[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (23)

Search Parameters:
Keywords = ABT-199/venetoclax

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 1580 KiB  
Article
Bcl-2 Up-Regulation Mediates Taxane Resistance Downstream of APC Loss
by Angelique R. Wise, Sara Maloney, Adam Hering, Sarah Zabala, Grace E. Richmond, Monica K. VanKlompenberg, Murlidharan T. Nair and Jenifer R. Prosperi
Int. J. Mol. Sci. 2024, 25(12), 6745; https://doi.org/10.3390/ijms25126745 - 19 Jun 2024
Viewed by 920
Abstract
Triple-negative breast cancer (TNBC) patients are treated with traditional chemotherapy, such as the taxane class of drugs. One such drug, paclitaxel (PTX), can be effective in treating TNBC; however, many tumors will develop drug resistance, which can lead to recurrence. In order to [...] Read more.
Triple-negative breast cancer (TNBC) patients are treated with traditional chemotherapy, such as the taxane class of drugs. One such drug, paclitaxel (PTX), can be effective in treating TNBC; however, many tumors will develop drug resistance, which can lead to recurrence. In order to improve patient outcomes and survival, there lies a critical need to understand the mechanism behind drug resistance. Our lab made the novel observation that decreased expression of the Adenomatous Polyposis Coli (APC) tumor suppressor using shRNA caused PTX resistance in the human TNBC cell line MDA-MB-157. In cells lacking APC, induction of apoptosis by PTX was decreased, which was measured through cleaved caspase 3 and annexin/PI staining. The current study demonstrates that CRISPR-mediated APC knockout in two other TNBC lines, MDA-MB-231 and SUM159, leads to PTX resistance. In addition, the cellular consequences and molecular mechanisms behind APC-mediated PTX response have been investigated through analysis of the BCL-2 family of proteins. We found a significant increase in the tumor-initiating cell population and increased expression of the pro-survival family member Bcl-2, which is widely known for its oncogenic behavior. ABT-199 (Venetoclax), is a BH3 mimetic that specifically targets Bcl-2. ABT-199 has been used as a single or combination therapy in multiple hematologic malignancies and has shown promise in multiple subtypes of breast cancer. To address the hypothesis that APC-induced Bcl-2 increase is responsible for PTX resistance, we combined treatment of PTX and ABT-199. This combination treatment of CRISPR-mediated APC knockout MDA-MB-231 cells resulted in alterations in apoptosis, suggesting that Bcl-2 inhibition restores PTX sensitivity in APC knockout breast cancer cells. Our studies are the first to show that Bcl-2 functional inhibition restores PTX sensitivity in APC mutant breast cancer cells. These studies are critical to advance better treatment regimens in patients with TNBC. Full article
(This article belongs to the Special Issue Molecular Research in Breast Cancer: Pathophysiology and Treatment)
Show Figures

Figure 1

Figure 1
<p>APC status in TNBC cell lines impacts therapeutic response. CRISPR/Cas9 knockout of APC in (<b>A</b>) MDA-MB-231 or (<b>B</b>) SUM159 cells. Protein from non-targeting control (NTC) and two gRNA-mediated clonal cell lines was run on an SDS-PAGE gel and probed for APC and actin. Blots are representative (n = 3). (<b>C</b>,<b>D</b>) Treatment with PTX induces apoptosis (annexin V/PI staining) in NTC cells. However, clonal APC knockout cells from both MDA-MB-231 (<b>C</b>) and SUM159 (<b>D</b>) show no induction of apoptosis. The graphs show relative PTX-induced apoptosis, compared to DMSO-treated cells. Experiments were performed 3 independent times, and a one-way ANOVA was used to determine significance (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>APC loss alters in vitro tumorigenic phenotypes. (<b>A</b>) Cell counting assay of MDA-MB-231 NTC and APC<sup>KO</sup> cells. Cells were plated and counted from day 3 through day 7 and showed no difference in growth. (<b>B</b>) Wound healing assay showed no change in the ability of cells to fill a scratch over 48 h. Representative images show the original scratch (0 h) and the filled wound (48 h). Images were taken with an EVOS inverted microscope with a 20× objective and Sony ICX285AL CCD camera (<b>C</b>) Clonogenic assay demonstrated increased individual colony area in the 231 clone 1 and clone 2 compared to the 231 NTC cells. Representative images taken with a fixed height camera and a light box show the stained colonies after 8 days in culture. (<b>D</b>) Overall change in fluorescence between control and test samples in an Aldefluor assay showed increased ALDH activity in the 231 clone 1 and clone 2 compared to the 231 NTC cells. Experiments were performed 3 independent times, and a one-way ANOVA was used to determine significance (* <span class="html-italic">p</span> &lt; 0.05; **** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Expression changes in proteins involved in cell cycle and apoptosis. (<b>A</b>) CDK1 expression is increased by western blot in the 231 clone 1 cells but not the 231 clone 2 cells compared to the 231 NTC cells. (<b>B</b>) Bcl-2 expression is increased in both MDA-MB-231 APC knockout clones compared to control. Experiments were performed 3 independent times, and a one-way ANOVA was used to determine significance (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Reversal of Resistance with ABT-199 Combination Treatment. The IC50 for the Bcl-2 specific BH3-mimetic (ABT-199) in MDA-MB-231 cells was determined (2.25 uM). This concentration was used alone or in combination with PTX treatment. APC control (231 NTC) or knockout cells (231 clone 1) were treated with ABT-199 and/or PTX for 24 h, and apoptosis was measured through annexin V/PI staining. After flow cytometric analysis, we observed that while the clones are resistant to PTX-induced apoptosis, the combination treatment induced a robust apoptotic response (n = 3) * <span class="html-italic">p</span> &lt; 0.05 with one-way ANOVA.</p>
Full article ">
24 pages, 5120 KiB  
Article
Identifying Targetable Vulnerabilities to Circumvent or Overcome Venetoclax Resistance in Diffuse Large B-Cell Lymphoma
by Clare M. Adams, Amanda McBride, Peter Michener, Irina Shkundina, Ramkrishna Mitra, Hyun Hwan An, Pierluigi Porcu and Christine M. Eischen
Cancers 2024, 16(11), 2130; https://doi.org/10.3390/cancers16112130 - 3 Jun 2024
Viewed by 881
Abstract
Clinical trials with single-agent venetoclax/ABT-199 (anti-apoptotic BCL2 inhibitor) revealed that diffuse large B-cell lymphoma (DLBCL) is not solely dependent on BCL2 for survival. Gaining insight into pathways/proteins that increase venetoclax sensitivity or unique vulnerabilities in venetoclax-resistant DLBCL would provide new potential treatment avenues. [...] Read more.
Clinical trials with single-agent venetoclax/ABT-199 (anti-apoptotic BCL2 inhibitor) revealed that diffuse large B-cell lymphoma (DLBCL) is not solely dependent on BCL2 for survival. Gaining insight into pathways/proteins that increase venetoclax sensitivity or unique vulnerabilities in venetoclax-resistant DLBCL would provide new potential treatment avenues. Therefore, we generated acquired venetoclax-resistant DLBCL cells and evaluated these together with intrinsically venetoclax-resistant and -sensitive DLBCL lines. We identified resistance mechanisms, including alterations in BCL2 family members that differed between intrinsic and acquired venetoclax resistance and increased dependencies on specific pathways. Although combination treatments with BCL2 family member inhibitors may overcome venetoclax resistance, RNA-sequencing and drug/compound screens revealed that venetoclax-resistant DLBCL cells, including those with TP53 mutation, had a preferential dependency on oxidative phosphorylation. Mitochondrial electron transport chain complex I inhibition induced venetoclax-resistant, but not venetoclax-sensitive, DLBCL cell death. Inhibition of IDH2 (mitochondrial redox regulator) synergistically overcame venetoclax resistance. Additionally, both acquired and intrinsic venetoclax-resistant DLBCL cells were similarly sensitive to inhibitors of transcription, B-cell receptor signaling, and class I histone deacetylases. These approaches were also effective in DLBCL, follicular, and marginal zone lymphoma patient samples. Our results reveal there are multiple ways to circumvent or overcome the diverse venetoclax resistance mechanisms in DLBCL and other B-cell lymphomas and identify critical targetable pathways for future clinical investigations. Full article
(This article belongs to the Section Cancer Therapy)
Show Figures

Figure 1

Figure 1
<p>Generation and characterization of DLBCL cell lines with acquired venetoclax resistance. (<b>A</b>) Venetoclax dose–response curves (MTS, 48 h, relative to DMSO vehicle control, quadruplicates, mean ± SEM) for DLBCL cell lines with intrinsic venetoclax sensitivity (blue shades) or resistance (red shades). IC50 values in parentheses. (<b>B</b>) Western blots of the indicated proteins from the 10 DLBCL lines in A. Each β-actin blot is associated with the blots above it. Two different BCL2 antibodies were needed to detect BCL2 in SUDHL4 and SUDHL6. (<b>C</b>) Venetoclax dose–response curves of parental (Par, blue) and acquired venetoclax-resistant (Res, red) DLBCL lines (MTS, 48 h, relative to DMSO vehicle control, quadruplicates, mean ± SEM). IC50 values of each indicated. (<b>D</b>) Caspase-3/7 activity after 12 h of venetoclax (Ven) treatment or DMSO vehicle control of the lines from C (triplicates/quadruplicates, mean ± SD). Low and high doses of venetoclax are the IC50 and 10× the IC50 of the parental line, respectively. Representative histograms following high-dose venetoclax treatment or DMSO shown. * <span class="html-italic">p</span> &lt; 2.30 × 10<sup>−5</sup>, compared to vehicle (DMSO) control. (<b>E</b>) Chromatograms of <span class="html-italic">BCL2</span> sequencing of the SUDHL4 parental and acquired venetoclax-resistant lines. (<b>F</b>) Western blots of BCL2 family members of the three parental (P) and acquired venetoclax-resistant (R) DLBCL lines (BCL2 Ab1 used for SU4 and SU16; BCL2 Ab2 used for SU6); note that some exposures in (<b>F</b>) were longer than in (<b>B</b>) to detect lower-expressed proteins. Each β-actin blot is associated with at least one of the blots above it.</p>
Full article ">Figure 2
<p>Combination treatment with BCL2 family member inhibitors is effective for DLBCL, follicular, and marginal zone lymphomas. (<b>A</b>) IC50s (μM) of acquired venetoclax-resistant (Res) and parental (Par) DLBCL lines and intrinsically venetoclax-resistant DLBCL lines following 48 h of treatment with navitoclax (Nav, B2XWi), A-1331852 (A852, BCLXi), or MIK665 (MIK, MCL1i). (<b>B</b>) ZIP synergy scores of venetoclax-sensitive and acquired venetoclax-resistant DLBCL lines (top) or intrinsically venetoclax-resistant (bottom) treated with venetoclax (Ven, BCL2i) + A852 (BCLXi) or MIK (MCL1i). See <a href="#app1-cancers-16-02130" class="html-app">Supplementary Table S4</a> for synergy scores from other synergy methods. (<b>C</b>–<b>F</b>) Intracellular flow cytometry of four anti-apoptotic BCL2 family members and treatment results with venetoclax (Ven, V), navitoclax (Nav, N), A-1331852 (A852, A), MIK665 (MIK, M), or untreated (UT) in fresh patient samples of DLBCL (<b>C</b>), normal B-cells (<b>D</b>), follicular lymphoma (<b>E</b>), and marginal zone lymphoma (<b>F</b>). Representative histograms shown with median fluorescence intensity (MFI) after subtracting the isotype control MFI value. Following B-cell enrichment, cell survival (MTS, relative to DMSO vehicle control, triplicates, mean ± SD) was measured 6–12 h after treatment with the compounds indicated. For (<b>C</b>) * <span class="html-italic">p</span> &lt; 0.01, (<b>E</b>) * <span class="html-italic">p</span> &lt; 0.05 (top) and * <span class="html-italic">p</span> &lt; 0.01 (bottom), and (<b>F</b>) * <span class="html-italic">p</span> &lt; 0.05, comparing each concentration used in the combination treatment to the same concentration of each single agent.</p>
Full article ">Figure 3
<p>Targetable oxidative phosphorylation vulnerability identified in venetoclax-resistant DLBCL lines and lymphoma patient samples. (<b>A</b>) Schematic of the workflow of RNA-seq analysis and drug/compound screens (images modified from BioRender.com) with the cell line comparisons indicated and the number of genesets identified (Venn diagram). The six overlapping Hallmark genesets are listed. (<b>B</b>,<b>C</b>) Cell death caused by ETC inhibitor or tigecycline-treated acquired venetoclax-resistant and parental DLBCL lines (<b>B</b>) or intrinsically venetoclax-resistant DLBCL lines (<b>C</b>) was measured with live/dead flow cytometry assay (72–96 h, triplicates, relative to DMSO vehicle control, mean ± SD). For (<b>B</b>) * <span class="html-italic">p</span> &lt; 0.01 (SUDHL6) and * <span class="html-italic">p</span> &lt; 0.05 (SUDHL16), comparing each resistant cell line to its parental counterpart at each concentration. For (<b>C</b>) * <span class="html-italic">p</span> &lt; 0.0001, comparing treated cells at each concentration to untreated cells. (<b>D</b>–<b>F</b>) Treatment of DLBCL (<b>D</b>,<b>E</b>) and marginal zone (<b>F</b>) lymphoma patient samples with venetoclax (Ven, V), mubritinib (Mub, M), BAY-87-2243 (BAY, B), and/or IACS-010759 (IACS, IA). Cell viability assays ((<b>D</b>,<b>F</b>), 24 h, triplicates, relative to DMSO vehicle control, mean ± SD) and Caspase-3/7 activity assay ((<b>E</b>), 24 h, triplicates, fold-change relative to DMSO vehicle control, mean ± SD). (<b>D</b>–<b>F</b>), * <span class="html-italic">p</span> &lt; 0.0001, comparing each combination treatment to both single agent treatments at the same concentrations.</p>
Full article ">Figure 4
<p>IDH2 is upregulated in acquired venetoclax-resistant DLBCL cells, and its inhibition synergizes with venetoclax to overcome resistance. (<b>A</b>) Number of overlapping and non-overlapping genes identified from Hallmark oxidative phosphorylation geneset evaluation after the comparisons shown in <a href="#cancers-16-02130-f003" class="html-fig">Figure 3</a>A were performed. Venn diagram (left) and significantly (FDR &lt; 0.05 with ≥1.5-fold change) altered genes are in the heatmaps (right). (<b>B</b>–<b>D</b>) Combination treatment with venetoclax (Ven, V) + IDH2 inhibitor (AGI-6780, AGI, A) in acquired venetoclax-resistant DLBCL lines. Survival assays ((<b>B</b>), left, 48 h, quadruplicates, relative to DMSO vehicle control, mean ± SEM) and 3D ZIP synergy plots ((<b>B</b>), right), live/dead flow cytometry analysis ((<b>C</b>), left, 48 h, triplicates, relative to DMSO vehicle control, mean ± SEM) and 3D ZIP synergy plots ((<b>C</b>), right), and Caspase-3/7 activity measured ((<b>D</b>), 48 h, triplicates, mean ± SD). Untreated (UT). (<b>E</b>–<b>G</b>) Treatment of DLBCL (<b>E</b>,<b>F</b>) and marginal zone lymphoma (<b>G</b>) patient samples with venetoclax (Ven, V) and AGI-6780 (AGI, A). Cell viability assays for IDH2i ((<b>E</b>,<b>G</b>), 24 h, triplicates, relative to DMSO vehicle control, mean ± SD) and Caspase-3/7 activity ((<b>F</b>), 24 h, triplicates, fold-change relative to DMSO vehicle control, mean ± SD). For (<b>D</b>–<b>G</b>) * <span class="html-italic">p</span> &lt; 0.0001, comparing each combination treatment to both single-agent treatments at the same concentrations.</p>
Full article ">Figure 5
<p>Targetable critical pathways in venetoclax-sensitive and -resistant DLBCL revealed from drug/compound screening. (<b>A</b>) Median fold-change of all compounds in the categories indicated with ≥1.5-fold increased sensitivity in acquired venetoclax-resistant DLBCL lines compared to venetoclax-sensitive lines identified by drug screens. (<b>B</b>,<b>C</b>) IC50s (μM) of acquired venetoclax-resistant (Res) and parental (Par) DLBCL lines and intrinsically venetoclax-resistant DLBCL lines following 48 h of treatment with CDK7/9 (<b>B</b>) or BCR (<b>C</b>) inhibitors. Transcription inhibitors: CDK9-IN-2 (CDK9i), SNS-032 (CDK2/7/9i), THZ1 (CDK7i). B-cell receptor signaling inhibitors: copanlisib (Copa, PI3Ki), ibrutinib (Ibru, BTKi), and R406 (SYKi). (<b>D</b>,<b>E</b>) MTS assays of enriched B-cells from fresh DLBCL (<b>D</b>,<b>E</b>), follicular (<b>D</b>,<b>E</b>), and marginal zone (<b>D</b>) lymphoma patient samples treated 24 h with venetoclax (Ven, V), THZ1 (T), CDK9 (CD), ibrutinib (Ib), R406 (R), and/or copanlisib (Copa, Co) (triplicates, relative to DMSO vehicle control, mean ± SD). For (<b>D</b>) * <span class="html-italic">p</span> &lt; 0.05 (top), * <span class="html-italic">p</span> &lt; 0.01 (middle), and * <span class="html-italic">p</span> &lt; 0.0001 (bottom), and for (<b>E</b>) * <span class="html-italic">p</span> &lt; 0.0001, comparing each combination treatment to both single-agent treatments at the same concentrations.</p>
Full article ">
13 pages, 3986 KiB  
Article
PROTAC-Mediated Dual Degradation of BCL-xL and BCL-2 Is a Highly Effective Therapeutic Strategy in Small-Cell Lung Cancer
by Sajid Khan, Lin Cao, Janet Wiegand, Peiyi Zhang, Maria Zajac-Kaye, Frederic J. Kaye, Guangrong Zheng and Daohong Zhou
Cells 2024, 13(6), 528; https://doi.org/10.3390/cells13060528 - 17 Mar 2024
Cited by 4 | Viewed by 2220
Abstract
BCL-xL and BCL-2 are validated therapeutic targets in small-cell lung cancer (SCLC). Targeting these proteins with navitoclax (formerly ABT263, a dual BCL-xL/2 inhibitor) induces dose-limiting thrombocytopenia through on-target BCL-xL inhibition in platelets. Therefore, platelet toxicity poses a barrier in advancing the clinical translation [...] Read more.
BCL-xL and BCL-2 are validated therapeutic targets in small-cell lung cancer (SCLC). Targeting these proteins with navitoclax (formerly ABT263, a dual BCL-xL/2 inhibitor) induces dose-limiting thrombocytopenia through on-target BCL-xL inhibition in platelets. Therefore, platelet toxicity poses a barrier in advancing the clinical translation of navitoclax. We have developed a strategy to selectively target BCL-xL in tumors, while sparing platelets, by utilizing proteolysis-targeting chimeras (PROTACs) that hijack the cellular ubiquitin proteasome system for target ubiquitination and subsequent degradation. In our previous study, the first-in-class BCL-xL PROTAC, called DT2216, was shown to have synergistic antitumor activities when combined with venetoclax (formerly ABT199, BCL-2-selective inhibitor) in a BCL-xL/2 co-dependent SCLC cell line, NCI-H146 (hereafter referred to as H146), in vitro and in a xenograft model. Guided by these findings, we evaluated our newly developed BCL-xL/2 dual degrader, called 753b, in three BCL-xL/2 co-dependent SCLC cell lines and the H146 xenograft models. 753b was found to degrade both BCL-xL and BCL-2 in these cell lines. Importantly, it was considerably more potent than DT2216, navitoclax, or DT2216 + venetoclax in reducing the viability of BCL-xL/2 co-dependent SCLC cell lines in cell culture. In vivo, 5 mg/kg weekly dosing of 753b was found to lead to significant tumor growth delay, similar to the DT2216 + venetoclax combination in H146 xenografts, by degrading both BCL-xL and BCL-2. Additionally, 753b administration at 5 mg/kg every four days induced tumor regressions. At this dosage, 753b was well tolerated in mice, without observable induction of severe thrombocytopenia as seen with navitoclax, and no evidence of significant changes in mouse body weights. These results suggest that the BCL-xL/2 dual degrader could be an effective and safe therapeutic for a subset of SCLC patients, warranting clinical trials in future. Full article
(This article belongs to the Section Cell Microenvironment)
Show Figures

Figure 1

Figure 1
<p>753b degrades both BCL-xL and BCL-2 leading to apoptosis in SCLC cells. (<b>a</b>) Chemical structures of DT2216 and 753b. Linker in 753b is highlighted in red. (<b>b</b>–<b>d</b>) Immunoblot analyses of BCL-X<sub>L</sub>, BCL-2, BCL-w, MCL-1, caspase 3 (C3), cleaved caspase 3 (CC3), PARP and cleaved (c)-PARP in SCLC H146 (<b>b</b>), H211 (<b>c</b>), and H1059 cells (<b>d</b>) after they were treated with indicated concentrations of 753b for 24 h. The β-tubulin was used as an equal loading control. (<b>e</b>–<b>g</b>) Densitometric analysis of BCL-xL and BCL-2 immunoblots in H146 (<b>e</b>), H211 (<b>f</b>), and H1059 cells (<b>g</b>) showing DC<sub>50</sub> and D<sub>max</sub> values for each protein. DC<sub>50,</sub> concentration required to degrade 50% of the protein; D<sub>max</sub>, maximum degradation in percentage; nd, not determined.</p>
Full article ">Figure 2
<p>753b is more potent than DT2216 or navitoclax to kill BCL-X<sub>L</sub>/2-dependent SCLC cells. (<b>a</b>–<b>e</b>) Viability of H146 (<b>a</b>), H211 (<b>b</b>), H1059 (<b>c</b>), and H378 (<b>d</b>) SCLC cells, and WI38 normal lung fibroblasts (<b>e</b>), after they were treated with increasing concentrations of 753b, DT2216, or navitoclax for 72 h. (<b>f</b>) IC<sub>50</sub> values for 753b, DT2216, and navitoclax in SCLC cell lines and WI38 cells are tabulated.</p>
Full article ">Figure 3
<p>753b showed comparable or enhanced efficacy as compared to DT2216 + venetoclax combination to kill BCL-xL/2-dependent SCLC cells. (<b>a</b>–<b>c</b>) Viability of H146 (<b>a</b>), H211 (<b>b</b>), and H1059 cells (<b>c</b>) after they were treated with increasing concentrations of DT2216, venetoclax, or their 1:1 combination. IC<sub>50</sub> values for the individual agents and combinations are tabulated. (<b>d</b>–<b>f</b>) Viability of H146 (<b>d</b>), H211 (<b>e</b>), and H1059 cells (<b>f</b>) after they were treated with increasing concentrations of 753b, venetoclax, or their 1:1 combination. IC<sub>50</sub> values for the individual agents and combinations are tabulated.</p>
Full article ">Figure 4
<p>753b is more potent than DT2216 and similar to DT2216 + venetoclax for inhibiting growth of BCL-X<sub>L</sub>/2-dependent H146 xenograft tumors in mice. (<b>a</b>) Tumor volume changes in H146 xenografts after treatment with vehicle, DT2216 (15 mg/kg, weekly i.e., q7d, i.p.), a combination of DT2216 with venetoclax (50 mg/kg, 5 days a week, p.o.), or 753b (5 mg/kg, q7d, i.p.). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 7, 6, 6, and 7 mice in vehicle, DT2216, DT2216 + venetoclax, and 753b groups, respectively, at the start of treatment). When the biggest tumor dimension reached 1.5 cm, the mice were sacrificed in accordance with IACUC protocol, and the remaining mice continued to be treated for up to 104 days. Tumor volume changes are shown up to post-treatment day 88, when 5 or more mice were alive in each treatment group. **** <span class="html-italic">p</span> &lt; 0.0001, ns: not significant as determined by one-way ANOVA and Dunnett’s multiple comparisons test at post-treatment day 32. (<b>b</b>) Kaplan-Meier survival analysis of mice as treated in (<b>a</b>). Survival time was recorded at the tumor endpoint, i.e., biggest tumor dimension of 1.5 cm or more. The median survival time is shown on the right. *** <span class="html-italic">p</span> &lt; 0.001, ns: not significant as determined by two-sided Student’s <span class="html-italic">t</span>-test. (<b>c</b>) Mouse body weight changes in H146 xenografts after treatment as in (<b>a</b>). Data are presented as mean ± SEM. (<b>d</b>) Immunoblot analysis of BCL-xL, BCL-2, and MCL-1 in H146 xenograft tumors two days after last treatment with vehicle, DT2216, DT2216 + venetoclax (DT + VEN), or 753b (<span class="html-italic">n</span> = 3 mice per group) as in (<b>a</b>). (<b>e</b>) Densitometric analysis of immunoblots in (<b>d</b>). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared to vehicle as determined by one-way ANOVA and Dunnett’s multiple comparisons test.</p>
Full article ">Figure 5
<p>753b induces regressions of larger H146 xenograft tumors in mice. (<b>a</b>) Tumor volume changes in H146 xenografts after treatment with vehicle or 753b (5 mg/kg, every four days i.e., q4d, i.p.). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 5 mice). ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle as determined by two-sided Student’s <span class="html-italic">t</span>-test. (<b>b</b>), Tumor weights at the end of experiment in (<b>a</b>). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 5 mice). *** <span class="html-italic">p</span> &lt; 0.001 compared to vehicle as determined by two-sided Student’s <span class="html-italic">t</span>-test. (<b>c</b>) The images of excised tumors from (<b>a</b>). (<b>d</b>) Mouse body weight changes in H146 xenografts after treatment as in (<b>a</b>).</p>
Full article ">
17 pages, 2520 KiB  
Article
Simultaneous Inhibition of Mcl-1 and Bcl-2 Induces Synergistic Cell Death in Hepatocellular Carcinoma
by Marlen Michalski, Magdalena Bauer, Franziska Walz, Deniz Tümen, Philipp Heumann, Petra Stöckert, Manuela Gunckel, Claudia Kunst, Arne Kandulski, Stephan Schmid, Martina Müller and Karsten Gülow
Biomedicines 2023, 11(6), 1666; https://doi.org/10.3390/biomedicines11061666 - 8 Jun 2023
Cited by 2 | Viewed by 1874
Abstract
Despite the recent approval of new therapies, the prognosis for patients with hepatocellular carcinoma (HCC) remains poor. There is a clinical need for new highly effective therapeutic options. Here, we present a combined application of BH3-mimetics as a potential new treatment option for [...] Read more.
Despite the recent approval of new therapies, the prognosis for patients with hepatocellular carcinoma (HCC) remains poor. There is a clinical need for new highly effective therapeutic options. Here, we present a combined application of BH3-mimetics as a potential new treatment option for HCC. BH3-mimetics inhibit anti-apoptotic proteins of the BCL-2 family and, thus, trigger the intrinsic apoptosis pathway. Anti-apoptotic BCL-2 proteins such as Bcl-2 and Mcl-1 are frequently overexpressed in HCC. Therefore, we analyzed the efficacy of the two BH3-mimetics ABT-199 (Bcl-2 inhibitor) and MIK665 (Mcl-1 inhibitor) in HCC cell lines with differential expression levels of endogenous Bcl-2 and Mcl-1. While administration of one BH3-mimetic alone did not substantially trigger cell death, the combination of two inhibitors enhanced induction of the intrinsic apoptosis pathway. Both drugs acted synergistically, highlighting the effectivity of this specific BH3-mimetic combination, particularly in HCC cell lines. These results indicate the potential of combining inhibitors of the BCL-2 family as new therapeutic options in HCC. Full article
Show Figures

Figure 1

Figure 1
<p>Determination of p53 and BCL-2 family protein expression in different HCC cell lines. (<b>A</b>–<b>E</b>) Western blot analyses of p53 and BCL-2 family members Bcl-2, Mcl-1, Bcl-X<sub>L</sub> and NOXA in Hep3B, HepG2 and Huh7. The three cell lines displayed a differential expression pattern. Shown is a representative western blot. (<b>F</b>–<b>I</b>) Densitometric analysis of western blots: Shown are the relative amounts of Bcl-2, Mcl-1, Bcl-X<sub>L</sub> and NOXA in Hep3B (red), HepG2 (blue) and Huh7 (green) normalized to β-actin (<span class="html-italic">n</span> = 3, mean ± SD).</p>
Full article ">Figure 2
<p>Combination of ABT-199 and MIK665 results in cytochrome c release in Hep3B cells. (<b>A</b>,<b>B</b>) Corresponding isotype controls. (<b>C</b>) DMSO treatment was used as vehicle control. (<b>D</b>) Cells were treated with 30 µM FCCP for 3 h to induce MOMP as a positive control. (<b>E</b>–<b>J</b>) Immunofluorescence of TOMM20 and cytochrome c in Hep3B cells. Cells were subjected to ABT-199 and MIK665 for 4 h at the indicated concentrations as single or combined treatment (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>Combination of the BH3-mimetics ABT-199 and MIK665 results in increased cleavage and activity of caspase-9 in HCC cell lines. (<b>A</b>–<b>C</b>) Western blot analyses of (cleaved) caspase-9. Shown is a representative western blot. Cells were treated with either one BH3-mimetic or a combination of ABT-199 or MIK665 inhibitors. DMSO was used as a vehicle control. (<b>D</b>,<b>E</b>) Activity assays of caspase-9 in HCC cell lines (Hep3B (red), HepG2 (blue), and Huh7 (green)) after 4 h of treatment with either ABT-199 or MIK665 alone or in combination at the indicated concentrations (the dotted line separates the single treatment from the combination treatment with ABT-199 and MIK665). Data show a specific increase in activity compared to DMSO-treated controls (<span class="html-italic">n</span> = 3, mean ± SEM) (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001 compared to treatment with one inhibitor alone).</p>
Full article ">Figure 4
<p>Combination of ABT-199 and MIK665 results in increased cleavage and activity of caspase-3 in HCC cell lines. (<b>A</b>–<b>C</b>) Western blot analyses of (cleaved) caspase-3. Shown is a representative western blot. Cells were treated with either one of the two BH3-mimetics ABT-199 or MIK665 alone or a combination of both inhibitors. DMSO was used as vehicle control. (<b>D</b>,<b>E</b>) Activity assays of caspase-3/7 in HCC cell lines (Hep3B (red), HepG2 (blue), and Huh7 (green)) after 4 h of treatment with either ABT-199 or MIK665 alone or in combination at the indicated concentrations (the dotted line separates the single treatment from the combination treatment with ABT-199 and MIK665). Data show a specific increase in activity compared to DMSO-treated controls (<span class="html-italic">n</span> = 3, mean ± SEM). (<b>F</b>–<b>H</b>) Western blot analyses of (cleaved) PARP. Shown is a representative western blot. Cells were incubated with ABT-199 and MIK665 at the indicated concentrations as single or combined treatment. DMSO was used as vehicle control. (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001 compared to single drug treatment).</p>
Full article ">Figure 5
<p>Treatment with combinations of ABT-199 and MIK665 induces cell death in Hep3B, HepG2 and Huh7. (<b>A</b>,<b>C</b>,<b>E</b>) Specific cell death was analyzed in HCC cells treated with a constant ABT-199 concentration (5 µM) and increasing MIK665 concentrations as indicated for up to 48 h (<span class="html-italic">n</span> = 3, mean ± SD). (<b>B</b>,<b>D</b>,<b>F</b>) Specific cell death was analyzed in HCC cells treated with a steady MIK665 concentration (6 µM) and increasing ABT-199 concentrations as indicated for up to 48 h (the dotted line separates the single treatment from the combination treatment with ABT-199 and MIK665) (<span class="html-italic">n</span> = 3, mean ± SD) (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001 compared to single drug treatment).</p>
Full article ">Figure 6
<p>Combination treatment with ABT-199 and MIK665 has synergistic effects in HCC cells. (<b>A</b>–<b>F</b>) Graphs show the IC<sub>50</sub> of monotherapy and combinations of ABT-199 and MIK665 administered to the HCC cell lines Hep3B, HepG2 and Huh7 for up to 48 h of treatment. For combination treatments, all cell lines were treated with a steady concentration of 5 µM ABT-199 or 6 µM MIK665 and increasing concentrations of the respective other BH3-mimetic (<span class="html-italic">n</span> = 3, mean ± SD). IC<sub>50</sub> calculation was executed with GraphPad Prism 8 (* <span class="html-italic">p</span> &lt;0.05; ** <span class="html-italic">p</span> &lt;0.01; *** <span class="html-italic">p</span> &lt;0.001; **** <span class="html-italic">p</span> &lt; 0.0001 compared to single drug treatment). (<b>G</b>,<b>H</b>) Graphs show dot plots of the calculated combination indices (CI) of the interpolated IC<sub>50</sub> of the ABT-199/MIK665 combinations at 24 h and 48 h of treatment. CI was calculated using CompuSyn (version 1) (<a href="http://www.combosyn.com/" target="_blank">http://www.combosyn.com/</a> (accessed on 26 May 2023)). CI values &lt; 1 indicate a synergistic effect of the combination treatment. Smaller CI values indicate stronger synergism. A CI &lt; 0.3 (dotted line) signifies strong to very strong synergism [<a href="#B44-biomedicines-11-01666" class="html-bibr">44</a>].</p>
Full article ">
31 pages, 6900 KiB  
Article
Pictolysin-III, a Hemorrhagic Type-III Metalloproteinase Isolated from Bothrops pictus (Serpentes: Viperidae) Venom, Reduces Mitochondrial Respiration and Induces Cytokine Secretion in Epithelial and Stromal Cell Lines
by Dan E. Vivas-Ruiz, Paola Rosas, Alex Proleón, Daniel Torrejón, Fanny Lazo, Ana Belén Tenorio-Ricca, Francisco Guajardo, Cristopher Almarza, Víctor Andrades, Jessica Astorga, Daniel Oropesa, Jorge Toledo, María Jesús Vera, Jorge Martínez, Ramiro Araya-Maturana, Karen Dubois-Camacho, Marcela A. Hermoso, Valéria G. Alvarenga, Eladio Flores Sanchez, Armando Yarlequé, Luciana Souza Oliveira and Félix A. Urraadd Show full author list remove Hide full author list
Pharmaceutics 2023, 15(5), 1533; https://doi.org/10.3390/pharmaceutics15051533 - 18 May 2023
Cited by 4 | Viewed by 2661
Abstract
From the venom of the Bothrops pictus snake, an endemic species from Peru, we recently have described toxins that inhibited platelet aggregation and cancer cell migration. In this work, we characterize a novel P-III class snake venom metalloproteinase, called pictolysin-III (Pic-III). It is [...] Read more.
From the venom of the Bothrops pictus snake, an endemic species from Peru, we recently have described toxins that inhibited platelet aggregation and cancer cell migration. In this work, we characterize a novel P-III class snake venom metalloproteinase, called pictolysin-III (Pic-III). It is a 62 kDa proteinase that hydrolyzes dimethyl casein, azocasein, gelatin, fibrinogen, and fibrin. The cations Mg2+ and Ca2+ enhanced its enzymatic activity, whereas Zn2+ inhibited it. In addition, EDTA and marimastat were also effective inhibitors. The amino acid sequence deduced from cDNA shows a multidomain structure that includes a proprotein, metalloproteinase, disintegrin-like, and cysteine-rich domains. Additionally, Pic-III reduces the convulxin- and thrombin-stimulated platelet aggregation and in vivo, it has hemorrhagic activity (DHM = 0.3 µg). In epithelial cell lines (MDA-MB-231 and Caco-2) and RMF-621 fibroblast, it triggers morphological changes that are accompanied by a decrease in mitochondrial respiration, glycolysis, and ATP levels, and an increase in NAD(P)H, mitochondrial ROS, and cytokine secretion. Moreover, Pic-III sensitizes to the cytotoxic BH3 mimetic drug ABT-199 (Venetoclax) in MDA-MB-231 cells. To our knowledge, Pic-III is the first SVMP reported with action on mitochondrial bioenergetics and may offer novel opportunities for promising lead compounds that inhibit platelet aggregation or ECM–cancer-cell interactions. Full article
(This article belongs to the Section Biologics and Biosimilars)
Show Figures

Figure 1

Figure 1
<p>Pictolysin-III purification from <span class="html-italic">B. pictus</span> venom. Pic-III was purified by a three-step purification procedure as described in the <a href="#sec2-pharmaceutics-15-01533" class="html-sec">Section 2</a>. (<b>A</b>) <span class="html-italic">B. pictus</span> venom (682 mg) was separated on Sephacryl S-200 resin. The fractions containing proteolytic and hemorrhagic activities were pooled for the next step. (<b>B</b>) Thereafter, 155 mg of lyophilized product was applied on a DEAE-Sepharose CL-6B column with a linear salt gradient from 0–0.3 M NaCl. (<b>C</b>) Active fractions (31 mg) were pooled and applied on a CM Sepharose CL-6B column. Active metalloproteinase fractions containing Pic-III (peak 1) were pooled, dialyzed against 1 mM CaCl<sub>2</sub> solution in water, and lyophilized. (<b>D</b>) The SDS-PAGE (12% gel) of purified Pic-III (5 μg) under non-reducing (NR) and reducing (R) conditions. (<b>E</b>) Peaks 1 and 2 of CM Sepharose were analyzed by Western blot using anti–Atr-III as the primary antibody. (<b>F</b>) SDS-PAGE of purified Pic-III under treatment with PNGase (+). (<b>G</b>) Transcript levels of Pic-III, TLE, PLA2, and LAAO from the venom of <span class="html-italic">B. pictus</span>. The levels of gene expression were analyzed by qRT-PCR real-time, using the housekeeping GAPDH. The data are presented as mean values ± SD, N = 3; TLE: thrombin-like enzyme; Pictobin; PLA2: Phospholipase A2; LAAO: L-amino acid oxidase from <span class="html-italic">B. pictus</span> venom. In (<b>G</b>), the data shown are the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05 vs. Pic-III; n.s., not significant.</p>
Full article ">Figure 2
<p>Biochemical characterization of Pictolysin-III. (<b>A</b>) Effect of pH on the proteolytic activity of Pic-III (1 mg/mL) using azocasein as the substrate. The enzyme was pre-incubated with different buffers before the determination of azocaseinolytic activity. (<b>B</b>) Effect of different temperatures on the proteolytic activity of Pic-III (1 mg/mL). (<b>C</b>) Effect of inhibitors and cations on Pic-III activity. (<b>D</b>) Inhibition of the proteolytic activity of Pic-III by PAS (INS-Peru). A total of half, one, and two neutralizing doses were tested. The enzyme (25 μg) was incubated with PAS for 30 min at 37 °C before the activity. The effect was described as % residual activity. (<b>E</b>) Pic-III (1 mg/mL), in the presence or absence of PNGase F (100 U), was incubated with azocasein to assess its proteolytic activity. (<b>F</b>) Effect of PNGase F on Pic-III proteolytic activity at 1 h of incubation. (<b>G</b>) Interaction between Pic-III and Pictobin (Picb) (1:1:1 molar ratio) with the endogenous α2-M inhibitor. Pic-III is not inhibited when it interacts with the serine protease, demonstrating a possible functional synergy. The data shown are the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. control (Pic-III); n.s., not significant. In (<b>A</b>,<b>B</b>), vs. proteolytic activity at pH = 4 and 5 °C, respectively.</p>
Full article ">Figure 3
<p>Biological activities of Pictolysin-III. Digestion of (<b>A</b>) H-Fg and (<b>B</b>) fibrin. Both digestion reactions were conducted using 1 µg of Pic-III and were analyzed with 14% SDS-PAGE. The lane description of gels is C: control; time digestion: 5-, 15-, 30-, and 60-min. Polypeptide chains of H-Fg control (α, β, and γ) and fibrin control (γ-γ dimer, α, and β) are indicated at the right. Pic-III degrades human (<b>C</b>) and bovine (<b>D</b>) fibrinogen in the same way. In both cases, the activity is significantly inhibited by the chelating agent EDTA. Effect of Pic-III on human platelet aggregation stimulated by thrombin (Thr, 1 U/mL) (<b>E</b>), convulxin (CVX, 6 µg/mL) (<b>F</b>), collagen-I (CoI, 10 μg/mL) (<b>G</b>), von Willebrand factor (vWF, 5.5 μg/mL) plus ristocetin (0.5 mg/mL) (<b>H</b>). Platelet aggregation was recorded by aggregometry. (<b>I</b>,<b>J</b>) Hemorrhagic activity of <span class="html-italic">B. pictus</span> venom and Pic-III. Injection of PBS was used as a control. The data shown are the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. CVX or Thr; n.s., not significant.</p>
Full article ">Figure 4
<p>The cDNA and deduced amino acid sequence of Pictolysin-III. The cDNA sequence with 1830 pb encodes an open reading frame for 610 amino acid residues. The deduced amino acid sequence is comprised of signal peptide (1–20), pro-peptide domain (26–153), metalloproteinase domain (190–392), disintegrin-like domain (404–486), and cysteine-rich domain (487–610). The deduced amino acid sequence is represented by one-letter code.</p>
Full article ">Figure 5
<p>Multiple alignment of Pictolysin-III primary structure with SVMP-III homologs from bothropic venoms and human ADAM. The conserved residues are colored red and the cysteine residues are shaded in red. The hypervariable region (HVR), Ca<sup>2+</sup> binding sites, and Zn<sup>2+</sup> binding sites are boxed in blue, red, and green, respectively. The M domain, D domain (D<sub>s</sub> and D<sub>a</sub>), and C domain (C<sub>w</sub> and C<sub>h</sub>) are drawn schematically. Abbreviations: B.atr: <span class="html-italic">Bothrops atrox</span>, C.atr: <span class="html-italic">Crotalus atrox</span>, B.jar: <span class="html-italic">Bothrops jararaca</span>, H.sap: <span class="html-italic">Homo sapiens</span>.</p>
Full article ">Figure 6
<p>Prediction of the structural model and evolutionary relationship for Pictolysin-III. (<b>A</b>) The three-dimensional structure of protein reveals the presence of three domains: M, D, and C. The M domain at the N-terminal (yellow sand, 1–203) has a Zn<sup>2+</sup>-binding site (145-HEMGHNLGIHH-155) and a Ca<sup>2+</sup>-binding site (E12, N203, D96, and C200). A linker sequence lies between the M and D domains (gray, 204–214). The D domain (215–298) could be divided into two subdomains, the “shoulder” (D<sub>s</sub>-domain) (cyan, 215–248) and the “arm” (D<sub>a</sub>-domain) (red, 249–298). The D domain presents two Ca<sup>2+</sup>-binding sites, one in the D<sub>s</sub>-domain (V215, L220, E222, and D228) and other in the D<sub>a</sub>-domain (D279, D294, E282, and V295). Another linker sequence lies between the D and C domains (gray, 299–315). The C domain (green, 316–421) at the C-terminal presents the hypervariable region (blue, 373–394). The binding sites were graphed as sticks. (<b>B</b>) Phylogenetic tree of P-III SVMPs based on nucleotidic sequences. ADAM 7 from <span class="html-italic">H. sapiens</span> was used as an out-group. Bootstrap values are shown at each node. Abbreviations: B.jar: <span class="html-italic">Bothrops jararaca</span>, B.atr: <span class="html-italic">Bothrops atrox</span>, C.atr: <span class="html-italic">Crotalus atrox</span>, D.acu: <span class="html-italic">Deinagkistrodon acutus</span>, B.ery: <span class="html-italic">Bothrops erythromelas</span>, D.rus: <span class="html-italic">Daboia russelii</span>, T.fla: <span class="html-italic">Trimeresurus flavoviridis</span>, E.car: <span class="html-italic">Echis carinatus</span>, N.naj: <span class="html-italic">Naja naja</span>, B.fas: <span class="html-italic">Bungarus fasciatus</span>, A.eng: <span class="html-italic">Atractaspis engaddensis</span>, P. olf: <span class="html-italic">Philodryas olfersii</span>, P. cha: <span class="html-italic">Philodryas chamissonis</span>, H.sap: <span class="html-italic">Homo sapiens</span>.</p>
Full article ">Figure 7
<p>Pictolysin-III modifies Caco-2 cell morphology. (<b>A</b>) Representative images of a time-lapse experiment of Caco-2 cells treated with 50 µg/mL Pic-III during 3 h with DIC microscopy. (<b>B</b>) Representative images of Caco-2 cell treatments were DAPI (yellow), rhodamine phalloidin (cyan), (<b>C</b>) Mask Deep Red (magenta), and (<b>D</b>) DIC (white). (<b>E</b>) Quantification of morphological changes in Caco-2 cells, respectively, treated with 10, 20, and 50 µg/mL Pic-III during 3 h (<span class="html-italic">n</span> = 329; 143; 193; 229 independent cells, respectively). Changes in spreading cell (area), (<b>F</b>) density of F-actin, (<b>G</b>) circularity (values range 1 round), and (<b>H</b>) elongation (normalized ratio of the square of the perimeter by the area). The data shown are mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001 vs. control.</p>
Full article ">Figure 8
<p>Effect of Pic-III on the MTT reduction in epithelial (MDA-MB-231 and Caco-2) and stromal (RMF-621) cells. Cells were treated with Pic-III (10, 20, and 50 μg/mL) for 48 h and cell viability was determined by MTT assay. Data are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001 vs. control; n.s., not significant.</p>
Full article ">Figure 9
<p>Pic-III reduces mitochondrial respiration. (<b>A</b>,<b>C</b>,<b>E</b>) Changes in the profile of respiration of MDA-MB-231, Caco-2, and RMF-621 cells induced by Pic-III. (<b>B</b>,<b>D</b>,<b>F</b>) Changes in the basal and maximal OCR in MDA-MB-231, Caco-2, and RMF-621 cells, respectively, treated with 10, 20, and 50 µg/mL Pic-III. Rot: Rotenone; AA: Antimycin A. Data are expressed as means ± SD. *** <span class="html-italic">p</span> &lt; 0.001 vs. control; n.s., not significant.</p>
Full article ">Figure 10
<p>Changes in glycolysis induced by Pic-III. (<b>A</b>,<b>C</b>,<b>E</b>) Changes in the profile of extracellular acidification rate (ECAR) of MDA-MB-231, Caco-2, and RMF-621 cells induced by Pic-III. (<b>B</b>,<b>D</b>,<b>F</b>) Effect of Pic-III on glycolysis and glycolytic capacity in MDA-MB-231, Caco-2, and RMF-621 cells, respectively. 2-DG: 2-deoxy-D-glucose. Data are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, vs. control; n.s., not significant.</p>
Full article ">Figure 11
<p>Pic-III reduces the intracellular ATP levels and affects the NAD(P)H and mitochondrial ROS. (<b>A</b>–<b>C</b>) Effect of Pic-III on ATP levels. (<b>D</b>–<b>F</b>) Effect of Pic-III (50 µg/mL), antimycin A (5 µM), and FCCP (5 µM) on NAD(P)H levels in MDA-MB-231, Caco-2, and RMF-621 cells. (<b>G</b>) Effect of Pic-III (50 µg/mL) on mitochondrial ROS in RMF-621 cells and (<b>H</b>) Caco-2 cells at 48 h, (<b>I</b>) Effect of the protonophoric agent FCCP (1 µM) on the inhibitory effect of the metabolic capacity of Pic-III (50 µg/mL) in RMF-621 cells. Data are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. control; n.s., not significant.</p>
Full article ">Figure 12
<p>Pic-III promotes the secretion of cytokines. (<b>A</b>,<b>B</b>) Effect of Pic-III on the expression of IL1β and TNFα genes in RMF-621 cells treated for 8 h, (<b>C</b>–<b>F</b>) Effect of Pic-III on cytokine production in RMF-621 and (<b>G</b>–<b>J</b>) Caco-2 cells. Cells were treated with Pic-III for 48 h with non-cytotoxic concentrations (RMF-621 cells: 20 µg/mL and Caco-2 cells: 50 µg/mL) and secreted cytokine levels (IL-10, IL-8, IL-1β, and TNFα) in the culture medium were determined by CBA assay. Data are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. control; n.s., not significant.</p>
Full article ">Figure 13
<p>Pic-III sensitizes to the BH3 mimetic drug ABT-199 in MDA-MB-231 cancer cells. Effect of Pic-III (50 µg/mL), ABT-199 (15 µM), and combination Pic-III + ABT-199 on the viability of MDA-MB-231 at 48 h of exposition. Data are expressed as means ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. control.</p>
Full article ">
13 pages, 675 KiB  
Article
Venetoclax Use in Paediatric Haemato-Oncology Centres in Poland: A 2022 Survey
by Katarzyna Bobeff, Agata Pastorczak, Zuzanna Urbanska, Walentyna Balwierz, Edyta Juraszewska, Jacek Wachowiak, Katarzyna Derwich, Magdalena Samborska, Krzysztof Kalwak, Iwona Dachowska-Kalwak, Paweł Laguna, Iwona Malinowska, Katarzyna Smalisz, Jolanta Gozdzik, Aleksandra Oszer, Bartosz Urbanski, Maciej Zdunek, Tomasz Szczepanski, Wojciech Mlynarski and Szymon Janczar
Children 2023, 10(4), 745; https://doi.org/10.3390/children10040745 - 19 Apr 2023
Cited by 3 | Viewed by 1942
Abstract
Venetoclax, the best established BH3-mimetic, is a practice-changing proapoptotic drug in blood cancers in adults. In paediatrics the data are fewer but exciting results were recently presented in relapsed or refractory leukaemias demonstrating significant clinical activity. Importantly, the in-terventions could be potentially molecularly [...] Read more.
Venetoclax, the best established BH3-mimetic, is a practice-changing proapoptotic drug in blood cancers in adults. In paediatrics the data are fewer but exciting results were recently presented in relapsed or refractory leukaemias demonstrating significant clinical activity. Importantly, the in-terventions could be potentially molecularly guided as vulnerabilities to BH3-mimetics were re-ported. Currently venetoclax is not incorporated into paediatric treatment schedules in Poland but it has been already used in patients that failed conventional therapy in Polish paediatric haemato-oncology departments. The aim of the study was to gather clinical data and correlates of all paediatric patients treated so far with venetoclax in Poland. We set out to gather this experience to help choose the right clinical context for the drug and stimulate further research. The questionnaire regarding the use of venetoclax was sent to all 18 Polish paediatric haemato-oncology centres. The data as available in November 2022 were gathered and analysed for the diagnoses, triggers for the intervention, treatment schedules, outcomes and molecular associations. We received response from 11 centres, 5 of which administered venetoclax to their patients. Clinical benefit, in most cases consistent with hematologic complete remission (CR), was reported in 5 patients out of ten, whereas 5 patient did not show clinical benefit from the intervention. Importantly, patients with CR included subtypes expected to show venetoclax vulnerability, such as poor-prognosis ALL with TCF::HLF fusion. We believe BH3-mimetics have clinical activity in children and should be available to pae-diatric haemato-oncology practitioners in well-selected applications. Full article
(This article belongs to the Section Pediatric Hematology & Oncology)
Show Figures

Figure 1

Figure 1
<p>Results of single nucleotide polymorphism array analysis of the leukaemic genomes in selected patients treated with venetoclax. (<b>A</b>) Whole genome view showing monosomy 7 and LOH of chromosome 17q in patient 1. (<b>B</b>) Chromosomal microarray analysis data plot for chromosome 17 displaying the gain of 17q22–q25.3, which involves exon 4 of the HLF gene in patient 2. (<b>C</b>) CMA data plot showing the 17p13.1–p13.3 deletion encompassing the TP53 gene in patient 3. Copy number abnormalities were identified by a decrease (deletion) or an increase (gain) in the Log2 ratio. The coordinates are in reference to GRCh37/hg19. 1.</p>
Full article ">
10 pages, 2115 KiB  
Article
Synergistic Interactions between the Hypomethylating Agent Thio-Deoxycytidine and Venetoclax in Myelodysplastic Syndrome Cells
by Xiaoyan Hu, Lin Li, Jewel Nkwocha, Kanika Sharma, Liang Zhou and Steven Grant
Hematol. Rep. 2023, 15(1), 91-100; https://doi.org/10.3390/hematolrep15010010 - 2 Feb 2023
Cited by 3 | Viewed by 2015
Abstract
Interactions between the novel hypomethylating agent (HMA) thio-deoxycytidine (T-dCyd) and the BCL-2 antagonist ABT-199 (venetoclax) have been examined in human myelodysplastic syndrome (MDS) cells. The cells were exposed to agents alone or in combination, after which apoptosis was assessed, and a Western blot [...] Read more.
Interactions between the novel hypomethylating agent (HMA) thio-deoxycytidine (T-dCyd) and the BCL-2 antagonist ABT-199 (venetoclax) have been examined in human myelodysplastic syndrome (MDS) cells. The cells were exposed to agents alone or in combination, after which apoptosis was assessed, and a Western blot analysis was performed. Co-administration of T-dCyd and ABT-199 was associated with the down-regulation of DNA methyltransferase 1 (DNMT1) and synergistic interactions documented by a Median Dose Effect analysis in multiple MDS-derived lines (e.g., MOLM-13, SKM-1, and F-36P). Inducible BCL-2 knock-down significantly increased T-dCyd’s lethality in MOLM-13 cells. Similar interactions were observed in the primary MDS cells, but not in the normal cord blood CD34+ cells. Enhanced killing by the T-dCyd/ABT-199 regimen was associated with increased reactive oxygen species (ROS) generation and the down-regulation of the anti-oxidant proteins Nrf2 and HO-1, as well as BCL-2. Moreover, ROS scavengers (e.g., NAC) reduced lethality. Collectively, these data suggest that combining T-dCyd with ABT-199 kills MDS cells through an ROS-dependent mechanism, and we argue that this strategy warrants consideration in MDS therapy. Full article
Show Figures

Figure 1

Figure 1
<p>MDS–derived MOLM–13 and MDS SKM–1 and F–36P cells were exposed to varying concentrations of ABT–199 and T–dCyd or Aza–T–dCyd administered at a fixed ratio for 48 h, after which Median Dose Effect analysis was used to characterize concentration index (CI) values in relation to the fraction affected (FA). CI values &lt; 1.0 denote synergistic interactions (<b>A</b>–<b>E</b>). (<b>F</b>) Cell death was monitored by Annexin V/PI staining and FCM. For <span class="html-italic">p</span> values, * = &lt;0.05; ** = &lt;0.01. (<b>G</b>,<b>H</b>) Cleaved PARP, caspase 3, and DNMT1 were detected by WB in MOLM–13 and SKM–1 cells exposed to T–dCyd ± ABT–199; β-actin controls are shown to document equivalent loading and transfer.</p>
Full article ">Figure 2
<p>Dox–inducible TET–on MOLM–13 cells expressing BCL–2 shRNA were exposed to the indicated concentrations of T–dCyd in the presence or absence of doxycycline. Cell death was determined by annexin V/PI staining and FCM. Inset: expression of BCL–2 by WB with or without doxycycline. * (<span class="html-italic">p</span> &lt; 0.05) and ** (<span class="html-italic">p</span> &lt; 0.01) = significantly greater than values for cells without doxycycline (<b>A</b>); cleaved PARP and Caspase 3 were detected by WB (<b>B</b>).</p>
Full article ">Figure 3
<p>Bone marrow cells from an MDS patient (MDS-EB-1) were exposed (24 h) to 500 nM T-dCyd with or without 50 nM venetoclax, after which they were stained with PE-CD34<sup>+</sup> (red) and Annexin (green) Abs. Cells were then viewed under an immunofluorescence microscope at 200× magnification (<b>A</b>). Cleaved PARP was detected by WB (<b>B</b>). Cord blood normal CD34<sup>+</sup> cells (<span class="html-italic">n</span> = 4) were exposed (24 h) to 1 uM T-dCyd with or without 50 nM venetoclax, after which the CD34<sup>+</sup> gated population was monitored for cell viability using Annexin V and 7-AAD staining (<b>C</b>). The percentage of viable cells in control and treated populations is reflected in the dot plot (<b>D</b>).</p>
Full article ">Figure 4
<p>MOLM–13 cells were exposed (24 h) to the designated concentrations of ABT and T–dCyd, after which mitochondrial ROS in the viable cell population was monitored by FCM and MFI (<b>A</b>). (<b>B</b>) Effects of NAC on mitoROS (<b>B</b>) and viability (<b>C</b>) following the same exposure. For <span class="html-italic">p</span> values, * = &lt;0.05; *** = &lt;0.001.</p>
Full article ">Figure 5
<p>(<b>A</b>) SKM-1 cells were exposed (24 h) to ABT-199 ± T-dCyd (1 μM each), after which cytosolic and nuclear fractions were monitored for NRF2, HO-1, and BCL-2 expressions by WB analysis. Images were quantified and analyzed by using ImageJ software. Data were normalized to the ratio of the indicated protein and β-actin/P84 versus control. (<b>B</b>) MOLM-13 cells were exposed (24 h) to T-dCyd (500 nM) ± 2 nM ABT-199, after which the nuclear disposition of NRF2 was monitored by immunofluorescence microscopy.</p>
Full article ">
18 pages, 2857 KiB  
Article
Rationale for Combining the BCL2 Inhibitor Venetoclax with the PI3K Inhibitor Bimiralisib in the Treatment of IDH2- and FLT3-Mutated Acute Myeloid Leukemia
by Katja Seipel, Yvo Brügger, Harpreet Mandhair, Ulrike Bacher and Thomas Pabst
Int. J. Mol. Sci. 2022, 23(20), 12587; https://doi.org/10.3390/ijms232012587 - 20 Oct 2022
Cited by 5 | Viewed by 3004
Abstract
In October 2020, the FDA granted regular approval to venetoclax (ABT-199) in combination with hypomethylating agents for newly-diagnosed acute myeloid leukemia (AML) in adults 75 years or older, or in patients with comorbidities precluding intensive chemotherapy. The treatment response to venetoclax combination treatment, [...] Read more.
In October 2020, the FDA granted regular approval to venetoclax (ABT-199) in combination with hypomethylating agents for newly-diagnosed acute myeloid leukemia (AML) in adults 75 years or older, or in patients with comorbidities precluding intensive chemotherapy. The treatment response to venetoclax combination treatment, however, may be short-lived, and leukemia relapse is the major cause of treatment failure. Multiple studies have confirmed the upregulation of the anti-apoptotic proteins of the B-cell lymphoma 2 (BCL2) family and the activation of intracellular signaling pathways associated with resistance to venetoclax. To improve treatment outcome, compounds targeting anti-apoptotic proteins and signaling pathways have been evaluated in combination with venetoclax. In this study, the BCL-XL inhibitor A1331852, MCL1-inhibitor S63845, dual PI3K-mTOR inhibitor bimiralisib (PQR309), BMI-1 inhibitor unesbulin (PTC596), MEK-inhibitor trametinib (GSK1120212), and STAT3 inhibitor C-188-9 were assessed as single agents and in combination with venetoclax, for their ability to induce apoptosis and cell death in leukemic cells grown in the absence or presence of bone marrow stroma. Enhanced cytotoxic effects were present in all combination treatments with venetoclax in AML cell lines and AML patient samples. Elevated in vitro efficacies were observed for the combination treatment of venetoclax with A1331852, S63845 and bimiralisib, with differing response markers for each combination. For the venetoclax and bimiralisib combination treatment, responders were enriched for IDH2 and FLT3 mutations, whereas non-responders were associated with PTPN11 mutations. The combination of PI3K/mTOR dual pathway inhibition with bimiralisib and BCL2 inhibition with venetoclax has emerged as a candidate treatment in IDH2- and FLT3-mutated AML. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of signaling pathways in myeloid cells. The inducible growth factor receptor FLT3 signals via PI3K-AKT-mTOR and RAS-MEK-ERK (black arrows). FLT3-ITD, a constitutively active growth factor receptor, additionally induces PTPN11-STAT5 (red arrows). Activated cytokine receptors signal via Janus kinase (JAK)-signal transducer and activator of transcription (STAT) pathway. Signal transduction leads to inhibition of the tumor suppressor p53 and induction of the anti-apoptotic BH3 proteins BCL2, BCL-XL, and MCL1, thereby promoting proliferation and cell growth of myeloid cells. Oncogenic functions are indicated in red, tumor suppressor functions in green, chemical inhibitors in blue.</p>
Full article ">Figure 2
<p>Susceptibility of leukemic cells to venetoclax combination treatment. Cell viability was determined in one lymphoma and seven AML cell lines treated for 20 h with single compounds and in combination with 100 nM venetoclax (VC) and 1 μM A1331825 (<b>A</b>), 1 μM bimiralisib (<b>B</b>), 1 μM C188-9 (<b>C</b>), 200 nM PTC596 (<b>D</b>), 100 nM S63845 (<b>E</b>), or 100 nM trametinib (<b>F</b>). Significance was calculated in a graph pad prism using grouped analysis with paired <span class="html-italic">t</span>-test comparing cell viabilities of VC treated and combination treated cells. A significance level of 0.05 indicates a 5% risk of concluding that a difference exists when there is no actual difference.</p>
Full article ">Figure 3
<p>Susceptibility of AML cell lines to targeted therapies in the presence of bone marrow stroma. MOLM-13 (<b>A</b>), ML-2 (<b>B</b>), OCI-AML3 (<b>C</b>), and SKM-1 cells (<b>D</b>) were treated for 20 h with single compounds and in combination with venetoclax (VC), A1331825 (A), bimiralisib (B), C-188-9 (C), PTC596 (P), S63845 (S), or trametinib (T). Cell viability was determined in AML cells grown in the absence or presence of HS-5 stroma. Concentrations of inhibitors were 100 nM for venetoclax, S63845, PTC596 and trametinib, 1 μM for bimiralisib, C-188-9, and A133825. Significance was calculated in a graph pad prism using grouped analysis with multiple unpaired <span class="html-italic">t</span>-test comparing cell viabilities of treated cells grown in the absence or presence of HS-5 stroma. Significance denoted for <span class="html-italic">p</span> &lt; 0.05 (*); <span class="html-italic">p</span> &lt; 0.005 (**); <span class="html-italic">p</span> &lt; 0.0005 (***); no significance denoted for <span class="html-italic">p</span> &gt; 0.05 (ns). A significance level of 0.05 indicates a 5% risk of concluding that a difference exists when there is no actual difference.</p>
Full article ">Figure 4
<p>Induction of apoptosis, cell cycle arrest and cell death in MOLM_13 cells treated with venetoclax and bimiralisib. Cytometric analysis of MOLM-13 cells treated for 20 h with 10 nM or 50 nM venetoclax (VC) and 0.1 μM or 1 μM bimiralisib alone or in combination. Depending on DAPI staining intensity cells were classified as subG1, G0/G1, S phase, or G2 phase (<b>A</b>). Treatment-induced cell death (subG1 fraction) (<b>B</b>), and G1 cell cycle arrest (<b>C</b>). Depending on Annexin V and PI staining intensity, cells were classified as vital (Ann lo, PI lo), early apoptotic (Ann hi, PI lo), late apoptotic (Ann hi, PI hi) or necrotic (Ann lo, PI hi) (<b>D</b>). Treatment-induced loss of vital cells (<b>E</b>) and amount of apoptotic cells (<b>F</b>) were significanty enhanced in the combination treatment. Significance of differences in median values was calculated by the Mann–Whitney test. Significance denoted for <span class="html-italic">p</span> &lt; 0.05 (*); <span class="html-italic">p</span> &lt; 0.005 (**); no significance denoted for <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 5
<p>Hematological cells in vitro response to venetoclax and various combination treatments. Cell viability was determined in mononuclear cells isolated from AML patients or healthy donor (HD) peripheral blood or bone marrow after 20 h treatment. The patient samples were sorted into two equally sized groups, one with major (strong) response (SR) and one with minor (normal) response (NR). Number of samples in each group are indicated in parentheses. Cells were treated in vitro with 100 nM venetoclax (VC) only (<b>A</b>), 100 nM venetoclax in combination with 1 μM A1331825 (<b>B</b>), 1 μM bimiralisib (<b>C</b>), 100 nM S63845 (<b>D</b>), 100 nM trametinib (<b>E</b>), 1 μM C-188-9 (<b>F</b>), or 200 nM PTC596 (<b>G</b>). Significance of differences in median values was calculated by Mann–Whitney test.</p>
Full article ">Figure 6
<p>Response markers in venetoclax monotherapy. Cell viability was determined in mononuclear cells isolated from AML patients peripheral blood or bone marrow after 20 h treatment with 100 nM venetoclax. Hematological cells were grouped according to single gene mutation status of IDH2 (<b>A</b>), NPM1 (<b>B</b>), IDH2 and NPM1 (<b>C</b>) FLT3 (<b>D</b>), TET2 (<b>E</b>), TP53 (<b>F</b>), ASXL1 (<b>G</b>), PTPN11 (<b>H</b>), RUNX1 (<b>I</b>) peripheral blast cell percentage (<b>J</b>), bone marrow blast cell percentage (<b>K</b>), and CD34 positivity (<b>L</b>). Number of samples in each group are indicated in parentheses. Significance of differences in median values was calculated by Mann–Whitney test.</p>
Full article ">Figure 7
<p>Response markers in venetoclax and A1331825 combination treatment. Cell viability was determined in mononuclear cells isolated from AML patients peripheral blood or bone marrow after 20 h treatment with 100 nM venetoclax and 1 μM A1331825 (<b>A</b>). Hematological cells were grouped according to single gene mutation status of IDH2 (<b>B</b>), FLT3 (<b>C</b>), ASXL1 (<b>D</b>), TET2 (<b>E</b>), PTPN11 (<b>F</b>), TP53 (<b>G</b>), peripheral blast cell percentage (<b>H</b>), bone marrow blast cell percentage (<b>I</b>). Number of samples in each group are indicated in parentheses. Significance of differences in median values was calculated by Mann–Whitney test.</p>
Full article ">Figure 8
<p>Response markers in venetoclax and bimiralisib (PQR309) combination treatment. Cell viability was determined in mononuclear cells isolated from AML patients peripheral blood or bone marrow after 20 h treatment with 100 nM venetoclax and 1 μM bimiralisib (<b>A</b>). Hematological cells were grouped according to single gene mutation status of IDH2 (<b>B</b>), NPM1 (<b>C</b>), FLT3 (<b>D</b>), PTPN11 (<b>E</b>), TET2 (<b>F</b>), ASXL1 (<b>G</b>), peripheral blast cell percentage (<b>H</b>), and bone marrow blast cell percentage (<b>I</b>). Number of samples in each group are indicated in parentheses. Significance of differences in median values was calculated by Mann–Whitney test.</p>
Full article ">Figure 9
<p>Response markers in venetoclax and S63845 combination treatment. Cell viability was determined in mononuclear cells isolated from AML patients peripheral blood or bone marrow after 20 h treatment with 100 nM venetoclax and 100 nM S63845 (<b>A</b>). Hematological cells were grouped according to single gene mutation status of IDH2 (<b>B</b>), NPM1 (<b>C</b>), FLT3 (<b>D</b>), TET2 (<b>E</b>), PTPN11 (<b>F</b>), ASXL1 (<b>G</b>), peripheral blast cell percentage (<b>H</b>), and bone marrow blast cell percentage (<b>I</b>). Number of samples in each group are indicated in parentheses. Significance of differences in median values was calculated by Mann–Whitney test.</p>
Full article ">
16 pages, 1840 KiB  
Article
Mitochondrial Kv1.3 Channels as Target for Treatment of Multiple Myeloma
by Stephanie Kadow, Fabian Schumacher, Melanie Kramer, Gabriele Hessler, René Scholtysik, Sara Oubari, Patricia Johansson, Andreas Hüttmann, Hans Christian Reinhardt, Burkhard Kleuser, Mario Zoratti, Andrea Mattarei, Ildiko Szabò, Erich Gulbins and Alexander Carpinteiro
Cancers 2022, 14(8), 1955; https://doi.org/10.3390/cancers14081955 - 13 Apr 2022
Cited by 8 | Viewed by 2320
Abstract
Despite several new developments in the treatment of multiple myeloma, all available therapies are only palliative without curative potential and all patients ultimately relapse. Thus, novel therapeutic options are urgently required to prolong survival of or to even cure myeloma. Here, we show [...] Read more.
Despite several new developments in the treatment of multiple myeloma, all available therapies are only palliative without curative potential and all patients ultimately relapse. Thus, novel therapeutic options are urgently required to prolong survival of or to even cure myeloma. Here, we show that multiple myeloma cells express the potassium channel Kv1.3 in their mitochondria. The mitochondrial Kv1.3 inhibitors PAPTP and PCARBTP are efficient against two tested human multiple myeloma cell lines (L-363 and RPMI-8226) and against ex vivo cultured, patient-derived myeloma cells, while healthy bone marrow cells are spared from toxicity. Cell death after treatment with PAPTP and PCARBTP occurs via the mitochondrial apoptotic pathway. In addition, we identify up-regulation of the multidrug resistance pump MDR-1 as the main potential resistance mechanism. Combination with ABT-199 (venetoclax), an inhibitor of Bcl2, has a synergistic effect, suggesting that mitochondrial Kv1.3 inhibitors could potentially be used as combination partner to venetoclax, even in the treatment of t(11;14) negative multiple myeloma, which represent the major part of cases and are rather resistant to venetoclax alone. We thus identify mitochondrial Kv1.3 channels as druggable targets against multiple myeloma. Full article
(This article belongs to the Section Molecular Cancer Biology)
Show Figures

Figure 1

Figure 1
<p>Mitochondrial Kv1.3 is expressed in the multiple myeloma cell lines L-363 and RPMI-8226. PAPTP and PCARBTP efficiently kill human multiple myeloma cells. (<b>a</b>) Whole cell lysates (left, 50 µg protein/lane) were lysed in RIPA-Buffer, mitochondria from L-363, RPMI-8226 and Jurkat cells (right, 30 µg protein/lane) were enriched as described; both were supplemented with 5× reducing SDS sample buffer, boiled and separated on 8.5% SDS-polyacrylamide gels and blotted on nitrocellulose membranes. Membranes were divided and, after blocking, membranes were incubated 1 h with anti-Kv1.3 or anti-Tim23 primary antibodies, respectively. After extensively washing, blots were incubated with secondary antibodies for 1 h, washed and developed with the Roti–Lumin system. Multiple myeloma cell lines RPMI-8226 and L-363 were treated with increasing concentrations PAPTP (<b>b</b>) or PCARBTP (<b>c</b>), solvent control (0.1% DMSO), 2 µM staurosporine as a positive control and 2 µM margatoxin as a membrane-impermeable Kv1.3 blocker. After 24 h, cells were stained with Annexin V-APC/7AAD and examined for cell death by flow cytometry. Results are reported as percentages with respect to untreated samples ± SD, (<span class="html-italic">n</span> = 3 independent experiments, each experiment in triplicate). Graph insert EC50: EC50 values of the indicated compounds in RPMI-8226 and L-363 were calculated by using GraphPad Prism version 9.3 for Windows (GraphPad Software, San Diego, CA, USA).</p>
Full article ">Figure 2
<p>PAPTP and PCARBTP efficiently kill patient-derived myeloma cells ex vivo, while healthy bone marrow cells are spared from toxic effects. Isolated mononuclear bone marrow cells were cultured in IMDM 20% FCS at a concentration of 2 × 10<sup>6</sup>/mL and incubated with PAPTP, PCARBTP as indicated (<b>a</b>,<b>b</b>), respectively or 0.1% DMSO (solvent control). After 24 h cells were harvested and stained with Annexin V-APC and 7-AAD for cell death and simultaneously for CD19, CD38, CD138, CD45, CD269 and CD56 to properly detect myeloma cells. Samples were analyzed by flow cytometry, malignant plasma cells were defined as CD38<sup>+</sup>, CD138<sup>+</sup>, CD19<sup>−</sup>, CD269<sup>+</sup> and CD56<sup>+</sup>. Shown is the mean ± SD of viable cells in % of untreated control, each dot representing the measure of one patient sample, <span class="html-italic">n</span> = 7–14, ns not significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001 Non-parametric ANOVA followed by Dunn’s multiple comparison and a Kruskal–Wallis test. ns: not significant.</p>
Full article ">Figure 3
<p>Inhibition of mitoKv1.3 with PAPTP and PCARBTP triggers apoptosis in human multiple myeloma cell lines L-363 and RPMI-8226. (<b>a</b>) Dissipation of the mitochondrial membrane potential was monitored following the TMRM signal over time. CCCP was used as the positive control. (<b>b</b>) Mitochondrial ROS production was monitored by the development of MitoSOX fluorescence over time. Antimycin A was used as the positive control. Values are reported as percentage MFI with respect to solvent control. (<b>c</b>) Caspase 3/7-activation was measured by flow cytometry 4, 8 and 12 h after treatment with PAPTP (1 µM), PCARPTP (2 µM) or staurosporine (0.5 µM) as a positive control. Shown is the mean ± SD of % caspase 3/7+ cells of 3 independent experiments, each experiment in triplicate.</p>
Full article ">Figure 4
<p>Resistance to PCARBTP is mediated by MDR-1, RPMI-8226-R1.0 and L-363-R3.5 are cross-resistant to PAPTP. (<b>a</b>,<b>b</b>) Parental L-363 and RPMI-8226 multiple myeloma cells and PCARBTP-resistant L-363-R3.5 and RPMI-8226-R1.0 cell lines were incubated with PCARBTP or 0.1% DMSO control in the presence or absence of elacridar as indicated. After 24 h, cell death was determined by flow cytometry with Annexin V-APC/7AAD-staining. Resistant RPMI-8226-R1.0 (<b>c</b>) and L-363-R3.5 cells (<b>d</b>) were treated with increasing concentrations PAPTP ± 100 nM elacridar for 24 h. After 24 h, cell death was determined by flow cytometry with Annexin V-APC/7AAD-staining. Values are normalized to solvent-treated control. Shown is the mean ± SD, (<span class="html-italic">n</span> = 3 independent experiments, each in duplicate).</p>
Full article ">Figure 4 Cont.
<p>Resistance to PCARBTP is mediated by MDR-1, RPMI-8226-R1.0 and L-363-R3.5 are cross-resistant to PAPTP. (<b>a</b>,<b>b</b>) Parental L-363 and RPMI-8226 multiple myeloma cells and PCARBTP-resistant L-363-R3.5 and RPMI-8226-R1.0 cell lines were incubated with PCARBTP or 0.1% DMSO control in the presence or absence of elacridar as indicated. After 24 h, cell death was determined by flow cytometry with Annexin V-APC/7AAD-staining. Resistant RPMI-8226-R1.0 (<b>c</b>) and L-363-R3.5 cells (<b>d</b>) were treated with increasing concentrations PAPTP ± 100 nM elacridar for 24 h. After 24 h, cell death was determined by flow cytometry with Annexin V-APC/7AAD-staining. Values are normalized to solvent-treated control. Shown is the mean ± SD, (<span class="html-italic">n</span> = 3 independent experiments, each in duplicate).</p>
Full article ">Figure 5
<p>Combined treatment of multiple myeloma cell lines with PAPTP + ABT-199 improves killing compared to PAPTP alone. Cells were cultivated in the presence of different concentrations of PAPTP ± ABT-199 as indicated. After 24 h, cell death was analyzed by flow cytometry with Annexin V-APC/7AAD staining. (<b>a</b>) In the Dose–Response Matrix Annexin V/7AAD, positive cells are given in % normalized to solvent treated control. Shown is the mean ± SD, (<span class="html-italic">n</span> = 3 independent experiments, each in duplicates). (<b>b</b>) Synergies were calculated using the onlinetool <a href="https://synergyfinder.fimm.fi" target="_blank">https://synergyfinder.fimm.fi</a> accessed on 24 February 2022 [<a href="#B26-cancers-14-01955" class="html-bibr">26</a>] and expressed as zero interaction potency-score [<a href="#B27-cancers-14-01955" class="html-bibr">27</a>].</p>
Full article ">
17 pages, 2345 KiB  
Article
Targeting the Non-Canonical NF-κB Pathway in Chronic Lymphocytic Leukemia and Multiple Myeloma
by Thomas A. Burley, Emma Kennedy, Georgia Broad, Melanie Boyd, David Li, Timothy Woo, Christopher West, Eleni E. Ladikou, Iona Ashworth, Christopher Fegan, Rosalynd Johnston, Simon Mitchell, Simon P. Mackay, Andrea G. S. Pepper and Chris Pepper
Cancers 2022, 14(6), 1489; https://doi.org/10.3390/cancers14061489 - 15 Mar 2022
Cited by 8 | Viewed by 3398
Abstract
In this study, we evaluated an NF-κB inducing kinase (NIK) inhibitor, CW15337, in primary chronic lymphocytic leukemia (CLL) cells, CLL and multiple myeloma (MM) cell lines and normal B- and T-lymphocytes. Basal NF-κB subunit activity was characterized using an enzyme linked immunosorbent assay [...] Read more.
In this study, we evaluated an NF-κB inducing kinase (NIK) inhibitor, CW15337, in primary chronic lymphocytic leukemia (CLL) cells, CLL and multiple myeloma (MM) cell lines and normal B- and T-lymphocytes. Basal NF-κB subunit activity was characterized using an enzyme linked immunosorbent assay (ELISA), and the effects of NIK inhibition were then assessed in terms of cytotoxicity and the expression of nuclear NF-κB subunits following monoculture and co-culture with CD40L-expressing fibroblasts, as a model of the lymphoid niche. CW15337 induced a dose-dependent increase in apoptosis, and nuclear expression of the non-canonical NF-κB subunit, p52, was correlated with sensitivity to CW15337 (p = 0.01; r2 = 0.39). Co-culture on CD40L-expressing cells induced both canonical and non-canonical subunit expression in nuclear extracts, which promoted in vitro resistance against fludarabine and ABT-199 (venetoclax) but not CW15337. Furthermore, the combination of CW15337 with fludarabine or ABT-199 showed cytotoxic synergy. Mechanistically, CW15337 caused the selective inhibition of non-canonical NF-κB subunits and the transcriptional repression of BCL2L1, BCL2A1 and MCL1 gene transcription. Taken together, these data suggest that the NIK inhibitor, CW15337, exerts its effects via suppression of the non-canonical NF-κB signaling pathway, which reverses BCL2 family-mediated resistance in the context of CD40L stimulation. Full article
(This article belongs to the Special Issue Therapeutic Targets in Chronic Lymphocytic Leukemia)
Show Figures

Figure 1

Figure 1
<p>Primary CLL cells and CLL and myeloma cell lines are all susceptible to NF-κB-inducing kinase (NIK) inhibition. (<b>A</b>) An example of Annexin V and 7-AAD bivariate plots obtained from primary CLL cells treated with increasing concentrations of CW15337. A concentration-dependent increase in the proportion of Annexin V<sup>+</sup>/7-AAD<sup>−</sup> and Annexin V<sup>+</sup>/7-AAD<sup>+</sup> was observed. (<b>B</b>) Sigmoidal dose–response curves illustrating the comparative effects of CW15337 on the H929, U266, RPMI8226, JJN3, MEC-1 cell lines and primary CLL cells. (<b>C</b>) Comparative analysis of the mean LD<sub>50</sub> values for CW15337 in the four multiple myeloma cell lines revealed heterogeneous sensitivity to the effects of CW15337 between the cell lines. (<b>D</b>) CW15337 was significantly more potent in MEC-1 cells and primary CLL cells when compared with normal B- and T-lymphocytes. (<b>E</b>) CW15337 inhibited the migration of MEC-1 cells in a concentration-dependent manner. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05. ns = not significant.</p>
Full article ">Figure 2
<p>CW15337 is preferentially cytotoxic in CLL cells that carry a BIRC3 or NOTCH1 mutation. (<b>A</b>) All of the CLL samples tested had low micromolar LD<sub>50</sub> values, with a mean of 1.44 μM for the entire cohort. (<b>B</b>) Comparative analysis of the sensitivity to CW15337 in CLL prognostic subsets revealed that samples derived from patients carrying a BIRC3 or NOTCH1 mutation were particularly sensitive to CW15337; the same samples showed increased in vitro resistance to fludarabine. *** <span class="html-italic">p</span> &lt; 0.001, and * <span class="html-italic">p</span> &lt; 0.05. ns = not significant.</p>
Full article ">Figure 3
<p>CW15337 preferentially inhibits the nuclear expression of p52 and RelB NF-κB subunits. (<b>A</b>) Each cell line evaluated showed a distinct pattern of constitutive NF-κB subunit activation in nuclear extracts. (<b>B</b>) Exposure to CW15337 resulted in a concentration-dependent reduction in two non-canonical NF-κB subunits, p52 and RelB. (<b>C</b>) MEC-1 cells treated with increasing concentrations of CW15337 showed a marked inhibition of p100 processing to p52. (<b>D</b>) Constitutive p52 expression in nuclear extracts from primary CLL samples (<span class="html-italic">n</span> = 15) were correlated with in vitro sensitivity to CW15337. **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05. ns = not significant.</p>
Full article ">Figure 4
<p>Co-culture on CD40L-expressing 3T3 cells drives MEC-1 cell activation and proliferation, which was reversed by the addition of CW15337. (<b>A</b>) Untransfected NIH/3T3 cells or NIH/3T3 cells transfected with human CD40L were plated prior to the addition of malignant B cells (1:10 ratio). (<b>B</b>) MEC-1 cells co-cultured on CD40L 3T3 cells showed significantly increased cell growth when compared to those cultured on untransfected 3T3 cells. MEC-1 cell growth was abolished by the addition of CW15337 to the CD40L 3T3 co-cultures. (<b>C</b>–<b>E</b>) In keeping with these findings, MEC-1 cells showed a marked increase in Ki67, HLA-DR and CD69 when co-cultured on CD40L 3T3 cells, which was reversed by the addition of CW15337. (<b>F</b>,<b>G</b>) CW15337 induced a concentration-dependent G1 arrest in MEC-1 cells co-cultured on CD40L 3T3 cells. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>CW15377 preferentially inhibits the non-canonical NF-κB pathway in tumor cells co-cultured on CD40L 3T3 cells. (<b>A</b>) Nuclear extracts were prepared from primary CLL cells following monoculture, co-culture on untransfected 3T3 cells and CD40L 3T3 cells. Extracts from CLL cells co-cultured on CD40L 3T3 cells showed a marked increase in both canonical and non-canonical subunit. (<b>B</b>,<b>C</b>) RNA sequencing showed a marked transcriptional activation of known NF-κB-target genes. Furthermore, NF-κB regulated genes were significantly overrepresented in the differentially expressed gene list. CD40L 3T3 co-culture of (<b>D</b>) MEC-1 cells and (<b>E</b>) RPMI8226 cells for 24 h with or without the addition of CW15337 for the last 8 h, showed that CW15337 preferentially inhibited the nuclear expression of p52 and RelB in a concentration-dependent manner. (<b>F</b>) QRT-PCR analysis of primary CLL cells (<span class="html-italic">n</span> = 6) confirmed that the addition of CW15337 could repress the transcriptional activation of BCL2L1, MCL1 and to a lesser extent, BCL2A1 induced by CD40L 3T3 co-culture. FDR = false discovery rate. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05. ns = not significant.</p>
Full article ">Figure 6
<p>CW15337 is synergistic with fludarabine and ABT-199 in the setting of CD40L 3T3 co-culture. Primary CLL cells showed a marked increase in resistance to (<b>A</b>) fludarabine and (<b>B</b>) ABT-199 when co-cultured on CD40L 3T3 cells. (<b>C</b>) In contrast, the sensitivity to CW15337 was not significantly affected by CD40L co-culture. The combination of CW15337 with (<b>D</b>) fludarabine and (<b>E</b>) ABT-199 showed synergistic interactions. (<b>F</b>) The combination of CW15337 and ABT-199 (100:1) was shown to be particularly synergistic under CD40L co-culture conditions, using SynergyFinder software (<a href="https://synergyfinder.fimm.fi" target="_blank">https://synergyfinder.fimm.fi</a>, accessed on 30 January 2022). BLISS scores ≥ 10 indicate synergy, the average synergy score of CW15337 and ABT-199 was 15.06. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
18 pages, 5114 KiB  
Article
Venetoclax-Resistant MV4-11 Leukemic Cells Activate PI3K/AKT Pathway for Metabolic Reprogramming and Redox Adaptation for Survival
by Hind A. Alkhatabi, Samir F. Zohny, Mohammed Razeeth Shait Mohammed, Hani Choudhry, Mohd Rehan, Aamir Ahmad, Farid Ahmed and Mohammad Imran Khan
Antioxidants 2022, 11(3), 461; https://doi.org/10.3390/antiox11030461 - 25 Feb 2022
Cited by 13 | Viewed by 4332
Abstract
Venetoclax (ABT199) is a selective B-cell lymphoma 2 (BCL-2) inhibitor. The US FDA recently approved it to be used in combination with low-dose cytarabine or hypomethylating agents in acute myeloid leukemia (AML) or elderly patients non-eligible for chemotherapy. However, acquiring resistance to venetoclax [...] Read more.
Venetoclax (ABT199) is a selective B-cell lymphoma 2 (BCL-2) inhibitor. The US FDA recently approved it to be used in combination with low-dose cytarabine or hypomethylating agents in acute myeloid leukemia (AML) or elderly patients non-eligible for chemotherapy. However, acquiring resistance to venetoclax in AML patients is the primary cause of treatment failure. To understand the molecular mechanisms inherent in the resistance to BCL-2 inhibitors, we generated a venetoclax-resistant cell line model and assessed the consequences of this resistance on its metabolic pathways. Untargeted metabolomics data displayed a notable impact of resistance on the PI3K/AKT pathway, the Warburg effect, glycolysis, the TCA cycle, and redox metabolism. The resistant cells showed increased NADPH and reduced glutathione levels, switching their energy metabolism towards glycolysis. PI3K/AKT pathway inhibition shifted resistant cells towards oxidative phosphorylation (OXPHOS). Our results provide a metabolic map of resistant cells that can be used to design novel metabolic targets to challenge venetoclax resistance in AML. Full article
(This article belongs to the Special Issue Mitochondrial Superoxide Dismutase in Cancer Biology and Therapy)
Show Figures

Figure 1

Figure 1
<p>ABT199-R cells are highly resistant to venetoclax. (<b>A</b>) MV4-11 and ABT-199R were treated with increasing doses of venetoclax (0–20 µM) for 48 h and assessed with a cell titer blue proliferation assay. The curves indicate the percentage survival of each cell line to increasing doses of venetoclax. (<b>B</b>) MV4-11 and ABT199-R cells were treated with indicated concentrations of venetoclax for 48 h and stained for an Annexin-V/7AAD assay. The percentage values determine the early and late apoptotic populations.</p>
Full article ">Figure 2
<p>Metabolomic analysis of MV4-11 cells vs. ABT199-R cells. (<b>A</b>) The metabolic heatmap profile of differentially accumulated metab olites between sensitive MV4-11 and resistant ABT199-R. (<b>B</b>) Volcano plots of metabolic differentiation between MV4-11 and ABT199-R. (<b>C</b>) ABT199-R cells were treated with several concentrations of PKI-402 for 48 h and assessed with cell titer blue viability assay. The curve indicates the IC50 value.</p>
Full article ">Figure 2 Cont.
<p>Metabolomic analysis of MV4-11 cells vs. ABT199-R cells. (<b>A</b>) The metabolic heatmap profile of differentially accumulated metab olites between sensitive MV4-11 and resistant ABT199-R. (<b>B</b>) Volcano plots of metabolic differentiation between MV4-11 and ABT199-R. (<b>C</b>) ABT199-R cells were treated with several concentrations of PKI-402 for 48 h and assessed with cell titer blue viability assay. The curve indicates the IC50 value.</p>
Full article ">Figure 3
<p>Metabolomic analysis of MV4-11, ABT199-R, and ABT199-R treated with PKI-402. (<b>A</b>) Metabolites were extracted and run in LTQ-XL linear ion trap LC-MS, showing their total ion chromatograms. (<b>B</b>) PCA analysis of comprehensive metabolites of MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>C</b>) Correlation heatmap of ABT-199R cells, MV4-11 cells, and ABT199-R cells treated with PKI-402. (<b>D</b>) Heatmap of differentially expressed metabolites in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>E</b>) VIP score based on PCA analysis of principal metabolites. (<b>F</b>) Top pathway enriched in metabolome analysis in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>G</b>) Pathway network analysis.</p>
Full article ">Figure 3 Cont.
<p>Metabolomic analysis of MV4-11, ABT199-R, and ABT199-R treated with PKI-402. (<b>A</b>) Metabolites were extracted and run in LTQ-XL linear ion trap LC-MS, showing their total ion chromatograms. (<b>B</b>) PCA analysis of comprehensive metabolites of MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>C</b>) Correlation heatmap of ABT-199R cells, MV4-11 cells, and ABT199-R cells treated with PKI-402. (<b>D</b>) Heatmap of differentially expressed metabolites in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>E</b>) VIP score based on PCA analysis of principal metabolites. (<b>F</b>) Top pathway enriched in metabolome analysis in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>G</b>) Pathway network analysis.</p>
Full article ">Figure 4
<p>p-AKT signaling pathway in ABT199-R cells. (<b>A</b>) Expression of PIP metabolites in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>B</b>) Western blot analysis of p-AKT and AKT in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>C</b>) Densitometer analysis of Western blot analysis of p-AKT/AKT in MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Venetoclax-resistant ABT199-R cells alter the glycolysis pathway. (<b>Left</b>) Quantitative levels of various metabolites involved in glycolysis pathways of MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>Right</b>) Gene expression (RT-PCR) of various genes involved in glycolysis. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>TCA cycle metabolism and OXPHOS are altered in ABT199-R cells. (<b>A</b>) The peak intensity of individual metabolites and their quantitative levels involved in the TCA cycle of MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>B</b>) Gene expression (RT-PCR) of various genes involved in mitochondrial OXPHOS. ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Pentose phosphate pathway (PPP) and energy metabolites modified in ABT199-R cells. (<b>A</b>) The density of metabolic features involved in regulating the energy metabolism of MV4-11 cells, ABT199-R cells, and ABT199-R cells treated with PKI-402. (<b>B</b>) Metabolites involved in PPP. ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>Redox homeostasis of venetoclax-resistant cells. (<b>A</b>) Glutathione (metabolite) involved in GSH pathway. (<b>B</b>) GLS-2 (gene) is involved in GSH pathway expression (RT-PCR). (<b>C</b>) Relative glutathione concentrations were measured using GSH assay from total protein. (<b>D</b>) ROS levels using CellRox Green Flow Cytometry assay. The channel Ch05 scans for red, and Ch02 scans for green. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 9
<p>Inhibition of PI3K-AKT induces apoptosis in resistant ABT199-R cells: ABT199-R cells were treated with the IC50 value of PKI-402 for 48 h and stained for an Annexin-V/PI assay. The single-cell images were captured using Amnis<sup>®</sup> FlowSight<sup>®</sup>. The percentage of apoptosis was calculated based on cell positivity. The channel Ch05 scans for red, and Ch02 scans for green. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
22 pages, 4812 KiB  
Article
Bcl-XL but Not Bcl-2 Is a Potential Target in Medulloblastoma Therapy
by Mike-Andrew Westhoff, Marie Schuler-Ortoli, Daniela Zerrinius, Amina Hadzalic, Andrea Schuster, Hannah Strobel, Angelika Scheuerle, Tiana Wong, Christian Rainer Wirtz, Klaus-Michael Debatin and Aurelia Peraud
Pharmaceuticals 2022, 15(1), 91; https://doi.org/10.3390/ph15010091 - 14 Jan 2022
Cited by 7 | Viewed by 2391
Abstract
Medulloblastoma (MB) is the most common solid tumour in children and, despite current treatment with a rather aggressive combination therapy, accounts for 10% of all deaths associated with paediatric cancer. Breaking the tumour cells’ intrinsic resistance to therapy-induced cell death should lead to [...] Read more.
Medulloblastoma (MB) is the most common solid tumour in children and, despite current treatment with a rather aggressive combination therapy, accounts for 10% of all deaths associated with paediatric cancer. Breaking the tumour cells’ intrinsic resistance to therapy-induced cell death should lead to less aggressive and more effective treatment options. In other tumour entities, this has been achieved by modulating the balance between the various pro- and anti-apoptotic members of the Bcl-2 family with small molecule inhibitors. To evaluate the therapeutic benefits of ABT-199 (Venetoclax), a Bcl-2 inhibitor, and ABT-263 (Navitoclax), a dual Bcl-XL/Bcl-2 inhibitor, increasingly more relevant model systems were investigated. Starting from established MB cell lines, progressing to primary patient-derived material and finally an experimental tumour system imbedded in an organic environment were chosen. Assessment of the metabolic activity (a surrogate readout for population viability), the induction of DNA fragmentation (apoptosis) and changes in cell number (the combined effect of alterations in proliferation and cell death induction) revealed that ABT-263, but not ABT-199, is a promising candidate for combination therapy, synergizing with cell death-inducing stimuli. Interestingly, in the experimental tumour setting, the sensitizing effect of ABT-263 seems to be predominantly mediated via an anti-proliferative and not a pro-apoptotic effect, opening a future line of investigation. Our data show that modulation of specific members of the Bcl-2 family might be a promising therapeutic addition for the treatment of MB. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Bcl-XL, but not Bcl-2, is a promising potential therapeutic target in medulloblastoma (MB) cell lines. (<b>a</b>) Summary of cell line characteristics. Abbreviations: SHH, sonic hedgehog; wt, wild type; amp, amplified; mut, mutant; m, male. (<b>b</b>) Immunoblotting of Bcl-2 family protein expression in MB cell lines. Extracts of four different MB cell lines were subjected to immunoblotting for Bcl-2 family proteins. β-actin served as loading control and the leukaemia cell line RS4;11 served as positive control for Bcl-2 and Mcl-1 expression. Anti-apoptotic members of the Bcl-2 family are marked by an asterisk. (<b>c</b>) The inhibitor of Bcl-2, ABT-199, does not show a therapeutic effect on MB cell lines. Cells were stimulated with 10 nM ABT-199 in the presence of either reduced serum concentrations (to 1.5% FCS for Daoy cells or 5% FCS for all others), or the chemotherapeutic drug doxorubicin (5 nM for Daoy cells or 10 nM for all others), or vincristine (1 nM for Daoy cells or 0.5 nM for all others) for either 24 (early) or 96 (late) h, as described in the Materials and Methods section. Left: Cell death was assessed by the surrogate readout of DNA fragmentation, evaluated via flow cytometric analysis of PI-stained nuclei. Right: The number of live cells was counted using the CASY1 DT cell counter. Using Bliss analysis, synergistic (green), additive (yellow) or antagonistic (red) effects were determined, while grey indicates that Bliss analysis was mathematically not possible to perform. Shown in (<b>b</b>) is a representative example of two independent experiments, while (<b>c</b>) summarizes at least three independent experiments performed in triplicate.</p>
Full article ">Figure 2
<p>The Bcl-XL/Bcl-2 inhibitor ABT-263 shows therapeutic potential in medulloblastoma (MB) cell lines. (<b>a</b>) Monotreatment with ABT-263 affects the metabolic activity of MB cell lines. The MB cell lines were seeded and stimulated with a serial dilution of ABT-263 for 24 and 96 h. The relative viability of the respective populations was normalized to the solvent-treated control population. The boxed column represents the concentration selected for further experiments (1 µM). (<b>b</b>) Combining ABT-263 and different stressors can enhance apoptosis. The seeded cell lines were cultured in the presence of 1 µM ABT-263, reduced serum concentrations (to 1.5% FCS for Daoy cells or 5% FCS for all others), doxorubicin (5 nM for Daoy cells or 10 nM for all others) and vincristine (1 nM for Daoy cells or 0.5 nM for all others), as well as combinations of them (concentrations described in the Materials and Methods section). Percentage of specific DNA fragmentation was determined 24 and 96 h after stimulation by flow cytometric analysis of PI-stained nuclei. (<b>c</b>) Combining ABT-263 and different stressors can reduce living cell numbers. The experimental setup was similar to (<b>b</b>). This was followed by assessing the total living cell number 24 and 96 h after stimulation by CASY1 DT measurement. Data was normalized to a solvent-treated control population for each time point. (<b>d</b>) The inhibitor of Bcl-XL/Bcl-2, ABT-263, shows a therapeutic effect on MB cell lines. Similarly to the data shown in <a href="#pharmaceuticals-15-00091-f001" class="html-fig">Figure 1</a>b, the effects of combining ABT-263 with different stressors on apoptosis (left) and living cell number (right), as shown in (<b>b</b>,<b>c</b>), respectively (early, 24 h; late, 96 h), were analysed for synergism. Using Bliss analysis, synergistic (green), additive (yellow) or antagonistic (red) effects were determined, while grey indicates that Bliss analysis was mathematically not possible to perform. Shown in (<b>a</b>) are the mean and SD of at least three independent experiments with six individual data points. In (<b>b</b>) are the mean and SD of three independent experiments performed in triplicates, and in (<b>c</b>) are the mean and SD of at least four independent experiments performed in triplicates, while (<b>d</b>) is a mathematical summary of (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 2 Cont.
<p>The Bcl-XL/Bcl-2 inhibitor ABT-263 shows therapeutic potential in medulloblastoma (MB) cell lines. (<b>a</b>) Monotreatment with ABT-263 affects the metabolic activity of MB cell lines. The MB cell lines were seeded and stimulated with a serial dilution of ABT-263 for 24 and 96 h. The relative viability of the respective populations was normalized to the solvent-treated control population. The boxed column represents the concentration selected for further experiments (1 µM). (<b>b</b>) Combining ABT-263 and different stressors can enhance apoptosis. The seeded cell lines were cultured in the presence of 1 µM ABT-263, reduced serum concentrations (to 1.5% FCS for Daoy cells or 5% FCS for all others), doxorubicin (5 nM for Daoy cells or 10 nM for all others) and vincristine (1 nM for Daoy cells or 0.5 nM for all others), as well as combinations of them (concentrations described in the Materials and Methods section). Percentage of specific DNA fragmentation was determined 24 and 96 h after stimulation by flow cytometric analysis of PI-stained nuclei. (<b>c</b>) Combining ABT-263 and different stressors can reduce living cell numbers. The experimental setup was similar to (<b>b</b>). This was followed by assessing the total living cell number 24 and 96 h after stimulation by CASY1 DT measurement. Data was normalized to a solvent-treated control population for each time point. (<b>d</b>) The inhibitor of Bcl-XL/Bcl-2, ABT-263, shows a therapeutic effect on MB cell lines. Similarly to the data shown in <a href="#pharmaceuticals-15-00091-f001" class="html-fig">Figure 1</a>b, the effects of combining ABT-263 with different stressors on apoptosis (left) and living cell number (right), as shown in (<b>b</b>,<b>c</b>), respectively (early, 24 h; late, 96 h), were analysed for synergism. Using Bliss analysis, synergistic (green), additive (yellow) or antagonistic (red) effects were determined, while grey indicates that Bliss analysis was mathematically not possible to perform. Shown in (<b>a</b>) are the mean and SD of at least three independent experiments with six individual data points. In (<b>b</b>) are the mean and SD of three independent experiments performed in triplicates, and in (<b>c</b>) are the mean and SD of at least four independent experiments performed in triplicates, while (<b>d</b>) is a mathematical summary of (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 3
<p>Characterisation of patient-derived medulloblastoma (MB) stem-like cells and short-term differentiated cells. (<b>a</b>) Schematic workflow of obtaining patient-derived MB cells. During therapeutically indicated surgery and following patient’s (or legal guardian’s) consent, a tumour sample surplus to histological characterisation and archiving was split into two. While one element was stored as reference material, the remaining tissue was further dissected and cleaned from blood cells. After several washes, fragments were sieved, and single cells/small cell clumps were cultured under non-adherent conditions in “stem cell” medium. Neurospheres that emerged after prolonged culturing were stored and characterised with reference to original tumour histology. Adherent, differentiated progeny were cultured from neurospheres via permitted adhesion and change of culture conditions, using “differentiation” medium. These differentiated populations were used for a maximum of ten generations to ensure genetic stability. Details are presented in the Materials and Methods section. Shown here are also exemplary pictures of the stem cell-like MB cells grown as free-flowing spheres and their adherently cultured differentiated progeny. Scale bar: 50 μm. (<b>b</b>) Neuropathology specimen of the patient’s tumour stained as indicated. The H&amp;E staining shows small, round and densely packed cells. GFAP staining is negative as for most MB tumours, while a relatively high count of Ki-67 positive areas indicates an increased amount of cycling cells. MAP2 confirms the CNS origin of the tumour, while ß-catenin indicates that the tumour exhibits SHH activity. In addition, Alcian and PAS was used to detect polysaccharides, elevated in cancers. Taken together, these data suggest a medulloblastoma belonging to the SHH subgroup. Abbreviations: H&amp;E—haematoxylin and eosin stain; GFAP—Glial Fibrillary Acidic Protein; Ki-67—Antigen KI-67; MAP2—Microtubule-Associated Protein 2; ß-catenin—βeta-catenin; Alcian &amp; PAS—Alcian blue stain and Perodic acid–Schiff stain, SHH–sonic hedgehog. Scale bar: 200 µm. (<b>c</b>) Expression protein profile in MB stem cell-like cells and differentiated cells shown as a heat map. Protein extracts of a stem cell-like cell population (SC) and its differentiated progeny (DC) were subjected to immunoblotting for three different types of protein profile assays, as described in the Materials and Methods section. Following visualisation, membranes were densitometrically evaluated and bioinformatically processed. The map shows relative expression (blue, higher than average; red, lower than average; as described in the Materials and Methods section) and represents the mean of two independent experiments.</p>
Full article ">Figure 3 Cont.
<p>Characterisation of patient-derived medulloblastoma (MB) stem-like cells and short-term differentiated cells. (<b>a</b>) Schematic workflow of obtaining patient-derived MB cells. During therapeutically indicated surgery and following patient’s (or legal guardian’s) consent, a tumour sample surplus to histological characterisation and archiving was split into two. While one element was stored as reference material, the remaining tissue was further dissected and cleaned from blood cells. After several washes, fragments were sieved, and single cells/small cell clumps were cultured under non-adherent conditions in “stem cell” medium. Neurospheres that emerged after prolonged culturing were stored and characterised with reference to original tumour histology. Adherent, differentiated progeny were cultured from neurospheres via permitted adhesion and change of culture conditions, using “differentiation” medium. These differentiated populations were used for a maximum of ten generations to ensure genetic stability. Details are presented in the Materials and Methods section. Shown here are also exemplary pictures of the stem cell-like MB cells grown as free-flowing spheres and their adherently cultured differentiated progeny. Scale bar: 50 μm. (<b>b</b>) Neuropathology specimen of the patient’s tumour stained as indicated. The H&amp;E staining shows small, round and densely packed cells. GFAP staining is negative as for most MB tumours, while a relatively high count of Ki-67 positive areas indicates an increased amount of cycling cells. MAP2 confirms the CNS origin of the tumour, while ß-catenin indicates that the tumour exhibits SHH activity. In addition, Alcian and PAS was used to detect polysaccharides, elevated in cancers. Taken together, these data suggest a medulloblastoma belonging to the SHH subgroup. Abbreviations: H&amp;E—haematoxylin and eosin stain; GFAP—Glial Fibrillary Acidic Protein; Ki-67—Antigen KI-67; MAP2—Microtubule-Associated Protein 2; ß-catenin—βeta-catenin; Alcian &amp; PAS—Alcian blue stain and Perodic acid–Schiff stain, SHH–sonic hedgehog. Scale bar: 200 µm. (<b>c</b>) Expression protein profile in MB stem cell-like cells and differentiated cells shown as a heat map. Protein extracts of a stem cell-like cell population (SC) and its differentiated progeny (DC) were subjected to immunoblotting for three different types of protein profile assays, as described in the Materials and Methods section. Following visualisation, membranes were densitometrically evaluated and bioinformatically processed. The map shows relative expression (blue, higher than average; red, lower than average; as described in the Materials and Methods section) and represents the mean of two independent experiments.</p>
Full article ">Figure 4
<p>The Bcl-XL/Bcl-2 inhibitor ABT-263, but not the Bcl-2 inhibitor ABT-199, shows therapeutic potential in primary medulloblastoma (MB) cells. (<b>a</b>) The inhibitor of Bcl-2, ABT-199, does not show a therapeutic effect on the metabolism of primary MB cells. The stem cell-like cells (SC322, left) or their differentiated progeny (DC322, right) were cultured in the presence of ABT-199, reduced growth factor/serum concentrations, doxorubicin or two concentrations of vincristine as well as ABT-199 combined with the different stressors (concentrations as indicated). Metabolic activity was assessed after 24 and 96 h and normalized to the solvent-treated control population. (<b>b</b>) The inhibitor of Bcl-XL/Bcl-2, ABT-263, shows a therapeutic effect on the metabolism of primary MB cells. The experimental setup was similar to 4A. Metabolic activity was assessed after 24 and 96 h and normalized to the solvent-treated control population. (<b>c</b>) ABT-263 shows a therapeutic effect on cell death induction of primary MB cells. The stem cell-like cells (SC322, left) or their differentiated progeny (DC322, right) were cultured in the presence of 1 µM ABT-263, reduced growth factor/serum concentrations (as indicated), 5 nM doxorubicin or 0.5 nM vincristine as well as ABT-263 combined with the different stressors. Percentage of specific DNA fragmentation was determined 24 and 96 h after stimulation by flow cytometric analysis of PI-stained nuclei. (<b>d</b>) ABT-263 shows a therapeutic effect on living cell number of primary MB cells. The experimental setup was similar to 4C. The total living cell number was assessed 24 and 96 h after stimulation by CASY1 DT measurement and normalized to solvent-treated control population for each time point. (<b>e</b>) Combining ABT-263 and various stressors frequently elicits a synergistic effect in primary MB cells. The effects of combining ABT-263 with different stressors on apoptosis (left) and living cell number (right), as shown in (<b>c</b>) and (<b>d</b>), respectively (early, 24 h; late, 96 h), were analysed for synergism. Using Bliss analysis, synergistic (green), additive (yellow) or antagonistic (red) effects were determined, while grey indicates that Bliss analysis was mathematically not possible to perform. In (<b>a</b>–<b>d</b>) columns represent mean and SD of three independent experiments performed in triplicates.</p>
Full article ">Figure 4 Cont.
<p>The Bcl-XL/Bcl-2 inhibitor ABT-263, but not the Bcl-2 inhibitor ABT-199, shows therapeutic potential in primary medulloblastoma (MB) cells. (<b>a</b>) The inhibitor of Bcl-2, ABT-199, does not show a therapeutic effect on the metabolism of primary MB cells. The stem cell-like cells (SC322, left) or their differentiated progeny (DC322, right) were cultured in the presence of ABT-199, reduced growth factor/serum concentrations, doxorubicin or two concentrations of vincristine as well as ABT-199 combined with the different stressors (concentrations as indicated). Metabolic activity was assessed after 24 and 96 h and normalized to the solvent-treated control population. (<b>b</b>) The inhibitor of Bcl-XL/Bcl-2, ABT-263, shows a therapeutic effect on the metabolism of primary MB cells. The experimental setup was similar to 4A. Metabolic activity was assessed after 24 and 96 h and normalized to the solvent-treated control population. (<b>c</b>) ABT-263 shows a therapeutic effect on cell death induction of primary MB cells. The stem cell-like cells (SC322, left) or their differentiated progeny (DC322, right) were cultured in the presence of 1 µM ABT-263, reduced growth factor/serum concentrations (as indicated), 5 nM doxorubicin or 0.5 nM vincristine as well as ABT-263 combined with the different stressors. Percentage of specific DNA fragmentation was determined 24 and 96 h after stimulation by flow cytometric analysis of PI-stained nuclei. (<b>d</b>) ABT-263 shows a therapeutic effect on living cell number of primary MB cells. The experimental setup was similar to 4C. The total living cell number was assessed 24 and 96 h after stimulation by CASY1 DT measurement and normalized to solvent-treated control population for each time point. (<b>e</b>) Combining ABT-263 and various stressors frequently elicits a synergistic effect in primary MB cells. The effects of combining ABT-263 with different stressors on apoptosis (left) and living cell number (right), as shown in (<b>c</b>) and (<b>d</b>), respectively (early, 24 h; late, 96 h), were analysed for synergism. Using Bliss analysis, synergistic (green), additive (yellow) or antagonistic (red) effects were determined, while grey indicates that Bliss analysis was mathematically not possible to perform. In (<b>a</b>–<b>d</b>) columns represent mean and SD of three independent experiments performed in triplicates.</p>
Full article ">Figure 4 Cont.
<p>The Bcl-XL/Bcl-2 inhibitor ABT-263, but not the Bcl-2 inhibitor ABT-199, shows therapeutic potential in primary medulloblastoma (MB) cells. (<b>a</b>) The inhibitor of Bcl-2, ABT-199, does not show a therapeutic effect on the metabolism of primary MB cells. The stem cell-like cells (SC322, left) or their differentiated progeny (DC322, right) were cultured in the presence of ABT-199, reduced growth factor/serum concentrations, doxorubicin or two concentrations of vincristine as well as ABT-199 combined with the different stressors (concentrations as indicated). Metabolic activity was assessed after 24 and 96 h and normalized to the solvent-treated control population. (<b>b</b>) The inhibitor of Bcl-XL/Bcl-2, ABT-263, shows a therapeutic effect on the metabolism of primary MB cells. The experimental setup was similar to 4A. Metabolic activity was assessed after 24 and 96 h and normalized to the solvent-treated control population. (<b>c</b>) ABT-263 shows a therapeutic effect on cell death induction of primary MB cells. The stem cell-like cells (SC322, left) or their differentiated progeny (DC322, right) were cultured in the presence of 1 µM ABT-263, reduced growth factor/serum concentrations (as indicated), 5 nM doxorubicin or 0.5 nM vincristine as well as ABT-263 combined with the different stressors. Percentage of specific DNA fragmentation was determined 24 and 96 h after stimulation by flow cytometric analysis of PI-stained nuclei. (<b>d</b>) ABT-263 shows a therapeutic effect on living cell number of primary MB cells. The experimental setup was similar to 4C. The total living cell number was assessed 24 and 96 h after stimulation by CASY1 DT measurement and normalized to solvent-treated control population for each time point. (<b>e</b>) Combining ABT-263 and various stressors frequently elicits a synergistic effect in primary MB cells. The effects of combining ABT-263 with different stressors on apoptosis (left) and living cell number (right), as shown in (<b>c</b>) and (<b>d</b>), respectively (early, 24 h; late, 96 h), were analysed for synergism. Using Bliss analysis, synergistic (green), additive (yellow) or antagonistic (red) effects were determined, while grey indicates that Bliss analysis was mathematically not possible to perform. In (<b>a</b>–<b>d</b>) columns represent mean and SD of three independent experiments performed in triplicates.</p>
Full article ">Figure 5
<p>Combining the Bcl-XL/Bcl-2 inhibitor ABT-263 and vincristine shows superior therapeutic potential over single treatment in a chorioallantoic membrane (CAM) assay. (<b>a</b>) The surface area of double treated tumours appears to be reduced compared to mock or single treated entities. A 20:80 mixture of medulloblastoma stem cell-like cells and their differentiated progeny (SC/DC322) was seeded on the CAM of fertilized chicken eggs and were treated daily for four consecutive days with the indicated substances via local surface application (ABT-263: 5 µM, vincristine: 2.5 nM). After the aforementioned time span, the tumour surface area of all tumours (as shown on the right) was determined as indicated (bottom left). While a trend towards reduction is clearly visible, the reduction of surface area was statistically not significant, as determined by a one-way ANOVA test (<span class="html-italic">p</span>-value 0.0623). (<b>b</b>) Combining ABT-263 and vincristine has an effect on tumour morphology and cycling cells, but not necessarily on apoptosis induction. Shown are representative sections taken from the tumour centre of an individual tumour per treatment group, stained for haematoxylin and eosin (H&amp;E), haematoxylin and Ki-67 (Ki-67) or haematoxylin and cleaved Caspase-3 (cCaspase-3). Scale: bars indicate 1000 µm, except in the columns labelled “(5x)”; here, bars indicate 200 µm. (<b>c</b>) Combining ABT-263 and vincristine increases the fraction of non-cycling cells. A histological evaluation of cells progressing through the cell cycle (Ki-67 positive) was performed. + indicates some positive cells clearly detectable, ++ indicates the majority of cells are positive, +++ almost all cells are positive; n/a, not available. Fields marked in yellow indicate the tumours depicted in (<b>b</b>). (<b>d</b>) High concentrations of ABT-263 can induce apoptosis. Depicted is a section of a tumour treated with ABT-263 only. Clearly visible is the positive staining for cleaved Caspase-3 (cCaspase-3) in the top layers of the section. These are the layers to which ABT-263 was directly applied and then allowed to diffuse through the tumour. Scale bar indicates 200 µm.</p>
Full article ">Figure 5 Cont.
<p>Combining the Bcl-XL/Bcl-2 inhibitor ABT-263 and vincristine shows superior therapeutic potential over single treatment in a chorioallantoic membrane (CAM) assay. (<b>a</b>) The surface area of double treated tumours appears to be reduced compared to mock or single treated entities. A 20:80 mixture of medulloblastoma stem cell-like cells and their differentiated progeny (SC/DC322) was seeded on the CAM of fertilized chicken eggs and were treated daily for four consecutive days with the indicated substances via local surface application (ABT-263: 5 µM, vincristine: 2.5 nM). After the aforementioned time span, the tumour surface area of all tumours (as shown on the right) was determined as indicated (bottom left). While a trend towards reduction is clearly visible, the reduction of surface area was statistically not significant, as determined by a one-way ANOVA test (<span class="html-italic">p</span>-value 0.0623). (<b>b</b>) Combining ABT-263 and vincristine has an effect on tumour morphology and cycling cells, but not necessarily on apoptosis induction. Shown are representative sections taken from the tumour centre of an individual tumour per treatment group, stained for haematoxylin and eosin (H&amp;E), haematoxylin and Ki-67 (Ki-67) or haematoxylin and cleaved Caspase-3 (cCaspase-3). Scale: bars indicate 1000 µm, except in the columns labelled “(5x)”; here, bars indicate 200 µm. (<b>c</b>) Combining ABT-263 and vincristine increases the fraction of non-cycling cells. A histological evaluation of cells progressing through the cell cycle (Ki-67 positive) was performed. + indicates some positive cells clearly detectable, ++ indicates the majority of cells are positive, +++ almost all cells are positive; n/a, not available. Fields marked in yellow indicate the tumours depicted in (<b>b</b>). (<b>d</b>) High concentrations of ABT-263 can induce apoptosis. Depicted is a section of a tumour treated with ABT-263 only. Clearly visible is the positive staining for cleaved Caspase-3 (cCaspase-3) in the top layers of the section. These are the layers to which ABT-263 was directly applied and then allowed to diffuse through the tumour. Scale bar indicates 200 µm.</p>
Full article ">
14 pages, 2990 KiB  
Article
A Novel Regimen for Treating Melanoma: MCL1 Inhibitors and Azacitidine
by Chiara R. Dart, Nabanita Mukherjee, Carol M. Amato, Anabel Goulding, Morgan MacBeth, Robert Van Gulick, Kasey L. Couts, James R. Lambert, David A. Norris, William A. Robinson and Yiqun G. Shellman
Pharmaceuticals 2021, 14(8), 749; https://doi.org/10.3390/ph14080749 - 30 Jul 2021
Cited by 3 | Viewed by 2728
Abstract
Although treatment options for melanoma patients have expanded in recent years with the approval of immunotherapy and targeted therapy, there is still an unmet need for new treatment options for patients that are ineligible for, or resistant to these therapies. BH3 mimetics, drugs [...] Read more.
Although treatment options for melanoma patients have expanded in recent years with the approval of immunotherapy and targeted therapy, there is still an unmet need for new treatment options for patients that are ineligible for, or resistant to these therapies. BH3 mimetics, drugs that mimic the activity of pro-apoptotic BCL2 family proteins, have recently achieved remarkable success in the clinical setting. The combination of BH3 mimetic ABT-199 (venetoclax) plus azacitidine has shown substantial benefit in treating acute myelogenous leukemia. We evaluated the efficacy of various combinations of BH3 mimetic + azacitidine in fourteen human melanoma cell lines from cutaneous, mucosal, acral and uveal subtypes. Using a combination of cell viability assay, BCL2 family knockdown cell lines, live cell imaging, and sphere formation assay, we found that combining inhibition of MCL1, an anti-apoptotic BCL2 protein, with azacitidine had substantial pro-apoptotic effects in multiple melanoma cell lines. Specifically, this combination reduced cell viability, proliferation, sphere formation, and induced apoptosis. In addition, this combination is highly effective at reducing cell viability in rare mucosal and uveal subtypes. Overall, our data suggest this combination as a promising therapeutic option for some patients with melanoma and should be further explored in clinical trials. Full article
(This article belongs to the Special Issue Novel Therapeutic Targets in Cancer)
Show Figures

Figure 1

Figure 1
<p>Single agent or combination treatment with ABT-199 and azacitidine (AZA) is ineffective in melanoma cell lines. (<b>A</b>,<b>B</b>) ATP assay of melanoma cell lines after 48 h of treatment with single agent AZA or ABT-199 (<b>A</b>), or their combination (<b>B</b>). Combination treatment was treated at 2.5 uM. Black dotted line indicates viability of 50%. Y axis indicates viability relative to DMSO control, set to 100%. X axis indicates drug treatment. Error bars represent +/− SEM. (<b>C</b>) Plot of the combination index (CI) values for the ABT-199 + AZA combination at 2.5 uM concentration. CI values were calculated using the CompuSyn software (version 1). CI values &gt;1 indicate antagonism, values 0.9–1 indicate an additive effect, and values &lt;0.9 indicate synergy. For clarity, values &gt;5 are cutoff at 5. Black dotted line indicates CI value of 1. Y axis indicates CI value, X axis indicates cell line.</p>
Full article ">Figure 2
<p>Knockdown of MCL1 or BCLXL sensitizes cells to single agent AZA treatment; knockdown of BCL2 has no effect on AZA sensitivity. (<b>A</b>) Knockdown A375 cell lines of BCL2, BCLXL, and MCL1, (<b>B</b>) Knockdown SKMEL-28 cell line of MCL1. (<b>A</b>,<b>B</b>) Knockdown cell lines were treated with AZA 0 uM- 2.5 uM for 48 h. Knockdown of MCL1 significantly sensitized cells to treatment with single agent AZA. Arrow indicates significance compared to sh Control. * indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01. Y axis indicates viability relative to DMSO control, set to 100%. X axis indicates AZA dosage. Immunoblot with control and knockdown cell lysates indicate knockdown of the target proteins. Error bars represent +/− SEM.</p>
Full article ">Figure 3
<p>Combination treatment with S63845 + AZA is effective in synergistically reducing melanoma cell viability. (<b>A</b>) Single agent S63845 and ABT-263 dosed at a concentrations of 0.156–2.5 uM in human melanoma cell lines. (<b>B</b>) Combination treatment with S63845 + AZA, and ABT-263 + AZA dosed at 2.5 uM each in human melanoma cell lines. (<b>A</b>,<b>B</b>) All cells were treated for 48 h. Dotted line indicates 50% viability. Y axis indicates viability relative to DMSO control, set to 100%. X axis indicates drug treatment. Error bars represent +/− SEM. (<b>C</b>) CI values for S63845 + AZA treatment. CI &lt;0.9 indicates synergy, 0.9–1 indicates additivity, &gt;1 indicates antagonism. CI values are calculated at the 2.5 uM dose using CompuSyn (version 1) software. Black dotted line indicates CI value of 1. Y axis indicates CI value, X axis indicates cell line and treatment.</p>
Full article ">Figure 4
<p>S63845 and clinical grade version S64315 (MIK665) have similar efficacy in melanoma cell lines. (<b>A</b>) ATP assay of S63845 or S64315 plus AZA in SKMEL-28. (<b>B</b>) ATP assay of S63845 or S64315 plus AZA in MB4667. Dose of AZA indicated in title of graph. Y axis indicates viability relative to DMSO control, set to 100%. X axis indicates S63845/S64315 dosage. Error bars represent +/− SEM.</p>
Full article ">Figure 5
<p>Treatment with S63845/S64315 plus AZA decreases proliferation and induces apoptosis. (<b>A</b>) IncuCyte live cell analysis of caspase 3/7 activity. Y axis indicates ratio of area expressing fluorescent signal. X axis indicates time in hours. (<b>B</b>) IncuCyte live cell analysis of cellular proliferation. Y axis represents confluence relative to 0 h. X axis indicates time in hours. For all graphs, error bars represent +/− SEM. Arrow indicates lowest significance compared to all other conditions. ** indicates <span class="html-italic">p</span> &lt; 0.01, *** indicates <span class="html-italic">p</span> &lt; 0.001 **** indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>S63845 plus AZA treatment disrupts sphere formation. (<b>A</b>) Quantification of primary spheres after 48 h of indicated treatments. Y axis represents sphere number relative to DMSO control, set to 100%. X axis indicates treatment. Error bars represent +/− SEM. (<b>B</b>) Brightfield images of representative wells of MB3616 spheres after 48 h of indicated treatment. Scalebar represents 100 um.</p>
Full article ">Figure 7
<p>Mucosal and uveal melanoma are most sensitive to MCL1i + AZA. Dot plot of IC50 values for the combination of S63845 + AZA, separated by melanoma subtype. Each dot represents one cell line. * represents <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
20 pages, 4721 KiB  
Article
Senolytic Targeting of Bcl-2 Anti-Apoptotic Family Increases Cell Death in Irradiated Sarcoma Cells
by Julie Lafontaine, Guillaume B. Cardin, Nicolas Malaquin, Jean-Sébastien Boisvert, Francis Rodier and Philip Wong
Cancers 2021, 13(3), 386; https://doi.org/10.3390/cancers13030386 - 21 Jan 2021
Cited by 28 | Viewed by 4516
Abstract
Radiotherapy (RT) is a key component of cancer treatment. Most of the time, radiation is given after surgery but for soft-tissue sarcomas (STS), pre-surgical radiation is commonly utilized. However, despite improvements in RT accuracy, the rate of local recurrence remains high and is [...] Read more.
Radiotherapy (RT) is a key component of cancer treatment. Most of the time, radiation is given after surgery but for soft-tissue sarcomas (STS), pre-surgical radiation is commonly utilized. However, despite improvements in RT accuracy, the rate of local recurrence remains high and is the major cause of death for patients with STS. A better understanding of cell fates in response to RT could provide new therapeutic options to enhance tumour cell killing by RT and facilitate surgical resection. Here, we showed that irradiated STS cell cultures do not die but instead undergo therapy-induced senescence (TIS), which is characterized by proliferation arrest, senescence-associated β-galactosidase activity, secretion of inflammatory cytokines and persistent DNA damage. STS-TIS was also associated with increased levels of the anti-apoptotic Bcl-2 family of proteins which rendered cells targetable using senolytic Bcl-2 inhibitors. As oppose to radiation alone, the addition of senolytic agents Venetoclax (ABT-199) or Navitoclax (ABT-263) after irradiation induced a rapid apoptotic cell death in STS monolayer cultures and in a more complex three-dimensional culture model. Together, these data suggest a new promising therapeutic approach for sarcoma patients who receive neoadjuvant RT. The addition of senolytic agents to radiation treatments may significantly reduce tumour volume prior to surgery and thereby improve the clinical outcome of patients. Full article
(This article belongs to the Special Issue Sarcomas: New Biomarkers and Therapeutic Strategies)
Show Figures

Figure 1

Figure 1
<p>Radiation induces cytostatic effects in sarcoma cell lines. (<b>A</b>) Representative picture of the three undifferentiated pleomorphic sarcoma (UPS) cell lines, untreated or 5 days after treatment (8 Gy). (<b>B</b>) Clonogenic survival assay of STS93 and STS117 treated with 0, 0.5, 2, 4 and 8 Gy of radiation. Cell survival is normalized to the clonogenic formation from untreated cells (0 Gy). (<b>C</b>,<b>D</b>) Cell death and cell cycle analyzed by flow cytometry 48 h after exposure to radiation (0, 2 and 10 Gy). In the graph, percentage of cell death (<b>C</b>) represents the sum of Annexin V positive cells (both DAPI positives and negatives) and DAPI positive cells (Annexin V negatives) from the quadrant plots of DAPI vs. Annexin V. (<b>E</b>) Cell proliferation curves of sarcoma cell lines expressing H2B-GFP and exposed to increasing doses of radiation (0, 2, 4, 6 and 8 Gy). Student’s t-test, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Data are representative of two to three experiments. Error bars represent ± standard deviation.</p>
Full article ">Figure 2
<p>Radiation induces a senescence-like phenotype in sarcoma cell lines. (<b>A</b>) SA-β-gal staining 10 days following irradiation (RT) with doses of 2, 4 and 10 Gy. (<b>B</b>) Analysis of a 24 h EdU pulse labelling 5 days following radiation (2, 4, 6 or 8 Gy). (<b>C</b>) DNA damage associated γH2AX (green) and 53BP1 (red) immunofluorescence 10 days after exposure to 8 Gy. (<b>D</b>) mRNA levels of IL-6 and IL-8 over time (D stands for day) relative to untreated controls evaluated by real-time qPCR following 8 Gy of radiation. (<b>E</b>) Secretion (pg/mL/10<sup>6</sup> cells) of IL-6 and IL-8 measured after 10 days (8 Gy). Not Detected (ND) indicated values below the standard curve. Data in (<b>B</b>) and (<b>D</b>) were analyzed using a two-tail Student’s t-test to compare RT treatment with the untreated control. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Data are representative of two to three experiments. Error bars indicate ± the standard deviation.</p>
Full article ">Figure 3
<p>Anti-apoptotic proteins of the Bcl-2 family are upregulated by radiation. (<b>A</b>) Relative mRNA levels of BCL-2 and BCL-XL over time evaluated by real-time qPCR following 8 Gy of radiation. The values represent fold change expression relative to untreated controls. (<b>B</b>) Western blot analysis of BCL-2 and BCL-XL protein levels of untreated control (-) at Day 5 and irradiated (RT) cells over time. Stain free is a representative band of total protein acquired by stain-free technology. Bar graph (in blue) represent the protein quantification relative to the total protein. Data in (<b>A</b>) were analyzed using two-tail Student’s t-test to compare untreated vs. RT treated cells. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001. Data representative of three experiments for qPCR and one experiment for western blot. Error bars indicate ± the standard deviation.</p>
Full article ">Figure 4
<p>ABT-263 or ABT-199 induces rapid and specific cell death in irradiated sarcoma cells. (<b>A</b>) Real-time cell death curve of propidium iodide (PI) incorporation in untreated (left) or pre-irradiated (RT; 8 Gy, 5 days before) (right) sarcoma cell lines treated with ABT-263 (0.5 μM) or ABT-199 (5 μM). (<b>B</b>) Representative images of PI staining in irradiated cells treated with vehicle, ABT-263 or ABT-199. (<b>C</b>) Caspase 3-7 activity in pre-irradiated cells (8 Gy) treated with ABT-263 (0.25 and 0.5 μM) or ABT-199 (2.5 and 5 μM). (<b>D</b>) Flow cytometry analysis of apoptosis in untreated cells, cells treated with ABT-199 (10 μM) or ABT-263 (1 μM) or RT alone (8 Gy) (monotherapy), and pre-irradiated cells treated with either senolytic in combination. (<b>E</b>) Cytotoxicity evaluated by percentage of PI positive cells over total number of cells (Cell viability (%)) in control or pre-irradiated cells (8 Gy) 48 h after treatment with increasing doses of ABT-199 or ABT-263. (<b>F</b>) Doses response curves for ATB-199 and ABT-263 treatment of RT (8 Gy) or untreated cells for in each cell lines evaluated from % of viability relative to DMSO exposed cells. (<b>G</b>) IC50 values for both compounds. (<b>H</b>) Heat map of Excess over bliss for combination treatments of radiation and different concentrations of ABT-263 and ABT-199. Data in (<b>E</b>) were analyzed using ANOVA for multiple comparison with the vehicle treated control. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Data are representative of three experiments and the Bliss score data are the mean of three independent experiments. Error bars indicate ± the standard deviation.</p>
Full article ">Figure 5
<p>Senolytics can induced cell death in irradiated 3D model of sarcoma cells. (<b>A</b>) Fold change in size over time of spheroids untreated or exposed to 8 Gy. (<b>B</b>) Size of spheroids irradiated (RT; 8 Gy) or combination of RT and ABT-263 (0.25 μM and 0.5 μM) or ABT-199 (5 μM and 10 μM). Measurements represent size 96 h after treatment. (<b>C</b>) Representative pictures of PI incorporation in irradiated spheroids treated with combination treatment over time. T = 0 represents cells prior to drug addition. The scale bar represents 300 μM. Data from (<b>A</b>) and (<b>B</b>) were analyzed using the two-tail Student’s t-test to compare treatment groups with the control group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Data are representative of three experiments. Error bars indicate ± the standard deviation.</p>
Full article ">Figure 6
<p>Potential clinical strategy for curative STS. Model in three steps proposed for management of treatment timeline of patient. For sarcomas patient, pre-operative radiation is given in several fractions and is followed by an interval of six to ten weeks before surgery. During this period of time, we hypothesize that the tumour will undergo radiation-induced senescence and this state will sensitize the cells for the use of senolytic drugs. This addition to conventional treatment may enhance tumour shrinkage and thus, potentially have positive effects on the success of the surgery as well as on reducing side effects cause by radiation over time.</p>
Full article ">
19 pages, 5217 KiB  
Article
Implementing Systems Modelling and Molecular Imaging to Predict the Efficacy of BCL-2 Inhibition in Colorectal Cancer Patient-Derived Xenograft Models
by Alice C. O’Farrell, Monika A. Jarzabek, Andreas U. Lindner, Steven Carberry, Emer Conroy, Ian S. Miller, Kate Connor, Liam Shiels, Eugenia R. Zanella, Federico Lucantoni, Adam Lafferty, Kieron White, Mariangela Meyer Villamandos, Patrick Dicker, William M. Gallagher, Simon A. Keek, Sebastian Sanduleanu, Philippe Lambin, Henry C. Woodruff, Andrea Bertotti, Livio Trusolino, Annette T. Byrne and Jochen H. M. Prehnadd Show full author list remove Hide full author list
Cancers 2020, 12(10), 2978; https://doi.org/10.3390/cancers12102978 - 14 Oct 2020
Cited by 10 | Viewed by 3768
Abstract
Resistance to chemotherapy often results from dysfunctional apoptosis, however multiple proteins with overlapping functions regulate this pathway. We sought to determine whether an extensively validated, deterministic apoptosis systems model, ‘DR_MOMP’, could be used as a stratification tool for the apoptosis sensitiser and BCL-2 [...] Read more.
Resistance to chemotherapy often results from dysfunctional apoptosis, however multiple proteins with overlapping functions regulate this pathway. We sought to determine whether an extensively validated, deterministic apoptosis systems model, ‘DR_MOMP’, could be used as a stratification tool for the apoptosis sensitiser and BCL-2 antagonist, ABT-199 in patient-derived xenograft (PDX) models of colorectal cancer (CRC). Through quantitative profiling of BCL-2 family proteins, we identified two PDX models which were predicted by DR_MOMP to be sufficiently sensitive to 5-fluorouracil (5-FU)-based chemotherapy (CRC0344), or less responsive to chemotherapy but sensitised by ABT-199 (CRC0076). Treatment with ABT-199 significantly improved responses of CRC0076 PDXs to 5-FU-based chemotherapy, but showed no sensitisation in CRC0344 PDXs, as predicted from systems modelling. 18F-Fluorodeoxyglucose positron emission tomography/computed tomography (18F-FDG-PET/CT) scans were performed to investigate possible early biomarkers of response. In CRC0076, a significant post-treatment decrease in mean standard uptake value was indeed evident only in the combination treatment group. Radiomic CT feature analysis of pre-treatment images in CRC0076 and CRC0344 PDXs identified features which could phenotypically discriminate between models, but were not predictive of treatment responses. Collectively our data indicate that systems modelling may identify metastatic (m)CRC patients benefitting from ABT-199, and that 18F-FDG-PET could independently support such predictions. Full article
(This article belongs to the Special Issue Patient-Derived Xenograft-Models in Cancer Research)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the experimental workflow. The sensitivity of tumour cells to undergo apoptosis was calculated after quantitative protein profiling using the deterministic systems model DR_MOMP [<a href="#B7-cancers-12-02978" class="html-bibr">7</a>]. First, protein levels were calculated by quantitative Western blotting in proteins extracted from colorectal cancer (CRC) patient-derived xenograft (PDX) tumours, using cell lines as standards in which absolute BCL-2 protein levels were previously determined. Protein profiles then served as input into the deterministic systems model, DR_MOMP. The model calculates a ‘stress dose’ (pro-apoptotic ‘BH3 only protein dose’) required to induce mitochondrial outer membrane permeabilization (MOMP), the process leading to mitochondrial apoptosis. Kinetics of the BCL-2 antagonist ABT-199 were included in the DR_MOMP model [<a href="#B10-cancers-12-02978" class="html-bibr">10</a>] to determine sensitisation of the apoptosis signalling network by the BCL-2 antagonist.</p>
Full article ">Figure 2
<p>DR_MOMP successfully predicated the sensitivity of two CRC PDXs to treatment with FOLFOX and ABT-199 in combination. (<b>A</b>) Two mCRC PDXs (CRC0344 and CRC0076) were assessed for BCL-2 family proteins levels by quantitative Western blot (<a href="#app1-cancers-12-02978" class="html-app">Figure S3</a>). (<b>B</b>) The susceptibility of each model to undergo MOMP in response to genotoxic chemotherapy alone or in combination treatment with ABT-199 was calculated using DR_MOMP. (<b>C</b>,<b>D</b>) Anti-tumour activity of ABT-199 alone and in combination with FOLFOX chemotherapy in the two models was assessed in vivo. Mice bearing either a CRC0076 or CRC0344 tumour were treated with vehicle control, ABT-199 100 mg/kg, FOLFOX or the combination ABT-199 + FOLFOX; <span class="html-italic">n</span> = 13–14/group, error bars = s.e.m., * <span class="html-italic">p</span> ≤ 0.05. The effect of treatment on absolute tumour volumes of the predicted combination only responder CRC0076 and the predicted FOLFOX alone responder CRC0344 PDX models are shown respectively (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 3
<p>Decrease in p53 upregulated modulator of apoptosis (PUMA) levels but no effect on BCL2-Interacting Mediator of cell death (BIM) or cleaved caspase-3, following ABT-199 treatment in CRC0076 PDX model. (<b>A</b>,<b>B</b>) Western blot analysis comparing the levels of PUMA and BIM in CRC0076 (<b>A</b>) and CRC0344 (<b>B</b>) PDX models. β-actin or glyceraldehyde 3-phosphate dehydrogenase (GAPDH) (see <a href="#app1-cancers-12-02978" class="html-app">Figure S4</a>) were used as loading controls. Error bars represent mean +/− 2 SD, * <span class="html-italic">p</span> ≤ 0.05 (ANOVA, Tukey post-hoc). (<b>C</b>,<b>D</b>) Analysis of cleaved caspase-3 levels in CRC0076 (<b>C</b>) and CRC0344 (<b>D</b>) PDX models. Representative images of H&amp;E (magnification 10×, scale bar 100 μm) and cleaved caspase-3 (magnification 5×, scale bar 200 μm) are shown for CRC0076 and CRC0344 in (<b>C</b>,<b>D</b>) respectively. The Allred method [<a href="#B15-cancers-12-02978" class="html-bibr">15</a>] was used to quantify the extent and intensity of staining, error bars represent mean +/− s.e.m (<span class="html-italic">n</span> = 3 tumours per treatment, <span class="html-italic">n</span> = 2–3 sections per tumour).</p>
Full article ">Figure 4
<p>Analysis of glucose uptake and metabolism in the predicted combination responder CRC0076 and the predicted FOLFOX alone responder CRC0344 PDX models. For imaging, <span class="html-italic">n</span> = 6 mice/group/PDX (at pre-treatment time point) received <sup>18</sup>F-FDG via tail vein injection, with image acquisition taking place 1 h after injection. Representative fused PET/CT images of CRC0076 and CRC0344 PDXs are shown (the same mice at pre- and on-treatment time points). White arrows indicate the tumours, scale bar indicates standard uptake values (SUVs) (<b>A</b>). N.B. Animals were imaged in a prone position. The coronal plane shown is positioned at the centre of the tumour which is located posterior to the bladder. Where the coronal plane intersects with the bladder, urinary excretion of FDG is observed as an intense signal. Comparison of relative change SUVmean at week 2 of treatment normalised to pre-treatment for CRC0076 and CRC0344 PDXs is presented. Error bars = s.e.m., * <span class="html-italic">p</span> ≤ 0.05 (paired <span class="html-italic">t</span>-tests, adjusted Holms) (<b>B</b>).</p>
Full article ">Figure 5
<p>Radiomic analysis of pre-treatment CT images of CRC0076 and CRC0344 can successfully distinguish between the PDXs using seven unique radiographic features. (<b>A</b>) Visual Representation of PDX radiomic analysis workflow. <span class="html-italic">n</span> = 45 baseline CT scans were partitioned into a training dataset (<span class="html-italic">n</span> = 36) and validation dataset (<span class="html-italic">n</span> = 9) before Recursive Feature Elimination was performed on the training dataset. (<b>B</b>) Representative CT axial slice of a mouse implanted with CRC PDX CRC0344. Region of interest (ROI, in red) was drawn semi-automatically on every slice containing tumour to enable segmentations of the tumour for radiomic feature extraction. (<b>C</b>) 3D representation of the mask used to segment the tumour. (<b>D</b>) Output of the recursive feature elimination (RFE) method. Using an RFE algorithm the ideal number of features was identified as 7 for radiomic model construction. (<b>E</b>): Graph depicting the result of the receiver operating characteristic (ROC) area under the curve (AUC) analysis of the 7 features identified by the model in a test cohort (<span class="html-italic">n</span> = 9). AUC is 0.8 (95% Confidence interval 0.408–1).</p>
Full article ">
Back to TopTop