cells from patients resistant to therapy display active Wnt signaling as demonstrated by Tcf reporter activity and the accumulation of nuclear β-catenin (Jamieson et al., 2004).
Over the last 20 years, a detailed outline of the canonical Wnt pathway has emerged. Although it is likely that most core components of the pathway have now been identified, much remains to be learned about the biochemical events that connect these components. Many of the gaps in our knowledge are due to the notorious difficulties in the production of purified Wnt proteins. Few good Wnt antibodies exist and, 25 years after the cloning of Wnt1, its structure remains unknown. The routing and the coincident posttranslational modifications of Wnt proteins in the secreting cell are incompletely understood. And the rules that dictate the movement of Wnt proteins between cells remain uncertain. However, a procedure to produce soluble Wnt has recently been developed (Willert et al., 2003), which creates avenues to address many of these issues.
The components of the destruction complex have been long known, yet the biochemistry of its activity has remained elusive. APC is an essential component of the destruction complex, but what is its biochemical activity? How relevant is Dsh for the coupling of Wnt receptors to the destruction complex? And what mechanism inhibits the phosphorylation of β-catenin by the destruction complex when a Wnt signal is being transduced?
In addition, a multitude of proposed pathway components, not discussed here, may activate, modify, or inhibit Wnt signaling or may be involved in crosstalk to other pathways. An updated, comprehensive list of these putative components and interactions appears on http://www.stanford.edu/∼rnusse/wntwindow.html. Often based on single studies, these candidate components remain to be independently confirmed.
Wnt signaling ultimately controls developmental fates through the transcription of cell type-specific programs of Tcf target genes. Recent developments in array-based technology allow detailed analysis of the nuclear transcriptional response to Wnt signals. With these technologies, it is expected that the dissection of the gene programs in various developmental or pathological events will provide a wealth of insight into the biology of these processes.
7.10.5 Wnt.β-Catenin Signaling. Components, Mechanisms, and Diseases
Signaling by the Wnt family of secreted glycolipoproteins via the transcription co-activator β-catenin controls embryonic development and adult homeostasis. Here we review recent progresses in this so-called canonical Wnt signaling pathway. We discuss Wnt ligands, agonists and antagonists and their interactions with Wnt receptors. We also dissect critical events that regulate β-catenin stability from Wnt receptors to the cytoplasmic β-catenin destruction complex, and nuclear machinery that mediates β-catenin-dependent transcription. Finally we highlight some key aspects of Wnt/β-catenin signaling in human diseases including congenital malformations, cancer and osteoporosis and potential therapeutic implications.
Signaling by the Wnt family of secreted glycolipoproteins is one of the fundamental mechanisms that direct cell proliferation, cell polarity and cell fate determination during embryonic development and tissue homeostasis (Logan and Nusse, 2004). As a result, mutations in the Wnt pathway are often linked to human birth defects, cancer and other diseases (Clevers, 2006). A critical and most studied Wnt pathway is canonical Wnt signaling, which functions by regulating the amount of the transcriptional co-activator β-catenin that controls key developmental gene expression programs. This review focuses on our current understanding of Wnt/β-catenin signaling, drawing mainly from genetic, developmental and biochemical analyses in Drosophila, Xenopus, mice and humans. For more comprehensive and historic perspective we refer readers to earlier reviews (Clevers, 2006; Logan and Nusse, 2004) and the Wnt homepage (www.stanford.edu/~rnusse/wntwindow.html). The nematode Caenorhabditis elegans exhibits similar but also divergent Wnt/β-catenin pathways, which are covered elsewhere (Mizumoto and Sawa, 2007) and in the accompanying review (Kimble 2009). Wnt also activates a number of non-canonical signaling pathways that are independent of β-catenin and have been recently reviewed (Seifert and Mlodzik, 2007; Wang and Nathans, 2007).
The central logic of Wnt/β-catenin signaling has emerged from two decades of studies (Figure 1). In the absence of Wnt, cytoplasmic β-catenin protein is constantly degraded by the action of the Axin complex, which is composed of the scaffolding protein Axin, the tumor suppressor adenomatous polyposis coli gene product (APC), casein kinase 1 (CK1), and glycogen synthase kinase 3 (GSK3). CK1 and GSK3 sequentially phosphorylate the amino terminal region of β-catenin, resulting in β-catenin recognition by β-Trcp, an E3 ubiquitin ligase subunit, and subsequent β-catenin ubiquitination and proteasomal degradation (He et al., 2004). This continual elimination of β-catenin prevents β-catenin from reaching the nucleus, and Wnt target genes are thereby repressed by the DNA-bound T cell factor/lymphoid enhancer factor (TCF/LEF) family of proteins (Figure 1a). The Wnt/β-catenin pathway is activated when a Wnt ligand binds to a seven-pass transmembrane Frizzled (Fz) receptor and its co-receptor, low-density lipoprotein receptor related protein 6 (LRP6) or its close relative LRP5. The formation of a likely Wnt-Fz-LRP6 complex together with the recruitment of the scaffolding protein Dishevelled (Dvl) results in LRP6 phosphorylation and activation and the recruitment of the Axin complex to the receptors. These events lead to inhibition of Axin-mediated β-catenin phosphorylation and thereby to the stabilization of β-catenin, which accumulates and travels to the nucleus to form complexes with TCF/LEF and activates Wnt target gene expression (Figure 1b).
Wnts are conserved in all metazoan animals. In mammals, complexity and specificity in Wnt signaling are in part achieved through 19 Wnt ligands, which are cysteine rich proteins of approxiamately 350-400 amino acids that contain an N-terminal signal peptide for secretion. Murine Wnt3a represents the first purified and biochemically characterized Wnt protein (Willert et al., 2003) owing to its relatively efficient secretion (in contrast to most other Wnt proteins). In addition to N-linked glycosylation, which is required for Wnt3a secretion (Komekado et al., 2007), Wnt3a undergoes two types of lipid modifications that likely account for the hydrophobicity and poor solubility of Wnt proteins (Hausmann et al., 2007). The first reported lipididation was the addition of palmitate to cysteine 77 (Willert et al., 2003). Its mutation had minimal effect on Wnt3a secretion but diminished the ability of Wnt3a to activate β-catenin signaling (Galli et al., 2007;Komekado et al., 2007; Willert et al., 2003). The second identified lipididation was a palmitoleoyl attached to serine 209, and its mutation resulted in Wnt3a accumulation in the endoplasmic reticulum (ER) and failure in secretion (Takada et al., 2006).
Drosophila Wingless (Wg) is the Wnt molecule most investigated in vivo (Hausmann et al., 2007). These studies plus work in nematodes have identified genes that regulate Wnt biogenesis and secretion. Porcupine (Porc) encodes a multipass transmembrane ER protein that contains an O-acyl transferase domain suggesting a role in Wg lipid modification (Hausmann et al., 2007). Porc deficiency results in Wg and Wnt3a accumulation in the ER and diminished Wnt3a palmitoleoylation at serine 209 (Takada et al., 2006), suggesting that Porc is responsible for this particular lipidation. Whether Porc or a distinct acyltransferase is involved in Wnt3a palmitoylation at cysteine 77 remains unknown.
Two additional proteins/protein complexes were identified for Wg/Wnt secretion: Wntless (Wls), also known as Evenness interrupted (Evi) or Sprinter (Srt), in Drosophila and the retromer complex in nematodes (Hausmann et al., 2007). Wls is a multipass transmembrane protein that localizes to the Golgi, endocytic compartments and the plasma membrane, and is essential for Wg secretion. The retromer complex, which is composed of five subunits, was defined first in yeast. It mediates membrane protein trafficking between endosomes and the Golgi apparatus (Hausmann et al., 2007). Several groups recently reported that the retromer complex is required for retrieval/recycling of Wls from the endosome to the Golgi (Belenkaya et al., 2008; Franch-Marro et al., 2008b; Pan et al., 2008a; Port et al., 2008; Yang et al., 2008), likely mediated by direct interaction between Wls and the retromer Vps35 subunit. Loss of retromer function causes Wls to be degraded in the lysosomes and results in reduction of Wls and thus Wnt secretion. These studies led to an emerging picture of Wnt biogenesis (Figure 2). Wnt is glycosylated and lipid modified by Porc in the ER, and is escorted by Wls from the Golgi to the plasma membrane for secretion. Wls is recycled by endocytosis and trafficked back to Golgi by the retromer. Note that porc, wls and retromer mutants largely phenocopywg/wnt mutants in flies and worms, attesting their dedicated roles in Wnt biogenesis.
Wnt proteins can function as morphogens that are capable of both short and long range signaling, as best demonstrated for Wg. Wg lipidation raises the issue of its diffusion and distribution through the aqueous extracellular space. Indeed purified Wnt3a exhibits increased activity via artificial liposomal packaging (Morrell et al., 2008). Two distinct Wg secretory pathways for short and long range signaling have been speculated but not fully substantiated. Wg may form multimers to bury lipid modifications inside (Katanaev et al., 2008), or bind to lipoprotein particles, which may be involved in Wg long range signaling (Panakova et al., 2005) (Figure 2). The membrane microdomain protein reggie-1/flotillin-2 specifically promotes Wg long-range secretion (Katanaev et al., 2008). The Wg receptors (see below) and heparan sulfate proteoglycans (HSPGs) such as Dally and Dally-like protein have important roles in the Wg morphogen concentration via regulating Wg degradation, diffusion, endocytosis/transcytosis, and may function in Wg signaling as potential low-affinity co-receptors (Lin, 2004). Note that reggie-1/flotillin-2, lipoprotein particles, Dally and Dally-like protein are important analogously for secreted Hedgehog morphogen, which is also lipid modified (Katanaev et al., 2008; Lin, 2004; Panakova et al., 2005).
Wnt receptors: Frizzled and LRP5/6
Two distinct receptor families are critical for Wnt/β-catenin signaling (Figure 3): the Frizzled (Fz or Fzd) seven-pass transmembrane receptors (Logan and Nusse, 2004) and the LDL receptor-related proteins 5 and 6 (LRP5 and LRP6) (He et al., 2004). The Wnt-receptor relationship is best illustrated for Wg, which binds toDrosophila Fz2 (Dfz2) and Dfz1 with high affinity (1-10 nM) and requires either Fz in a redundant manner (Logan and Nusse, 2004). Wg reception also absolutely depends on Arrow, the LRP5/6 homolog (He et al., 2004). The mammalian genome harbors 10 Fz genes, most of which have variable capacities to activate β-catenin signaling when co-overexpressed with Wnt and LRP5/6 (e.g., Binnerts et al., 2007) and functional redundancy among Fz members is likely prevalent (Logan and Nusse, 2004). Between the two LRPs, LRP6 plays a more dominant role and is essential for embryogenesis whereas LRP5 is dispensable for embryogenesis but critical for adult bone homeostasis. Nonetheless LRP5 and LRP6 are partially redundant as their functions together are required for mouse gastrulation (He et al., 2004). Most data, including Wnt binding to LRP5/6 and Wnt1-Fz8-LRP6 complex formation in vitro and observations that engineered Fz-LRP5/6 proximity is sufficient to activate β-catenin signaling (Cong et al., 2004; Holmen et al., 2005;Tolwinski et al., 2003), support the model that Wnt induces the formation of Fz-LRP5/6 complex (He et al., 2004) (Figure 1). But unambiguous demonstration of this receptor complex in vivo is lacking. It is noteworthy that Wnt3a palmitoylation (at cysteine 77) is important for binding to both Fz and LRP6 (Cong et al., 2004; Komekado et al., 2007), explaining in part the importance of this lipid modification
Secreted Wnt antagonists and agonists nihms196288f3
A particular Wnt may activate β-catenin and/or non-canonical pathways depending on the receptor complement (van Amerongen et al., 2008). Fz function is involved in β-catenin and non-canonical pathways. The Fz-LRP5/6 co-receptor model stipulates that a Wnt-Fz pair capable of recruiting LRP5/6 activates the β-catenin pathway, consistent with the specific requirement of LRP5/6 in Wnt/β-catenin signaling (He et al., 2004). However some evidence suggests that LRP6 antagonizes non-canonical Wnt signaling in vivo, possibly via competing for Wnt ligands (Bryja et al., 2009) or an unknown mechanism (Tahinci et al., 2007). Other Wnt receptors exist such as Ryk and ROR2, which are not required for, but in some cases may antagonize, Wnt/β-catenin signaling (van Amerongen et al., 2008).
Wnt antagonists and agonists
Several secreted protein families antagonize or modulate Wnt/β-catenin signaling (Figure 3). sFRPs (secreted Frizzled related proteins), and WIF (Wnt inhibitory protein) bind to Wnt, and in the case of sFRPs, also to Fz (Figure 3), and thereby function as Wnt antagonists for both β-catenin and non-canonical signaling (Bovolenta et al., 2008). Loss-of-function studies in mice have revealed significant redundancy for the sFRP genes (Satoh et al., 2008). The Wnt-binding property suggests that sFRPs and WIF may also regulate Wnt stability and diffusion/distribution extracellularly beyond just Wnt inhibitors. Some sFRPs have been shown to have Wnt-independent activity such as regulators of extracellular proteinases (Bovolenta et al., 2008).
Two distinct classes of Wnt inhibitors are the Dickkopf (Dkk) family and the Wise/SOST family (Figure 3). Dkk proteins, exemplified by Dkk1, are LRP5/6 ligands/antagonists and are considered specific inhibitors for Wnt/β-catenin signaling. Although two different models for Dkk1 action have been proposed (Mao et al., 2002; Semenov et al., 2001), recent biochemical and genetic studies (Ellwanger et al., 2008; Semenov et al., 2008; Wang et al., 2008) have argued against the model that Dkk1 inhibits Wnt signaling via inducing LRP6 internalization/degradation through transmembrane Kremen (Krm) proteins (Mao et al., 2002). Dkk1 disruption of Wnt-induced Fz-LRP6 complex remains a more likely mechanism (Semenov et al., 2001), with Krm playing a minor modulatory role in specific tissues (Ellwanger et al., 2008). Wise and SOST constitute another family of LRP5/6 ligands/antagonists (Itasaki et al., 2003; Li et al., 2005; Semenov et al., 2005). Like Dkk1, SOST is able to disrupt Wnt-induced Fz-LRP6 complex in vitro (Semenov et al., 2005). Both Dkk1 and SOST are strongly implicated in human diseases (see below).
Shisa proteins represent a distinct family of Wnt antagonists (Figure 3), which trap Fz proteins in the ER and prevent Fz from reaching the cell surface, thereby inhibiting Wnt signaling cell-autonomously (Yamamoto et al., 2005). Shisa proteins also antagonize FGF (fibroblast growth factor) signaling by trapping FGF receptors in the ER. Other Wnt antagonists with multivalent activities exist. Xenopus Cerberus binds to and inhibits Wnt as well as Nodal and BMP (bone morphogenetic protein) (Piccolo et al., 1999), and IGFBP-4 (Insulin-like growth-factor-binding protein-4) antagonizes Wnt signaling via binding to both Fz and LRP6, in addition to modulating IGF signaling (Zhu et al., 2008).
Norrin and R-spondin (Rspo) proteins are two families of agonists for Wnt/β-catenin signaling (Figure 3). Norrin is a specific ligand for Fz4 and acts through Fz4 and LRP5/6 during retinal vascularization (Xu et al., 2004). Rspo proteins exhibit synergy with Wnt, Fz and LRP6 (Kazanskaya et al., 2004; Kim et al., 2005;Nam et al., 2006; Wei et al., 2007), and show genetic interaction with LRP6 during embryogenesis (Bell et al., 2008), but their mechanism of action is controversial. Results that Rspo binds to both Fz and LRP6 (Nam et al., 2006), to LRP6 primarily (Wei et al., 2007), or to neither (Kazanskaya et al., 2004) have been reported. Another model suggests that Rspo is a ligand for Krm and antagonizes Dkk/Krm-mediated LRP6 internalization (Binnerts et al., 2007), but this seems unlikely given that Krm1 and Krm2 double knockout mice are viable and do not exhibit Rspo mutant phenotypes, and Rspo activates β-catenin signaling in cells lacking both Krm genes (Bell et al., 2008; Ellwanger et al., 2008). Rspo genes are often co-expressed with and depend on Wnt for expression (Kazanskaya et al., 2004), and may represent a means of positive feedback that reinforces Wnt signaling. Mutations in Norrin and Rspo genes cause distinct hereditary diseases (see below).
Wnt signaling
Wnt-off state: β-catenin phosphorylation/degradation by the Axin complex
Cytosolic β-catenin phosphorylation/degradation and its regulation by Wnt are the essence of Wnt signaling (Figure 1). The scaffolding protein Axin uses separate domains to interact with GSK3, CK1α, and β-catenin and coordinates sequential phosphorylation of β-catenin at serine 45 by CK1α and then at threonine 41, serine 37 and serine 33 by GSK3 (Kimelman and Xu, 2006). β-catenin phosphorylation at serine 33 and 37 creates a binding site for the E3 ubiquitin ligase β-Trcp, leading to β-catenin ubiquitination and degradation (Figure 4). Mutations of β-catenin at and surrounding these serine and threonine residues are frequently found in cancers, generating mutant β-catenin that escapes phosphorylation and degradation (Table 1). Axin also contains an RGS (regulator of G protein signaling) domain that interacts with APC, a large multifunctional scaffolding protein that itself binds β-catenin. These core Axin complex components (Kimelman and Xu, 2006) share a common goal of ensuring β-catenin phosphorylation and degradation. Indeed both APC and Axin are tumor suppressor genes, and APC mutations are particularly prevalent in colorectal cancer (Table 1).
Figure 4Regulation of Axin complex assembly for β-catenin degradation
Table 1 Human diseases associated with mutations of the Wnt signaling components
Several aspects of the Axin complex deserve further discussion. (i) In addition to β-catenin, GSK3 and CK1 also phosphorylate Axin and APC, leading to increased association of Axin and APC with β-catenin and thus enhanced β-catenin phosphorylation/degradation (Huang and He, 2008; Kimelman and Xu, 2006) (Figure 4). (ii) Two abundant serine/threonine phosphatases, PP1 and PP2A, both of which associate with Axin and/or APC, counteract the action of GSK3 and/or CK1 in the Axin complex. Thus PP1 dephosphorylates Axin and promotes the disassembly of the Axin complex (Luo et al., 2007), whereas PP2A dephosphorylates β-catenin (Su et al., 2008), each resulting in reduced β-catenin degradation (Figure 4). One should note that PP2A may have multiple and opposing roles in the Wnt pathway depending on the particular associated regulatory subunits and substrates (Kimelman and Xu, 2006). (iii) The assembly of the Axin complex appears to be multivalent and robust. In fly embryos that are null for Axin, expression, at physiological levels, of Axin mutants lacking either the APC-, GSK3-, or β-catenin-binding domain restores a significant degree of normal patterning, implying a quasi-functional Axin complex assembly via multivalent interactions; furthermore, some of these Axin deletion mutants can complement each other and restore fly viability, possibly via Axin dimerization or multimerization (Peterson-Nedry et al., 2008). Indeed Axin has multiple potential dimerization domains (Luo et al., 2005) and the Axin DIX domain may form multimeric polymers (Schwarz-Romond et al., 2007a). (iv) Axin concentration is exceedingly low compared to other components in Xenopus oocytes, indicating that Axin is rate limiting for the complex assembly. This feature may ensure that changes in the Axin protein level will not fluctuate the availability of GSK3 (or other components) for non-Wnt functions, thereby further insulating Wnt and other signaling events (Lee et al., 2003). It is unknown, however, whether the drastic difference between the concentration of Axin versus the other components applies universally, and whether different cells employ quantitative differences in the ratio of Axin and other components to shape their unique Wnt response kinetics (such as the speed and level of β-catenin accumulation). Indeed in Drosophila photoreceptors, APC appears to be present at minimal levels such that a 50% reduction alters the graded Wg response (Benchabane et al., 2008).
Other proteins such as WTX (Wilms tumor gene on the X chromosome) may have roles in β-catenin degradation. Loss of WTX and activating β-catenin mutations seem to have non-overlapping occurrence in Wilms tumor (a pediatric kidney cancer) (Rivera et al., 2007). WTX binds to β-catenin, Axin, APC and β-Trcp to promote β-catenin ubiquitination, although its biochemical role remains unknown (Major et al., 2007). Another Axin-binding protein Diversin can facilitate β-catenin degradation via recruiting CK1ε to phosphorylate β-catenin (Schwarz-Romond et al., 2002).
APC function and APC-Axin cross regulation
The biochemical nature of APC has been enigmatic. A recent study suggested that APC protectsβ-catenin from dephosphorylation by PP2A thereby enhancing β-catenin phosphorylation/degradation (Su et al., 2008) (Figure 4), consistent with the observation that Axin overexpression causes β-catenin degradation even in cells lacking APC function (Behrens et al., 1998). Surprisingly APC (upon phosphorylation by CK1/GSK3) and Axin bind to and compete for the same β-catenin interaction interface, leading to a proposal that APC acts as a “ratchet” to remove phosphorylated β-catenin from Axin for ubiquitination and for making Axin available for a further round of β-catenin phosphorylation (Kimelman and Xu, 2006; Xing et al., 2003). A different model was proposed based on differential β-catenin binding affinity by unphosphorylated versus phosphorylated APC (Ha et al., 2004). APC has also been shown to promote β-catenin nuclear export and to act as a chromatin-associated suppressor for β-catenin target genes, thus functioning in the nucleus (see below).
Another paradoxical observation is that APC has a positive function in physiological and ectopic Wg/Wnt signaling through the promotion of Axin degradation (Lee et al., 2003; Takacs et al., 2008) (Figure 4). One model suggests that this represents a fail-safe mechanism to buffer dramatic β-catenin fluctuations when APC levels vary (Lee et al., 2003). Thus a decrease in the APC level results in higher Axin amounts, compensating for β-catenin degradation. APC-mediated Axin degradation depends on the APC amino terminal domain that is not involved inβ-catenin degradation (Takacs et al., 2008). It is intriguing that colon cancer cells are rarely null for APC but rather retain the amino terminal half, and may have hijacked a part of this fail-safe regulation for tumorigenesis. Conversely Axin can also facilitate APC degradation upon overexpression (Choi et al., 2004), constituting perhaps the other side of the Axin-APC regulation circuit (Figure 4). Mechanisms for Axin and APC degradation, which are proteosome-dependent, have not been characterized.
Wnt-on state
Activation of Wnt receptors
Wnt signaling requires both Fz and LRP6 (or LRP5), likely through a Wnt-induced Fz-LRP6 complex (Figure 1). Wnt-induced LRP6 phosphorylation is a key event in receptor activation (Tamai et al., 2004). LRP6, LRP5 and Arrow each have five reiterated PPPSPxS motifs (P, proline; S, serine or threonine, x, a variable residue), which are essential for LRP6 function and are each transferrable to a heterologous receptor to result in constitutive β-catenin signaling (MacDonald et al., 2008; Tamai et al., 2004; Zeng et al., 2005). These dually phosphorylated PPPSPxS motifs are docking sites for the Axin complex (Davidson et al., 2005;Tamai et al., 2004; Zeng et al., 2005), thereby recruiting Axin to LRP6 upon Wnt stimulation (Mao et al., 2001) (Figure 5).
The kinases responsible for PPPSPxS phosphorylation have been identified unexpectedly as GSK3 and CK1 (Davidson et al., 2005; Zeng et al., 2005). Although one study argued that only CK1 phosphorylation is Wnt-induced (Davidson et al., 2005), most available data support that Wnt induces PPPSP phosphorylation (Binnerts et al., 2007; Khan et al., 2007; Pan et al., 2008b; Wei et al., 2007), which is carried out by GSK3 and primes xS phosphorylation by CK1, thereby leading to dually induced phosphorylation (Zeng et al., 2005) (Figure 5). Although potential involvement of additional kinases cannot be ruled out, experiments in GSK3α/β null cells indicate that GSK3 accounts for most, if not all, PPPSP phosphorylation (Zeng et al., 2008; Zeng et al., 2005). As in β-catenin phosphorylation, Axin-bound GSK3 appears to mediate LRP6 phosphorylation (Zeng et al., 2008). Thus PPPSPxS phosphorylation exhibits a mirror image of β-catenin phosphorylation in sequential order, in priming requirement, and importantly in functionality, but apparently by the same Axin-GSK3 complex (Huang and He, 2008) (Figure 5). This unusual mechanism, using the same kinase complex for both positive and negative regulation, is reminiscent of another morphogenetic pathway, Hedgehog signaling in Drosophila (Price, 2006), and implies a simple vi
Aurelian