[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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (79)

Search Parameters:
Keywords = BH3 mimetics

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 669
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 ">
15 pages, 4010 KiB  
Article
Modeling the Binding of Anticancer Peptides and Mcl-1
by Shamsa Husain Ahmed Alhammadi, Bincy Baby, Priya Antony, Amie Jobe, Raghad Salman Mohammed Humaid, Fatema Jumaa Ahmed Alhammadi and Ranjit Vijayan
Int. J. Mol. Sci. 2024, 25(12), 6529; https://doi.org/10.3390/ijms25126529 - 13 Jun 2024
Viewed by 586
Abstract
Mcl-1 (myeloid cell leukemia 1), a member of the Bcl-2 family, is upregulated in various types of cancer. Peptides representing the BH3 (Bcl-2 homology 3) region of pro-apoptotic proteins have been demonstrated to bind the hydrophobic groove of anti-apoptotic Mcl-1, and this interaction [...] Read more.
Mcl-1 (myeloid cell leukemia 1), a member of the Bcl-2 family, is upregulated in various types of cancer. Peptides representing the BH3 (Bcl-2 homology 3) region of pro-apoptotic proteins have been demonstrated to bind the hydrophobic groove of anti-apoptotic Mcl-1, and this interaction is responsible for regulating apoptosis. Structural studies have shown that, while there is high overall structural conservation among the anti-apoptotic Bcl-2 (B-cell lymphoma 2) proteins, differences in the surface groove of these proteins facilitates binding specificity. This binding specificity is crucial for the mechanism of action of the Bcl-2 family in regulating apoptosis. Bim-based peptides bind specifically to the hydrophobic groove of Mcl-1, emphasizing the importance of these interactions in the regulation of cell death. Molecular docking was performed with BH3-like peptides derived from Bim to identify high affinity peptides that bind to Mcl-1 and to understand the molecular mechanism of their interactions. The interactions of three identified peptides, E2gY, E2gI, and XXA1_F3dI, were further evaluated using 250 ns molecular dynamics simulations. Conserved hydrophobic residues of the peptides play an important role in their binding and the structural stability of the complexes. Understanding the molecular basis of interaction of these peptides will assist in the development of more effective Mcl-1 specific inhibitors. Full article
(This article belongs to the Special Issue New Insights into Anti-cancer Drug Discovery and Development)
Show Figures

Figure 1

Figure 1
<p>Structure of Mcl-1 (grey) with a modified Bim BH3 peptide, SAH-MS1-18 (cyan) (PDB: 5W89) [<a href="#B35-ijms-25-06529" class="html-bibr">35</a>].</p>
Full article ">Figure 2
<p>Sequence of the top scoring Bim-based peptides. The heptad convention used to refer to positions in the BH3 peptide is shown. Numbering uses the convention (abcdefg)n. Complete heptad repeats 2 to 4 are indicated above the sequences.</p>
Full article ">Figure 3
<p>(<b>A</b>) The crystal structure of Mcl-1 (grey), complexed with a modified Bim BH3 peptide SAH-MS1-18 (cyan) (PDB ID:5W89) [<a href="#B35-ijms-25-06529" class="html-bibr">35</a>] as well as docked E2gI (orange), E2gY (green), and XXA1 F3dI (pink). (<b>B</b>) Docked pose of E2gY in Mcl-1. (<b>C</b>) Docked pose of E2gI in Mcl-1. (<b>D</b>) Docked pose of XXA1 F3dI in Mcl-1. Mcl-1 residues are colored in blue.</p>
Full article ">Figure 4
<p>(<b>A</b>) Root mean square deviation (RMSD) of protein Cα atoms obtained from 250 ns simulations of the crystal structure (PDB ID: 5W89) with bound SAH-MS1-18 (cyan) and docked E2gI (orange), E2gY (green), and XXA1 F3dI (pink). (<b>B</b>) Root mean square fluctuation (RMSF) of protein residues obtained from 250 ns simulations. Crystal structure (PDB ID: 5W89) with bound SAH-MS1-18 (cyan) and docked E2gI (orange), E2gY(green), and XXA1 F3dI (pink).</p>
Full article ">Figure 5
<p>Percentage of simulation time during which intermolecular polar and hydrophobic contacts were retained between Mcl-1 and peptides in the 250 ns systems. (<b>A</b>) Mcl-1/SAH-MS1-18 inhibitor, and (<b>B</b>) Mcl-1/E2gI.</p>
Full article ">Figure 6
<p>Percentage of simulation time during which intermolecular polar and hydrophobic contacts were retained between Mcl-1 and peptides in the 250 ns systems. (<b>A</b>) Mcl-1/E2gY, and (<b>B</b>) Mcl-1/XXA1 F3dI.</p>
Full article ">
19 pages, 3970 KiB  
Article
Coupling Kinesin Spindle Protein and Aurora B Inhibition with Apoptosis Induction Enhances Oral Cancer Cell Killing
by João P. N. Silva, Bárbara Pinto, Luís Monteiro, Patrícia M. A. Silva and Hassan Bousbaa
Cancers 2024, 16(11), 2014; https://doi.org/10.3390/cancers16112014 - 25 May 2024
Viewed by 829
Abstract
Many proteins regulating mitosis have emerged as targets for cancer therapy, including the kinesin spindle protein (KSP) and Aurora kinase B (AurB). KSP is crucial for proper spindle pole separation during mitosis, while AurB plays roles in chromosome segregation and cytokinesis. Agents targeting [...] Read more.
Many proteins regulating mitosis have emerged as targets for cancer therapy, including the kinesin spindle protein (KSP) and Aurora kinase B (AurB). KSP is crucial for proper spindle pole separation during mitosis, while AurB plays roles in chromosome segregation and cytokinesis. Agents targeting KSP and AurB selectively affect dividing cells and have shown significant activity in vitro. However, these drugs, despite advancing to clinical trials, often yield unsatisfactory outcomes as monotherapy, likely due to variable responses driven by cyclin B degradation and apoptosis signal accumulation networks. Accumulated data suggest that combining emerging antimitotics with various cytostatic drugs can enhance tumor-killing effects compared to monotherapy. Here, we investigated the impact of inhibiting anti-apoptotic signals with the BH3-mimetic Navitoclax in oral cancer cells treated with the selective KSP inhibitor, Ispinesib, or AurB inhibitor, Barasertib, aiming to potentiate cell death. The combination of BH3-mimetics with both KSP and AurB inhibitors synergistically induced substantial cell death, primarily through apoptosis. A mechanistic analysis underlying this synergistic activity, undertaken by live-cell imaging, is presented. Our data underscore the importance of combining BH3-mimetics with antimitotics in clinical trials to maximize their effectiveness. Full article
(This article belongs to the Topic Recent Advances in Anticancer Strategies)
Show Figures

Figure 1

Figure 1
<p>KSP and Aurora B show increased expression in oral squamous carcinoma cell lines. The mRNA expression levels of KSP (<b>a</b>) and Aurora B (<b>c</b>) were assessed through qRT-PCR in the oral cancer cell lines SCC09 and SCC25 and compared to the non-tumor human oral keratinocyte (HOK) cells. The quantification of the protein levels of KSP (<b>b</b>, left) and Aurora B (<b>d</b>, left) was performed by the Western blotting assay, with the protein α-tubulin as control. The representative Western blot images for KSP (<b>b</b>, right) and Aurora B (<b>d</b>, right) are presented. The data presented indicate the mean value along with the standard deviation (mean ± SD) obtained from three independent experiments. Statistical analysis was performed using one-way ANOVA followed by Tukey’s multiple comparisons test. The significance levels were as follows: * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.01, and *** for <span class="html-italic">p</span> &lt; 0.001. Original western blots are presented in File S1.</p>
Full article ">Figure 2
<p>Mitotic defects induced by KSP and Aurora B inhibition. Illustrative immunofluorescence images showing SCC25 cells phenotype after 24 h treatment with 1.875 nM of Ispinesib or 1000 nM of Barasertib (<b>a</b>,<b>b</b>). DAPI was used to stain DNA (blue), while α-tubulin was stained to allow the visualization of microtubules (green), and CREST (red) for kinetochores localization. Bar, 5 μm.</p>
Full article ">Figure 3
<p>Dose−response curves of Ispinesib, Barasertib, and Navitoclax in SCC25 (<b>a</b>) and SCC09 (<b>b</b>) cell lines. The percentage of cell viability vs. the concentration of the different inhibitors (logarithmic scale) is shown. The R<sup>2</sup> values are shown for each curve indicating the fit of the model to the data.</p>
Full article ">Figure 4
<p>The combinatorial approaches Ispinesib + Navitoclax and Barasertib + Navitoclax enhance cytotoxicity in the SCC25 and SCC09 cell lines. Cell viability (%) following 48 h of drug exposure both alone or in combination (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>), assessed by MTT assay with at least three independent experiments. The synergy scores were calculated using the Bliss model of the Combenefit software 2.021. Asterisks denote synergistic effects with statistical significance of * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>). The representative images of colony formation assays following 6 days with SCC25 cells (<b>i</b>) are presented. The quantification of survival fraction (%) following treatment with drugs both alone and in combination is illustrated (<b>j</b>,<b>k</b>). The data presented are the average ± standard deviation of three separate experiments. Statistical analysis was performed using one-way ANOVA followed by Tukey’s post hoc test for multiple comparisons. The significance levels were as follows: ** for <span class="html-italic">p</span> &lt; 0.01; *** for <span class="html-italic">p</span> &lt; 0.001; and **** for <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Addition of Navitoclax to Ispinesib increases cell death during mitosis in oral cancer cells. The representative images acquired by phase-contrast microscopy after drug exposure for 24 h for SCC25 (<b>a</b>). Mitotic index quantification for the SCC25 cell line (<b>b</b>); 0.2% of DMSO (drug solvent) was used as the negative control, while 1 μM of Nocodazole (mitotic blocker drug) was used as the positive control. The data presented are the average ± standard deviation of three separate experiments. Statistical analysis was performed using one-way ANOVA followed by Tukey’s post hoc test for multiple comparisons. **** <span class="html-italic">p</span> &lt; 0.0001. The measurement of the duration of mitosis following the indicated drug treatments by time-lapse microscopy (<b>c</b>). The assessment of cell fate (%) over 48 h using indicated treatments (<b>f</b>). The representative time-lapse image sequences acquired during 48 h of exposure to drugs both alone and in combination (<b>d</b>). The assess-ment of cell fate (%) over 48 h using indicated treatments (<b>e</b>). The data presented are the average ± standard deviation of three separate experiments. Statistical analysis was performed using two-way ANOVA followed by Tukey’s post hoc test for multiple comparisons. #### (<span class="html-italic">p</span> &lt; 0.0001) statistically significant difference in the cells that underwent death in mitosis (%) between 1500 nM Navitoclax or 1.875 nM Ispinesib and 1.875 nM Ispinesib + 1500 nM Navitoclax. *** (<span class="html-italic">p</span> &lt; 0.001) postmitotic survival cell (%) difference between 1500 nM Navitoclax and 1.875 nM Ispinesib + 1500 nM Navitoclax. **** (<span class="html-italic">p</span> &lt; 0.0001) postmitotic survival cell (%) difference between 1.875 nM Ispinesib and 1.875 nM Ispinesib + 1500 nM Navitoclax. The combination of Ispinesib and Navitoclax enhances cell death in the SCC25 oral cancer cell line. The quantification of Annexin-V-positive cells (<b>f</b>). Cytograms demonstrative of the oral cancer cells double stained with Annexin V-FITC and propidium iodide (PI) (<b>g</b>). The quadrants Q are defined as Q1 = living cells (Annexin V- and PI-negative), Q2 = early-stage apoptosis (Annexin V-positive/PI-negative), and Q3 = late-stage apoptosis/secondary necrosis (Annexin V- and PI-positive). The data presented are the average ± standard deviation of three separate experiments. Statistical analysis was performed using one-way ANOVA followed by Tukey’s post hoc test for multiple comparisons. The significance levels were as follows: **** for <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Addition of Navitoclax to Barasertib increases post-slippage death in oral cancer cells. The measurement of the duration of mitosis following the indicated drug treatments by time-lapse microscopy (<b>a</b>). The assessment of cell fate (%) over 48 h using indicated treatments (<b>b</b>). The representative time-lapse image sequences acquired during 48 h of exposure to drugs both alone and in combination (<b>c</b>). The data presented are the average ± standard deviation of three separate experiments. Statistical analysis was performed using two-way ANOVA followed by Tukey’s post hoc test for multiple comparisons. # (<span class="html-italic">p</span> &lt; 0.05) statistically significant difference in the cells that underwent post-slippage death (%) between 3000 nM Navitoclax or 1000 nM Barasertib and 1000 nM Barasertib + 3000 nM Navitoclax. **** (<span class="html-italic">p</span> &lt; 0.0001) postmitotic survival cell (%) difference between 3000 nM Navitoclax and 1000 nM Barasertib + 3000 nM Navitoclax. <span>$</span><span>$</span><span>$</span><span>$</span> (<span class="html-italic">p</span> &lt; 0.0001) post-slippage survival cell (%) difference between 1000 nM Barasertib and 1000 nM Barasertib + 3000 nM Navitoclax. The addition of Barasertib to Navitoclax enhances cell death in the oral cancer cell line SCC25. The quantification of Annexin-V-positive cells (<b>d</b>). Cytograms demonstrative of the oral cancer cells double stained with Annexin V-FITC and propidium iodide (PI) (<b>e</b>). The quadrants Q are defined as Q1 = living cells (Annexin V- and PI-negative), Q2 = early-stage apoptosis (Annexin V-positive/PI-negative), and Q3 = late-stage apoptosis/secondary necrosis (Annexin V- and PI-positive). The data presented are the average ± standard deviation of three separate experiments. Statistical analysis was performed using one-way ANOVA followed by Tukey’s post hoc test for multiple comparisons. The significance levels were as follows: * for <span class="html-italic">p</span> &lt; 0.05; ** for <span class="html-italic">p</span> &lt; 0.01; and **** for <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
19 pages, 877 KiB  
Review
Venetoclax-Related Neutropenia in Leukemic Patients: A Comprehensive Review of the Underlying Causes, Risk Factors, and Management
by Laura Giuseppina Di Pasqua, Murwan Mahmoud Abdallah, Fausto Feletti, Mariapia Vairetti and Andrea Ferrigno
Pharmaceuticals 2024, 17(4), 484; https://doi.org/10.3390/ph17040484 - 10 Apr 2024
Viewed by 1537
Abstract
Venetoclax is a Bcl-2 homology domain 3 (BH3) mimetic currently approved for the treatment of chronic lymphocytic leukemia (CLL) and acute myeloid leukemia (AML) that has proven to be highly effective in reinstating apoptosis in leukemic cells through the highly selective inhibition of [...] Read more.
Venetoclax is a Bcl-2 homology domain 3 (BH3) mimetic currently approved for the treatment of chronic lymphocytic leukemia (CLL) and acute myeloid leukemia (AML) that has proven to be highly effective in reinstating apoptosis in leukemic cells through the highly selective inhibition of the anti-apoptotic protein B-cell lymphoma-2 (Bcl-2). Clinically, venetoclax has provided lasting remissions through the inhibition of CLL and AML blasts. However, this activity has often come at the cost of grade III/IV neutropenia due to hematopoietic cells’ dependence on Bcl-2 for survival. As life-threatening infections are an important complication in these patients, an effective management of neutropenia is indispensable to maximize patient outcomes. While there is general consensus over dose reduction and scheduling modifications to minimize the risk of neutropenia, the impact of these modifications on survival is uncertain. Moreover, guidelines do not yet adequately account for patient-specific and disease-specific risk factors that may predict toxicity, or the role combination treatment plays in exacerbating neutropenia. The objective of this review is to discuss the venetoclax-induced mechanism of hematological toxicity, the potential predictive risk factors that affect patient vulnerability to neutropenia, and the current consensus on practices for management of neutropenia. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Figure 1

Figure 1
<p>The intrinsic apoptotic pathway triggers the activation of the pro-apoptotic members of the BCL-2 protein family, known as BH3-only proteins. These BH3-only proteins interact with and neutralize pro-survival BCL-2 proteins, effectively releasing the key apoptotic effectors BAK and BAX. BAK and BAX then assemble into large complexes that induce ruptures in the mitochondrial outer membrane (mitochondrial permeability transition pore, MPTP), leading to the release of apoptogenic factors, such as cytochrome c. Notably, certain BH3-only proteins have been shown to directly bind to and activate BAK and BAX to induce MPTP as well. The BH3-only mimetic venetoclax displaces and reactivates pro-apoptotic proteins bound to the BH3-binding groove of BCL2, leading to the assembly of MPTP.</p>
Full article ">Figure 2
<p>Mechanism of venetoclax-induced neutropenia. The anti-apoptotic Bcl-2 expression is readily detectable at the protein level on neutrophil precursors, while it is very low in mature neutrophils. Bcl-2 inhibition triggers apoptosis in neutrophile precursors, resulting in neutropenia.</p>
Full article ">
15 pages, 5275 KiB  
Article
FLT3 and IRAK4 Inhibitor Emavusertib in Combination with BH3-Mimetics in the Treatment of Acute Myeloid Leukemia
by Katja Seipel, Harpreet Mandhair, Ulrike Bacher and Thomas Pabst
Curr. Issues Mol. Biol. 2024, 46(4), 2946-2960; https://doi.org/10.3390/cimb46040184 - 29 Mar 2024
Viewed by 1224
Abstract
Targeting the FLT3 receptor and the IL-1R associated kinase 4 as well as the anti-apoptotic proteins MCL1 and BCL2 may be a promising novel approach in the treatment of acute myeloid leukemia (AML). The FLT3 and IRAK4 inhibitor emavusertib (CA4948), the MCL1 inhibitor [...] Read more.
Targeting the FLT3 receptor and the IL-1R associated kinase 4 as well as the anti-apoptotic proteins MCL1 and BCL2 may be a promising novel approach in the treatment of acute myeloid leukemia (AML). The FLT3 and IRAK4 inhibitor emavusertib (CA4948), the MCL1 inhibitor S63845, the BCL2 inhibitor venetoclax, and the HSP90 inhibitor PU-H71 were assessed as single agents and in combination for their ability to induce apoptosis and cell death in leukemic cells in vitro. AML cells represented all major morphologic and molecular subtypes, including FLT3-ITD and NPM1 mutant AML cell lines and a variety of patient-derived AML cells. Emavusertib in combination with MCL1 inhibitor S63845 or BCL2 inhibitor venetoclax induced cell cycle arrest and apoptosis in MOLM-13 cells. In primary AML cells, the response to emavusertib was associated with the presence of the FLT3 gene mutation with an allelic ratio >0.5 and the presence of NPM1 gene mutations. S63845 was effective in all tested AML cell lines and primary AML samples. Blast cell percentage was positively associated with the response to CA4948, S63845, and venetoclax, with elevated susceptibility of primary AML with blast cell fraction >80%. Biomarkers of the response to venetoclax included the blast cell percentage and bone marrow infiltration rate, as well as the expression levels of CD11b, CD64, and CD117. Elevated susceptibility to CA4948 combination treatments with S63845 or PU-H71 was associated with FLT3-mutated AML and CD34 < 30%. The combination of CA4948 and BH3-mimetics may be effective in the treatment in FLT3-mutated AML with differential target specificity for MCL1 and BCL2 inhibitors. Moreover, the combination of CA4948 and PU-H71 may be a candidate combination treatment in FLT3-mutated AML. Full article
(This article belongs to the Special Issue Molecular Research and Pathological Mechanism of Leukemia)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>In vitro dose response to emavusertib (CA4948) treatment in AML cell lines. AML cells were treated with the dual IRAK4 and FLT3 inhibitor CA4948 at the indicated dosages for 20 h (<b>A</b>). Susceptibility of AML cells to CA4948 under exposure to lipopolysaccharides (<b>B</b>,<b>C</b>). Cell viability was determined in AML cells after exposure to 50–250 pg/mL LPS and 50 to 150 nM CA4948 in FLT3-ITD positive MOLM-13 (<b>B</b>) and FLT3 wild-type ML-2 (<b>C</b>) cells. Cell viability data are average values of multiple repeat measurements per dosage. The standard deviation was 3–6%. Significance of differences denoted for <span class="html-italic">p</span> &lt; 0.01 (**); and <span class="html-italic">p</span> &lt; 0.0001 (****).</p>
Full article ">Figure 2
<p>Schematic presentation of the FLT3 and SCFR (CD117, c-KIT) as well as IL1R and TLR4 signaling pathways and downstream effects. FLT3-ITD and FLT3-TKD are constitutively active growth factor receptors. SCFR (c-KIT) is an inducible growth factor receptor activated by stem cell factor (SCF). Both tyrosine receptor kinases activate PI3K-AKT, RAS-MEK-ERK, and STAT5, leading to cell growth and proliferation via inhibition of the tumor suppressor TP53 and induction of the apoptosis regulators MCL1 and BCL2. IL1R and TLR4 are inducible receptors signaling via MyD88 and IRAK4. HSP90 protein can bind and stabilize client proteins, including AKT, BCL2, FLT3, JAK, MDM2, STAT5, SHP2, and BRAF. Hsp90 proteins are indicated in yellow, oncogenic protein functions in red, tumor suppressor functions in green ovals, and targeted inhibitors in blue rectangles. Sharp arrows and blunt arrows indicate target induction and inhibition, respectively.</p>
Full article ">Figure 3
<p>Susceptibility of AML cell lines to various treatment combinations. Cell viability was determined in AML cell lines MOLM-13, ML-2, and OCI-AML3 after 20 h of treatment with single compounds and in combination with 50–150 nM CA4948 (CA) and 10–100 nM venetoclax (VC100), 100–200 nM PU-H71 (PU), or 50–100 nM S63845 (S100). Susceptibility of MOLM-13 cells to various treatment combinations in the absence or presence of HS-5 bone marrow stroma (<b>B</b>). Cell viability data are average values of multiple repeat measurements per dosage. Synergistic effects of combination treatments are depicted in MOLM-13 (<b>A</b>). Oppositional effects in ML-2 (<b>C</b>) and OCI-AML3 (<b>D</b>) cells. Cell viability data are average values of multiple repeat measurements per dosage. Significance of differences denoted for <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 4
<p>Induction of cell cycle arrest, apoptosis, and cell death in AML cells treated with CA4948 in combination with PU-H71, S63845, or venetoclax. Cytometric analysis of MOLM-13 cells treated with 100 nM PU-H71 (PU), 50 nM S63845 (S), or 25 nM venetoclax (VC) and stained with Annexin-V and PI (<b>A</b>,<b>C</b>) or DAPI (<b>B</b>,<b>D</b>). According to 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). According to DAPI staining intensity, cells were classified as subG1 (&lt;2 N), G0/G1 (2 N), S phase (2–4 N), or G2 phase (4 N). Significance of differences denoted for <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 5
<p>Primary AML cells’ in vitro response to Emavusertib and combination treatments. Cell viability was determined in mononuclear cells isolated from the peripheral blood or bone marrow of 23 AML patients (A1 to A23) after 20 h of treatment. (<b>A</b>) Primary AML cells treated with 100 nM CA4948 (CA100) and 100 nM S63845 (S100), alone and in combination. (<b>B</b>) Primary AML cells treated with 100 nM CA4948 (CA100) and 100 nM venetoclax (VC100), alone and in combination. (<b>C</b>) Primary AML cells treated with 100 nM CA4948 (CA100) and 100 nM PU-H71 (PU100), alone and in combination. Treatment for 20 h with 100 nM CA4948 and 100 nM S63845 (<b>D</b>), 100 nM CA4948 and 100 nM venetoclax (<b>E</b>), and 100 nM CA4948 and 100 nM PU-H71 (<b>F</b>). Significance was calculated by paired <span class="html-italic">t</span>-test. The patient samples were sorted into two groups, susceptible (S) and resistant (R), in a single compound treatment with CA4948 (<b>G</b>), S63845 (<b>H</b>), venetoclax (<b>I</b>), and PU-H71 (<b>J</b>), as well as in different combination treatments with CA4948 and S63845 (<b>K</b>), CA4948 and venetoclax (<b>L</b>), and CA4948 and PU-H71 (<b>M</b>). Significance of differences in median values was calculated using the Mann–Whitney test. Significance is denoted for <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>Biomarkers of response to emavusertib and combination treatments. Cell viability was determined in mononuclear cells isolated from AML patients’ peripheral blood or bone marrow after 20 h of treatment with 100 nM CA4948 (CA100), 100 nM S63845 (S100), 100 nM venetoclax (VC100), and 100 nM PU-H71 (PU100), alone and in combination. Response was correlated to <span class="html-italic">FLT3</span> gene mutation status (<b>A</b>), <span class="html-italic">NPM1</span> gene mutation status (<b>B</b>), blast cell percentage (<b>C</b>), bone marrow infiltration (<b>D</b>), CD11b (<b>E</b>), CD34 (<b>F</b>), CD64 (<b>G</b>), and CD117 (<b>H</b>). Best fit lines are according to simple linear regression analysis. Association of response and biomarkers was analyzed by unpaired <span class="html-italic">t</span>-test (<b>I</b>). Significance denoted for <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 (***). Legend for treatment combinations (<b>J</b>).</p>
Full article ">
20 pages, 1155 KiB  
Review
Innovative Combinations, Cellular Therapies and Bispecific Antibodies for Chronic Lymphocytic Leukemia: A Narrative Review
by Andrea Visentin, Sara Frazzetto, Livio Trentin and Annalisa Chiarenza
Cancers 2024, 16(7), 1290; https://doi.org/10.3390/cancers16071290 - 26 Mar 2024
Viewed by 1151
Abstract
In the last few years, several agents targeting molecules that sustain the survival and the proliferation of chronic lymphocytic leukemia (CLL) cells have become clinically available. Most of these drugs target surface proteins, such as CD19 or CD20, via monoclonal or bispecific monoclonal [...] Read more.
In the last few years, several agents targeting molecules that sustain the survival and the proliferation of chronic lymphocytic leukemia (CLL) cells have become clinically available. Most of these drugs target surface proteins, such as CD19 or CD20, via monoclonal or bispecific monoclonal antibodies (BsAbs), CAR T cells, intracellular proteins like BTK by using covalent or non-covalent inhibitors or BCL2 with first or second generation BH3-mimetics. Since the management of CLL is evolving quickly, in this review we highlighted the most important innovative treatments including novel double and triple combination therapies, CAR T cells and BsAbs for CLL. Recently, a large number of studies on novel combinations and newer strategic options for CLL therapy have been published or presented at international conferences, which were summarized and linked together. Although the management of treatment with a single continuous agent is easier, the emergence of protein mutations, long-term toxicities and costs are important concerns that favor the use of a fixed duration therapy. In the future, a measurable residual disease (MRD)-guided treatment cessation and MRD-based re-initiation of targeted therapy seems to be a more feasible approach, allowing identification of the patients who might benefit from continuous therapy or who might need a consolidation with BsAbs or CAR T cells to clear the neoplastic clone. Full article
(This article belongs to the Special Issue Towards a Tailored Treatment of Chronic Lymphocytic Leukemia)
Show Figures

Figure 1

Figure 1
<p>Current and novel targeted therapies in CLL. The neoplastic cells of chronic lymphocytic leukemia express on their surfaces CD5, CD20 (even if a low levels), CD23 and CD200. CD20 can be targeted by anti-CD20 monoclonal antibodies (mAbs) like rituximab, ofatumumab or obinutuzumab. These cells move from the peripheral blood to the bone morrow and secondary lymphoid organs thanks to integrins, such CD49d, cytokines and chemokine receptors, like CXCR4 [<a href="#B10-cancers-16-01290" class="html-bibr">10</a>]. One of the main membrane receptors is the B-cell receptor (BCR), which after antigen engagement actives several kinases such as Lyn, Syk, PI3K and BTK, which play a key role in mediating the survival and the proliferation of CLL cells [<a href="#B11-cancers-16-01290" class="html-bibr">11</a>,<a href="#B12-cancers-16-01290" class="html-bibr">12</a>]. BTK (Bruton’s Tyrosine kinase) can be targeted by covalent inhibitors such as ibrutinib, acalabrutinib and zanubrutinib but also the non-covalent inhibitors pirtobrutinib and nemtabrutinib and CDACs (chimeric degradation activation compounds) like BGB-16673, NX-2127 and NX-5948 [<a href="#B13-cancers-16-01290" class="html-bibr">13</a>,<a href="#B14-cancers-16-01290" class="html-bibr">14</a>]. CLL cells are also characterized by the overexpression of the anti-apoptotic protein BCL2, mainly due to the loss of mir15-a and mir16-1, whose main target is the mRNA derived from the <span class="html-italic">BCL2</span> gene. The BH3 mimetics, such as venetoclax, sonrotoclax and lisaftoclax are able to disrupt the binding of BCL2 with BID, triggering apoptosis. The tumor microenvironment represented by the nurse-like cells, macrophages, mesenchymal stromal cells and T-cells is another important player in sustaining the survival of CLL cells and favoring the resistance to therapies. Bispecific antibodies (BsAbs), such as blinatumomab, mosunetuzumab, glofitamab epcoritamab and plamatomab favor the activation of T cells against neoplastic cells. In a similar way, chimeric antigen receptors (CARs), once expressed by T cells (CAR T-cells), such as lisocabtagene maraleucel (liso-cel), improve the T-cell killing of CLL cells.</p>
Full article ">
21 pages, 2973 KiB  
Article
Effective Targeting of Melanoma Cells by Combination of Mcl-1 and Bcl-2/Bcl-xL/Bcl-w Inhibitors
by Zhe Peng, Bernhard Gillissen, Antje Richter, Tobias Sinnberg, Max S. Schlaak and Jürgen Eberle
Int. J. Mol. Sci. 2024, 25(6), 3453; https://doi.org/10.3390/ijms25063453 - 19 Mar 2024
Cited by 2 | Viewed by 1159
Abstract
Recent advances in melanoma therapy have significantly improved the prognosis of metastasized melanoma. However, large therapeutic gaps remain that need to be closed by new strategies. Antiapoptotic Bcl-2 proteins critically contribute to apoptosis deficiency and therapy resistance. They can be targeted by BH3 [...] Read more.
Recent advances in melanoma therapy have significantly improved the prognosis of metastasized melanoma. However, large therapeutic gaps remain that need to be closed by new strategies. Antiapoptotic Bcl-2 proteins critically contribute to apoptosis deficiency and therapy resistance. They can be targeted by BH3 mimetics, small molecule antagonists that mimic the Bcl-2 homology domain 3 (BH3) of proapoptotic BH3-only proteins. By applying in vitro experiments, we aimed to obtain an overview of the possible suitability of BH3 mimetics for future melanoma therapy. Thus, we investigated the effects of ABT-737 and ABT-263, which target Bcl-2, Bcl-xL and Bcl-w as well as the Bcl-2-selective ABT-199 and the Mcl-1-selective S63845, in a panel of four BRAF-mutated and BRAF-WT melanoma cell lines. None of the inhibitors showed significant effectiveness when used alone; however, combination of S63845 with each one of the three ABTs almost completely abolished melanoma cell survival and induced apoptosis in up to 50–90% of the cells. Special emphasis was placed here on the understanding of the downstream pathways involved, which may allow improved applications of these strategies. Thus, cell death induction was correlated with caspase activation, loss of mitochondrial membrane potential, phosphorylation of histone H2AX, and ROS production. Caspase dependency was demonstrated by a caspase inhibitor, which blocked all effects. Upregulation of Mcl-1, induced by S63845 itself, as reported previously, was blocked by the combinations. Indeed, Mcl-1, as well as XIAP (X-linked inhibitor of apoptosis), were strongly downregulated by combination treatments. These findings demonstrate that melanoma cells can be efficiently targeted by BH3 mimetics, but the right combinations have to be selected. The observed pronounced activation of apoptosis pathways demonstrates the decisive role of apoptosis in the loss of cell viability by BH3 mimetics. Full article
(This article belongs to the Special Issue Advances in Melanoma and Skin Cancers)
Show Figures

Figure 1

Figure 1
<p>Combinations of BH3 mimetics decrease cell viability and induce apoptosis in melanoma cells at 24 h. Melanoma cell lines A-375, Mel-HO, MeWo, and SK-Mel-23 were seeded in 24-well plates and were treated with S63845 (S63, 1 µM) as well as with ABT-199, ABT-263, or ABT-737 (0.01, 0.1, 1 µM) as indicated. (<b>A</b>) After 24 h, cell viability was determined via calcein-AM staining and flow cytometry. Values represent the percentage of cells with high calcein staining (viable cells). Effects on cell viability are displayed as a percentage of non-treated controls (100%). (<b>B</b>) After 24 h, apoptotic cells were identified as sub-G1 cells in cell cycle analyses via flow cytometry after propidium-iodide staining. (<b>A</b>,<b>B</b>) At least two series of experiments were performed, each one consisting of independent triplicate values. Mean values of all individual values (at least 6) are shown here. Statistical significance is indicated by asterisks (*; <span class="html-italic">p</span> &lt; 0.05) and was calculated for the single treatments as compared to non-treated control cells. The combination treatments were compared both to the respective single treatments with ABT (light bars) and to S63845 alone. Largely comparable findings were obtained after 48 h treatments (<a href="#app1-ijms-25-03453" class="html-app">Supplementary Figure S2</a>). Example flow cytometry readings of cells treated for 48 h with combinations of ABTs and S63845 (1 µM concentrations) are shown as overlays vs. controls. Non-viable and viable cell populations as well as cell cycle phases G1 (gap 1), S (synthesis), G2 (gap 2), and sub-G1 cells are indicated.</p>
Full article ">Figure 2
<p>Early induction of apoptosis plays a leading role. Mel-HO and SK-Mel-23 cells were seeded in 24-well plates treated with ABT-199, ABT-263, ABT-737, and/or S63845 (1 µM concentrations; ---, no ABTs were added). (<b>A</b>) Cell death analysis by Annexin V/PI staining and flow cytometry was performed at 6 h, 12 h, and 24 h. Early apoptotic cells were determined as AnnV(+)/PI(−), while late apoptotic or necrotic cells corresponded to AnnV(+)/PI(+) staining. Mean values (in %) and SDs were calculated of three series of experiments, each one consisting of double values (six values in a group). Statistically significant changes observed in combination treatments, as compared to the respective single treatment with ABT and to S63845 alone, are indicated for the two cell death fractions (*; <span class="html-italic">p</span> &lt; 0.01). (<b>B</b>) Representative flow cytometry histograms of combination-treated cells and control cells are shown. Readings are separated into four quadrants (according to PI and AnnV positivity). Thus, the lower left quadrant corresponds to AnnV(−)/PI(−) cells (viable cells), the lower right quadrant corresponds to AnnV(+)/PI(−) cells (early apoptosis), and the upper right quadrant corresponds to AnnV(+)/PI(+) cells (late apoptosis).</p>
Full article ">Figure 3
<p>Combinations of BH3 mimetics induce mitochondrial apoptosis in vitro. The four cell lines were treated with S63845 (S63), ABT-199, ABT-263, ABT-737, and combinations (1 µM concentrations; ---, no ABTs were added). (<b>A</b>) Mitochondrial membrane potential (MMP) was determined at 4 h and at 24 h via TMRM<sup>+</sup> staining and flow cytometry. Values represent the percentage of cells with low MMP. Mean values and SDs (in %) were calculated for two series of experiments, each one consisting of independent triplicate values (six values in a group). Statistical significance (* <span class="html-italic">p</span> &lt; 0.05) was calculated for combination treatments as compared to the respective single treatments (ABTs alone and S63845 alone). (<b>B</b>) Examples of flow cytometry readings are shown as overlays of combination treatments (dark graphs) vs. the controls (open graphs). Cell populations with low MMP are indicated. (<b>C</b>) For microscopic visualization of low MMP, Mel-HO and SK-Mel-23 cells were stained with JC-1 and counterstained with Hoechst-33342 at 6 h and at 12 h of treatment. Cell nuclei are stained in blue, while mitochondria with high (normal) MMP are stained in red. Green staining indicates cytosolic JC1 localization upon release of JC1 from mitochondria with low MMP. Two independent experiments revealed highly comparable results.</p>
Full article ">Figure 4
<p>Reactive oxygen species are produced after combination treatment. A-375, Mel-HO, MeWo, and SK-Mel-23 cells were treated with ABT-199, ABT-263, ABT-737, S63845 (S63), and combinations (1 µM concentrations; ---, no ABTs were added). (<b>A</b>) Cellular levels of ROS were determined at 4 h and at 24 h via H<sub>2</sub>DCF-DA staining and flow cytometry. Values represent the percentage of cells with high ROS. Mean values and SDs (in %) were calculated for two independent experiments, each one consisting of triplicate values (six values in a group). Statistical significance (* <span class="html-italic">p</span> &lt; 0.05) was calculated for combination treatments as compared to the respective single treatments (ABTs alone and S63845 alone). (<b>B</b>) Examples of flow cytometry readings are shown on the right side for the combination treatments (dark graphs) vs. the controls (open graphs). Cell populations with high ROS are indicated.</p>
Full article ">Figure 5
<p>Pro-apoptotic pathways are efficiently activated only after combination treatment. Mel-HO and SK-Mel-23 cells were treated for 8 h with ABT-199, ABT-263, ABT-737, S63845, and combinations (1 µM concentrations). Total protein extracts were analyzed by Western blotting for cleaved caspase-3, total caspase-8, and caspase-9 (Csp). Further, processing of PARP from 116 to 89 kD and phosphorylation of histone H2AX (γ-H2AX) were analyzed. Equal protein amounts (30 µg per lane) were separated by SDS-PAGE, and consistent blotting was proven by Ponceau staining as well as via evaluation of GAPDH expression. Molecular weights are indicated in kD. Each two independent series of protein extracts and Western blots revealed highly comparable results.</p>
Full article ">Figure 6
<p>Cell death induced by BH3 mimetic combinations in melanoma cells is caspase-dependent in vitro. Mel-HO and SK-Mel-23 cells were treated with combinations of S63845 with ABT-199, ABT-263, and ABT-737 (1 µM concentrations). When indicated, the pan-caspase inhibitor QVD-Oph (QVD, 5 µM) was applied at 1 h before other treatments started. Cell viability (calcein staining; (<b>A</b>)), apoptosis induction (cell cycle analysis; (<b>B</b>)), loss of MMP (TMRM<sup>+</sup> staining; (<b>C</b>)) and production of ROS (H<sub>2</sub>DCF-DA staining; (<b>D</b>)) were determined via flow cytometry at 24 h. Mean values and SDs (in %) were calculated from two series of experiments, each one consisting of independent triplicate values. Statistical significance (* <span class="html-italic">p</span> &lt; 0.05) was calculated for QVD treatments as compared to ABT/S63845 combinations without QVD (at least 6 individual values in a group). On the right side, example flow cytometry readings of ABT/S63845 combination treatments +/− QVD are shown as overlays. Viable and non-viable cell populations (<b>A</b>), sub-G1, G1, S and G2 cell populations (<b>B</b>), cell populations with low MMP (<b>C</b>), and cell populations with high ROS levels (<b>D</b>) are indicated.</p>
Full article ">Figure 7
<p>Apoptosis proteins Mcl-1 and XIAP are downregulated by combination treatments. Mel-HO and SK-Mel-23 cells were treated with ABT-199, ABT-263, ABT-737, S63845, and combinations (1 µM concentrations, treatment time: 8 h). Total protein extracts were analyzed by Western blotting for expression of Mcl-1 (41 kD), Bcl-2 (26 kD), Bcl-w (18 kD), Bcl-x<sub>L</sub> (30 kD), and XIAP (53 kD). Equal protein amounts (30 µg per lane) were separated by SDS-PAGE, and consistent blotting was proven via Ponceau staining as well as through evaluation of GAPDH expression. Molecular weights are indicated in kD. Two independent series of protein extracts and Western blots revealed highly comparable results.</p>
Full article ">
19 pages, 1233 KiB  
Review
Venetoclax Resistance in Acute Myeloid Leukemia
by Sylvain Garciaz, Marie-Anne Hospital, Yves Collette and Norbert Vey
Cancers 2024, 16(6), 1091; https://doi.org/10.3390/cancers16061091 - 8 Mar 2024
Cited by 1 | Viewed by 3987
Abstract
Venetoclax is a BH3-mimetics agent interacting with the anti-apoptotic protein BCL2, facilitating cytochrome c release from mitochondria, subsequent caspases activation, and cell death. Venetoclax combined with azacitidine (VEN-AZA) has become a new standard treatment for AML patients unfit for intensive chemotherapy. In the [...] Read more.
Venetoclax is a BH3-mimetics agent interacting with the anti-apoptotic protein BCL2, facilitating cytochrome c release from mitochondria, subsequent caspases activation, and cell death. Venetoclax combined with azacitidine (VEN-AZA) has become a new standard treatment for AML patients unfit for intensive chemotherapy. In the phase III VIALE-A study, VEN-AZA showed a 65% overall response rate and 14.7 months overall survival in comparison with 22% and 8 months in the azacitidine monotherapy control arm. Despite these promising results, relapses and primary resistance to venetoclax are frequent and remain an unmet clinical need. Clinical and preclinical studies have been conducted to identify factors driving resistance. Among them, the most documented are molecular alterations including IDH, FLT3, TP53, and the newly described BAX mutations. Several non-genetic factors are also described such as metabolic plasticity, changes in anti-apoptotic protein expression, and dependencies, as well as monocytic differentiation status. Strategies to overcome venetoclax resistance are being developed in clinical trials, including triplet therapies with targeted agents targeting IDH, FLT3, as well as the recently developed menin inhibitors or immunotherapies such as antibody–drug conjugated or monoclonal antibodies. A better understanding of the molecular factors driving venetoclax resistance by single-cell analyses will help the discovery of new therapeutic strategies in the future. Full article
(This article belongs to the Section Clinical Research of Cancer)
Show Figures

Figure 1

Figure 1
<p>Mechanism of action of venetoclax. Venetoclax acts as a protein/protein interaction inhibitor that specifically binds the anti-apoptotic protein BCL2 and releases the pro-apoptotic proteins including BAX and BAK that induce mitochondrial outer membrane permeabilization (MOMP) and cell death by intrinsic apoptosis.</p>
Full article ">Figure 2
<p>Main genetic alterations influencing VEN resistance. Intrinsic apoptosis relies on the balance between pro- and anti-apoptotic proteins that will induce mitochondrial outer membrane permeabilization (MOMP). In response to apoptotic stimuli, BAX and BAK form pores in the mitochondrial membrane, inducing mitochondrial outer membrane permeabilization (MOMP). This phenomenon is increased by the members of the BCL2 protein family containing a single BH3 domain named “BH3-only proteins”, which have a pro-apoptotic role (BIM, NOXA, PUMA, BID). Conversely, MOMP is blocked by a series of proteins that have an anti-apoptotic role including BCL2, BCL-XL, and MCL1. <span class="html-italic">IDH</span> mutations are associated with a higher sensitivity to VEN mostly by their role in mitochondrial metabolism and the tricarboxylic acid cycle (TCA). <span class="html-italic">TP53</span> mutations confer resistance to VEN by transcriptionally decreasing pro-apoptotic proteins such as NOXA or PUMA. <span class="html-italic">FLT3</span> mutations are associated with VEN resistance because of increased signalization unbalancing BCL2 and MCL1 dependencies. <span class="html-italic">BAX</span> mutations disrupt the COOH-terminal α9 helix necessary to BAX translocation from the cytosol into the mitochondrial outer membrane, decreasing MOMP.</p>
Full article ">Figure 3
<p>Main classes of drug synergizing with venetoclax in clinical studies. Hypomethylating agents such as azacitidine have been shown to induce an integrated stress response, upregulating the transcription of pro-apoptotic proteins NOXA and PUMA, which increases the efficacy of BCL2 inhibition. IDH inhibitors act by blocking the formation of the oncometabolite <span class="html-italic">D</span>-2-HG in IDH-mutated AML. IDH-mutated AML has been shown to be more susceptible to intrinsic apoptosis due to defects in the TCA cycle. FLT3 mutations induce MCL1 protein upregulation through the RAS/AKT signalization cascade. FLT3 inhibitors are deemed to cause an indirect MCL1 inhibition by degrading MCL1.</p>
Full article ">
21 pages, 4725 KiB  
Article
Maximizing Anticancer Response with MPS1 and CENPE Inhibition Alongside Apoptosis Induction
by Bárbara Pinto, João P. N. Silva, Patrícia M. A. Silva, Daniel José Barbosa, Bruno Sarmento, Juliana Carvalho Tavares and Hassan Bousbaa
Pharmaceutics 2024, 16(1), 56; https://doi.org/10.3390/pharmaceutics16010056 - 29 Dec 2023
Cited by 2 | Viewed by 1455
Abstract
Antimitotic compounds, targeting key spindle assembly checkpoint (SAC) components (e.g., MPS1, Aurora kinase B, PLK1, KLP1, CENPE), are potential alternatives to microtubule-targeting antimitotic agents (e.g., paclitaxel) to circumvent resistance and side effects associated with their use. They can be classified into mitotic blockers, [...] Read more.
Antimitotic compounds, targeting key spindle assembly checkpoint (SAC) components (e.g., MPS1, Aurora kinase B, PLK1, KLP1, CENPE), are potential alternatives to microtubule-targeting antimitotic agents (e.g., paclitaxel) to circumvent resistance and side effects associated with their use. They can be classified into mitotic blockers, causing SAC-induced mitotic arrest, or mitotic drivers, pushing cells through aberrant mitosis by overriding SAC. These drugs, although advancing to clinical trials, exhibit unsatisfactory cancer treatment outcomes as monotherapy, probably due to variable cell fate responses driven by cyclin B degradation and apoptosis signal accumulation networks. We investigated the impact of inhibiting anti-apoptotic signals with the BH3-mimetic navitoclax in lung cancer cells treated with the selective CENPE inhibitor GSK923295 (mitotic blocker) or the MPS1 inhibitor BAY1217389 (mitotic driver). Our aim was to steer treated cancer cells towards cell death. BH3-mimetics, in combination with both mitotic blockers and drivers, induced substantial cell death, mainly through apoptosis, in 2D and 3D cultures. Crucially, these synergistic concentrations were less toxic to non-tumor cells. This highlights the significance of combining BH3-mimetics with antimitotics, either blockers or drivers, which have reached the clinical trial phase, to enhance their effectiveness. Full article
(This article belongs to the Topic Recent Advances in Anticancer Strategies)
Show Figures

Figure 1

Figure 1
<p>CENPE and MPS1 are overexpressed in NSCLC lung cancer cell lines. mRNA expression of CENPE (<b>a</b>) and MPS1 (<b>c</b>) was determined by qRT-PCR in A549 and NCI-H460 cancer cell lines, and was compared to that in non-tumor HPAEpiC cells. Protein levels of CENPE (<b>b</b>) and MPS1 (<b>d</b>) were quantified by Western blotting assay, using α-tubulin as control. Data represent the mean ± SD of three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons 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.</p>
Full article ">Figure 2
<p>GSK923295 + navitoclax and BAY1217389 + navitoclax combinations potentiate cytotoxicity in 2D A549 lung cancer cell cultures. Cell viability (%) of single or combination therapies after 48 h of drug exposure (<b>a</b>,<b>b</b>), from three independent experiments as determined by MTT assay. Synergy scores calculated by the Bliss model of Combenefit software 2.021 with statistical relevance of * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001. Asterisk indicates synergism effects (<b>c</b>,<b>d</b>). Colony formation assays were performed using A549 cells following 7 days (<b>e</b>). Quantification of survival fraction (%) after single or combination treatments as indicated (<b>f</b>,<b>g</b>). Data represent the mean ± SD of three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons 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; 0.0001.</p>
Full article ">Figure 3
<p>Combination of GSK923295 + navitoclax reduces mitotic arrest duration and prevents slippage by accelerating cell death in mitosis in lung cancer cells. Representative phase-contrate microscopy images after 24 h of drugs alone or in combination (<b>a</b>). Quantification of mitotic index; 0.25% DMSO (compound solvent) and 1 μM Nocodazole (mitotic blocker agent) were used as negative and positive controls, respectively (<b>b</b>). Quantification of mitosis duration after respective drug treatments via time-lapse microscopy (<b>c</b>). Cyclin B1 levels as determined by Western blotting assay. (<b>d</b>) Quantification of cell fate (%) for 48 h using different treatments as indicated (<b>e</b>). Representative time-lapse image sequences of A549 cells immediately after drug treatments (<b>f</b>). Data represent the mean ± SD of at least three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons 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; **** <span class="html-italic">p</span> &lt; 0.0001. <span>$</span><span>$</span><span>$</span> difference (<span class="html-italic">p</span> &lt; 0.001) of post-slippage survival cells (%) between untreated and 125 nM GSK923295. <span>$</span><span>$</span><span>$</span><span>$</span> difference (<span class="html-italic">p</span> &lt; 0.0001) of post-slippage survival cells (%) between 125 nM GSK923295 and 125 nM GSK923295 + 1000 nM navitoclax. * difference in death in mitosis cells (%) between 125 nM GSK923295 and 125 nM GSK923295 + 1000 nM navitoclax. #### difference (<span class="html-italic">p</span> &lt; 0.0001) in post-mitotic survival cells (%) between 125 nM GSK923295 and 125 nM GSK923295 + 1000 nM navitoclax.</p>
Full article ">Figure 4
<p>GSK923295 + navitoclax combination enhance A549 lung cancer cell death. Representative cytograms of A549 cell line double stained with Annexin V-FITC and propidium iodide (PI) (<b>a</b>). The quadrants Q are defined as Q1 = live (Annexin V- and PI-negative), Q2 = early stage of apoptosis (Annexin V-positive/PI-negative), Q3 = late stage of apoptosis (Annexin V- and PI-positive) and Q4 = necrosis (Annexin V-negative/PI-positive). Quantification of Annexin-V-positive cells (<b>b</b>). Representative images of A549 apoptotic cells after 48 h treatment, via TUNEL assay to detect DNA fragmentation (green). DNA (blue) was stained with DAPI. Bar, 5 μm (<b>c</b>). Quantification of A549 TUNEL-positive cells (<b>d</b>). Quantification of caspase-9 activity was normalized against the protein content of the extract. Additionally, normalization was performed against the value obtained in the untreated group, setting it as 1 for each assay (<b>e</b>). Data represent the mean ± SD of three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons test. *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Combination of BAY1217389 + navitoclax induces post-mitotic death in 2D lung cancer cell cultures. Quantification of mitosis duration after respective drug treatments via time-lapse microscopy (<b>a</b>). Quantification of cell fate (%) for 48 h using different treatments as indicated (<b>b</b>). Representative time-lapse image sequences of A549 cells immediately after drug treatments (<b>c</b>). Data represent the mean ± SD of at least three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons test. **** <span class="html-italic">p</span> &lt; 0.0001. ns: not significant difference in mitosis duration between 500 nM BAY1217389 and 500 nM BAY1217389 + 1000 nM navitoclax, and in post-slippage survival cells (%) between untreated and 500 nM BAY1217389 and between 500 nM BAY1217389 and 500 nM BAY1217389 + 1000 nM navitoclax. **** difference in post-mitotic death cells (%) between 500 nM BAY1217389 and 500 nM BAY1217389 + 1000 nM navitoclax. #### difference (<span class="html-italic">p</span> &lt; 0.0001) in post-mitotic survival cells (%) between 500 nM BAY1217389 and 500 nM BAY1217389 + 1000 nM navitoclax.</p>
Full article ">Figure 6
<p>BAY1217389 + navitoclax combination enhances A549 lung cancer cell death by apoptosis. Representative cytograms of A549 cells double stained with Annexin V-FITC and propidium iodide (PI) (<b>a</b>). The quadrants Q are defined as Q1 = live (Annexin V- and PI-negative), Q2 = early stage of apoptosis (Annexin V-positive/PI-negative), Q3 = late stage of apoptosis (Annexin V- and PI-positive), and Q4 = necrosis (Annexin V-negative/PI-positive). Quantification of Annexin-V-positive cells (<b>b</b>). Representative images of A549 apoptotic cells after 48 h treatment, via TUNEL assay to detect DNA fragmentation (green). DNA (blue) was stained with DAPI. Bar, 5 μm (<b>c</b>). Quantification of A549 TUNEL-positive cells (<b>d</b>). Quantification of caspase-9 activity was normalized against the protein content of the extract. Additionally, normalization was performed against the value obtained in the untreated group, setting it as 1 for each assay (<b>e</b>). Data represent the mean ± SD of three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons 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; 0.0001.</p>
Full article ">Figure 7
<p>GSK923295 + navitoclax and BAY1217389 + navitoclax combinations enhance cytotoxicity in A549 3D spheroids. Cell viability (%) of single or combination therapies after 48 h of drug exposure (<b>a</b>,<b>d</b>) from three independent experiments as determined by MTT assay. Synergy scores calculated by the Bliss model of Combenefit software with statistical relevance of * <span class="html-italic">p</span> &lt; 0.05. Asterisk indicates synergism effects (<b>b</b>,<b>e</b>). Three-dimensional spheroid viability (%) after 4000 nM, 8000 nM, and 16,000 nM drug concentration treatments (<b>c</b>,<b>f</b>). Data represent the mean ± SD of three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons test. ** <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.</p>
Full article ">Figure 8
<p>Combination of GSK923295 + navitoclax and BAY1217389 + navitoclax potentiates 3D lung cancer spheroid toxicity. Representative images of A549 3D spheroids at days 0 and 2 post-treatment with mono- or combination drugs (100 μm) (<b>a</b>,<b>d</b>). Representative cytograms (<b>b</b>,<b>e</b>) and quantification (<b>c</b>,<b>f</b>) of Annexin V−positive cells after 48 h of drug exposure. The quadrants Q are defined as Q1 = live (Annexin V− and PI−negative), Q2 = early stage of apoptosis (Annexin V−positive/PI−negative), Q3 = late stage of apoptosis (Annexin V− and PI−positive), and Q4 = necrosis (Annexin V−negative/PI−positive). Data represent the mean ± SD of three independent experiments, one-way ANOVA followed by Tukey’s multiple comparisons 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 **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
17 pages, 10759 KiB  
Article
Obatoclax Rescues FUS-ALS Phenotypes in iPSC-Derived Neurons by Inducing Autophagy
by Cristina Marisol Castillo Bautista, Kristin Eismann, Marc Gentzel, Silvia Pelucchi, Jerome Mertens, Hannah E. Walters, Maximina H. Yun and Jared Sterneckert
Cells 2023, 12(18), 2247; https://doi.org/10.3390/cells12182247 - 11 Sep 2023
Cited by 1 | Viewed by 1839
Abstract
Aging is associated with the disruption of protein homeostasis and causally contributes to multiple diseases, including amyotrophic lateral sclerosis (ALS). One strategy for restoring protein homeostasis and protecting neurons against age-dependent diseases such as ALS is to de-repress autophagy. BECN1 is a master [...] Read more.
Aging is associated with the disruption of protein homeostasis and causally contributes to multiple diseases, including amyotrophic lateral sclerosis (ALS). One strategy for restoring protein homeostasis and protecting neurons against age-dependent diseases such as ALS is to de-repress autophagy. BECN1 is a master regulator of autophagy; however, is repressed by BCL2 via a BH3 domain-mediated interaction. We used an induced pluripotent stem cell model of ALS caused by mutant FUS to identify a small molecule BH3 mimetic that disrupts the BECN1-BCL2 interaction. We identified obatoclax as a brain-penetrant drug candidate that rescued neurons at nanomolar concentrations by reducing cytoplasmic FUS levels, restoring protein homeostasis, and reducing degeneration. Proteomics data suggest that obatoclax protects neurons via multiple mechanisms. Thus, obatoclax is a candidate for repurposing as a possible ALS therapeutic and, potentially, for other age-associated disorders linked to defects in protein homeostasis. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Identification of BH3 mimetic compounds reducing the P525L FUS-eGFP SG phenotype in iPSC-derived neurons. (<b>a</b>) Effect on the cell viability of the compounds in iPSC-derived neurons. Compounds were tested at different concentrations for 24 h. Cell viability was assessed with calcein AM-Red. The mean of three independent experiments (n = 3) is shown; error bars indicate the standard error of the mean (SEM). Related to <a href="#app1-cells-12-02247" class="html-app">Figures S1 and S2</a>. (<b>b</b>) Compounds were tested for 24 h at concentrations that were well tolerated. Fluorescent confocal micrographs show that the indicated compounds reduce FUS-eGFP-positive SGs in iPSC-derived neurons treated for 24 h. GA = gambogic acid. Scale bar = 10 μm. (<b>c</b>) Quantification of individual FUS-eGFP-positive SGs from three independent experiments (n = 3); error bars indicate SEM. Significance was tested using the Kruskal-Wallis test with a Dunn post-test. **** denotes the significance of <span class="html-italic">p</span> &lt; 0.0001 between unstressed and arsenite. **** denotes the significance between treatments and arsenite. *, ** and **** indicate <span class="html-italic">p</span> &lt; 0.5, &lt;0.01, and &lt;0.0001, respectively. Related to <a href="#app1-cells-12-02247" class="html-app">Figures S1–S5</a>.</p>
Full article ">Figure 2
<p>Obatoclax is associated with ALS-associated phenotypes in P525L FUS-eGFP iPSC-derived neurons. (<b>a</b>) Levels of cytoplasmic and nuclear FUS-eGFP in iPSC-derived neurons treated with DMSO and obatoclax at 10 nM for 24 h. The mean of six independent experiments (n = 6) is shown; error bars indicate SEM. Treatments were analyzed via an unpaired student’s <span class="html-italic">t</span>-test. * indicates <span class="html-italic">p</span> &lt; 0.05. n.s. indicates not significant. Scale bar = 20 μm. (<b>b</b>) Cleaved caspase 3 (CC3) levels in iPSC-derived neurons treated with DMSO and obatoclax at 10 nM for 24 h. The mean of six independent experiments (n = 6) is shown; error bars indicate SEM. Treatments were analyzed via an unpaired student’s t-test. * indicates <span class="html-italic">p</span> &lt; 0.05. Scale bar = 20 μm. Related to <a href="#app1-cells-12-02247" class="html-app">Figure S6</a>.</p>
Full article ">Figure 3
<p>Obatoclax restores protein homeostasis in P525L FUS-eGFP iPSC-derived neurons. p62 protein levels in iPSC-derived neurons treated with DMSO and obatoclax at 10 nM for 24 h. The mean of three independent experiments (n = 3) is shown; error bars indicate SEM. Treatments were analyzed via an unpaired student’s <span class="html-italic">t</span>-test. ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Obatoclax induces autophagy in P525L FUS e-GFP iPSC-derived neurons. (<b>a</b>) Obatoclax was tested at 10 nM for 6, 9, 24, and 48 hours. We assessed the LC3B-II protein levels using Western blotting. Results show the mean of six independent experiments (n = 6); error bars indicate SEM. Treatments were analyzed via one-way ANOVA with a Dunnett post-test, * indicates <span class="html-italic">p</span> &lt; 0.5; n.s. indicates not significant. (<b>b</b>) Obatoclax was tested at 10 nM for 24 hours with and without Bafilomycin A at 10 nM for 24h. We assessed the LC3B-II protein levels using Western blotting. Results show the mean of three independent experiments (n = 3); error bars indicate SEM. Treatments were analyzed via one-way ANOVA with a Dunnet post-test; * indicates <span class="html-italic">p</span> &lt; 0.05. ** indicates <span class="html-italic">p</span> &lt; 0.01. n.s. indicates not significant. BafA1 = Bafilomycin A1. (<b>c</b>) Obatoclax was tested at 10 nM for 24 hours with and without Bafilomycin A at 10 nM for 48 hours. We assessed the LC3B-II protein levels using Western blotting. Results show the mean of five independent experiments (n = 5); error bars indicate SEM. Treatments were analyzed via one-way ANOVA with a Dunnet post-test, ** indicates <span class="html-italic">p</span> &lt; 0.01. n.s., not significant. BafA1 = Bafilomycin A1. (<b>d</b>) Obatoclax was tested at 10 nM for 24 hours. We assessed the LAMP1 protein using Western blotting. Results show the mean of eight independent experiments (n = 8); error bars indicate SEM. n.s. indicates not significant. Related to <a href="#app1-cells-12-02247" class="html-app">Figures S7 and S8</a>.</p>
Full article ">Figure 5
<p>Obatoclax induces autophagy in P525L FUS-eGFP iPSC-derived neurons. Confocal fluorescent micrographs showing LC3 puncta in P525L FUS-eGFP iPSC-derived neurons treated with DMSO and obatoclax at 10 nM for 24 hours. Obatoclax was tested at 10 nM for 24 hours. We assessed the LC3B-II protein levels using Western blotting. The mean of three independent experiments (n = 3); error bars indicate SEM. Treatments were analyzed via an unpaired student’s <span class="html-italic">t</span>-test. * indicates <span class="html-italic">p</span> &lt; 0.05. Scale bar = 10 μm.</p>
Full article ">Figure 6
<p>Obatoclax disrupts the BECN1-BCL2 complex in iPSC-derived neurons. Proximity ligation assay to assess the interaction of BECN1 and BCL2 in iPSC-derived neurons treated with DMSO and obatoclax at 10 nM for 24 h. The mean of eight independent experiments (n = 8); error bars indicate SEM. Treatments were analyzed via an unpaired Student’s <span class="html-italic">t</span>-test. * indicates <span class="html-italic">p</span> &lt; 0.05. Scale bar = 10 μm.</p>
Full article ">Figure 7
<p>Proteomic analysis of iPSC-derived neurons. (<b>a</b>) Principle component analysis (PCA) plot of the proteome of the samples analyzed: WT FUS-eGFP treated with DMSO, P525L FUS-eGFP treated with DMSO, and P525L FUS-eGFP treated with obatoclax at 10 nM for 24 h. Three independent experiments. See also <a href="#app1-cells-12-02247" class="html-app">Tables S1 and S2</a>. (<b>b</b>) Volcano plot of the comparison between WT FUS-eGFP treated with DMSO and P525L FUS-eGFP treated with DMSO. (<b>c</b>) Volcano plot of the comparison between P525L FUS treated with DMSO and P525L FUS-eGFP treated with obatoclax at 10 nM for 24 h. (<b>d</b>) Differentially expressed proteins related to the proteasome pathway. Proteasome subunit alpha type-1 (PSMA1, <span class="html-italic">p</span> = 0.000103), proteasome activator complex subunit 1 (PSME1, <span class="html-italic">p</span> = 0.000251), proteasome activator complex subunit 2 (PSME2, <span class="html-italic">p</span> = 0.000132), and proteasome activator complex subunit 3 (PSME3, <span class="html-italic">p</span> = 0.000054) are downregulated in P525L FUS-eGFP compared with WT FUS-eGFP. (<b>e</b>) Differentially expressed proteins related to the mucin type O-glycan biosynthesis pathway. Lactosylceramide alpha-2,3-sialyltransferase (ST3GAL1, <span class="html-italic">p</span> = 0.005826) and beta-1,4-galactosyltransferase 5 (B4GALT5, <span class="html-italic">p</span> = 0.000386) are upregulated in P525L FUS-eGFP treated with obatoclax compared with P525L FUS-eGFP treated with DMSO. (<b>f</b>) Proteins related to endocytic trafficking. EH domain-containing protein 4 (EHD4, 0.000021) and AP-1 complex subunit sigma-2 (AP1S1, <span class="html-italic">p</span> = 0.004056) are upregulated in P525L FUS-eGFP treated with obatoclax compared with P525L FUS-eGFP treated with DMSO. (<b>g</b>) MRG/MORF4L-binding protein (MRGBP, <span class="html-italic">p</span> = 0.000154) related to DNA double-strand break repair. (<b>h</b>) DnaJ homolog subfamily C member 9 (DNAJC9, <span class="html-italic">p</span> = 0.001442) related to histone chaperone network and heat shock-induced response. Related to <a href="#app1-cells-12-02247" class="html-app">Tables S1 and S2</a>.</p>
Full article ">
16 pages, 7121 KiB  
Article
HSP90 Inhibitor PU-H71 in Combination with BH3-Mimetics in the Treatment of Acute Myeloid Leukemia
by Katja Seipel, Scarlett Kohler, Ulrike Bacher and Thomas Pabst
Curr. Issues Mol. Biol. 2023, 45(9), 7011-7026; https://doi.org/10.3390/cimb45090443 - 23 Aug 2023
Cited by 2 | Viewed by 1444
Abstract
Targeting the molecular chaperone HSP90 and the anti-apoptotic proteins MCL1 and BCL2 may be a promising novel approach in the treatment of acute myeloid leukemia (AML). The HSP90 inhibitor PU-H71, MCL1 inhibitor S63845, and BCL2 inhibitor venetoclax were assessed as single agents and [...] Read more.
Targeting the molecular chaperone HSP90 and the anti-apoptotic proteins MCL1 and BCL2 may be a promising novel approach in the treatment of acute myeloid leukemia (AML). The HSP90 inhibitor PU-H71, MCL1 inhibitor S63845, and BCL2 inhibitor venetoclax were assessed as single agents and in combination for their ability to induce apoptosis and cell death in leukemic cells. AML cells represented all major morphologic and molecular subtypes including FLT3-ITD and TP53 mutant AML cell lines and a variety of patient-derived AML cells. Results: PU-H71 and combination treatments with MCL1 inhibitor S63845 or BCL2 inhibitor venetoclax induced cell cycle arrest and apoptosis in susceptible AML cell lines and primary AML. The majority of the primary AML samples were responsive to PU-H71 in combination with BH3 mimetics. Elevated susceptibility to PU-H71 and S63845 was associated with FLT3 mutated AML with CD34 < 20%. Elevated susceptibility to PU-H71 and venetoclax was associated with primary AML with CD117 > 80% and CD11b < 45%. The combination of HSP90 inhibitor PU-H71 and MCL1 inhibitor S63845 may be a candidate treatment for FLT3-mutated AML with moderate CD34 positivity while the combination of HSP90 inhibitor PU-H71 and BCL2 inhibitor venetoclax may be more effective in the treatment of primitive AML with high CD117 and low CD11b positivity. Full article
(This article belongs to the Special Issue Advances in Molecular Pathogenesis Regulation in Cancer, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>PU-H71 dose response in AML cell lines. AML cells were treated with the HSP90 inhibitor PU-H71 at the indicated dosages for 20 h. Cell viability data are average values of multiple repeat measurements per dosage.</p>
Full article ">Figure 2
<p>Susceptibility of AML cell lines to various treatment combinations. Cell viability was determined in AML cell lines MOLM-13 (<b>A</b>), ML-2 (<b>B</b>), OCI-AML3 (<b>C</b>), SKM-1 (<b>D</b>), MOLM-16 (<b>E</b>) and PL-21 (<b>F</b>) after 20 h of treatment with single compounds and in combination with 0.1–0.3 μM PU-H71 (PU) and 0.l μM S63845 (S0.1), 0.1 μM venetoclax (VC0.1), or 1 μM venetoclax (VC1). Significance of differences denoted for <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 3
<p>Induction of cell cycle arrest, gene expression, and protein degradation in AML cells treated with HSP90 inhibitor PU-H71 (PU). Cytometric analysis of AML cell lines MOLM13 (<b>A</b>), OCI-AML3 (<b>B</b>), and SKM-1 (<b>C</b>) after 20 h treatment with 100–300 nM PU-H71. According to DAPI staining, cell intensities were classified as the G0/G1 (2N) or G2 (4N) phase. Relative quantitation of <span class="html-italic">CDKN1A</span> (<b>D</b>) and <span class="html-italic">TP53</span> (<b>E</b>) gene expression. Relative quantitation of FLT3, BCL2, and MCL1 protein levels in MOLM-13 cells treated with 100–200 nM PU-H71 (<b>F</b>). Significance of differences denoted for <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 4
<p>Induction of apoptosis and cell death in AML cells treated with PU-H71 in combination with S63845 or venetoclax. (<b>A</b>–<b>C</b>): Cytometric analysis of MOLM-13 cells after 20 h treatment with 50–100 nM PU-H71 and 5–10 nM S63845 (S) and stained with annexin-V and PI (<b>A</b>,<b>B</b>) or DAPI (<b>C</b>). (<b>D</b>–<b>F</b>): Cytometric analysis of MOLM-13 cells treated with 100–200 nM PU-H71 (PU) and 5–25 nM venetoclax (VC) and stained with annexin-V (<b>D</b>,<b>E</b>) or DAPI (<b>F</b>). According to 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). According to DAPI staining, cell intensities were classified as subG1 (&lt;2N), G0/G1 (2N), S phase (2–4N) or G2 phase (4N). Significance of differences denoted for <span class="html-italic">p</span> &lt; 0.001 (***); and <span class="html-italic">p</span> &lt; 0.0001 (****).</p>
Full article ">Figure 5
<p>Hematological cells’ in vitro responses to PU-H71, S63845, and venetoclax. Cell viability was determined in mononuclear cells isolated from AML patients (A1-A27) and healthy donors (H1-H5) peripheral blood or bone marrow after 20 h treatment with 100 nM PU-H71 (PU), 100 nM S63845 (S), PU-H71 combined with S63845 (PU-S) (<b>A</b>), or PU-H71 combined with venetoclax (PU-VC) (<b>B</b>). The patient samples were sorted into response groups, namely substantial response (SR), intermediate response (IR), and minor (normal) response (NR) compared to healthy donors, treated with PU-H71 and S63845 (<b>C</b>), PU-H71 and venetoclax (<b>D</b>), 100 nM PU-H71 (<b>E</b>), 100 nM S63845 (<b>F</b>), and 100 nM venetoclax (<b>G</b>). The significance of differences in median values was calculated by Mann–Whitney test. Significance denoted for <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>FLT3 status and blast cell percentage as biomarkers of responses to PU-H71 and BH3 mimetic combination treatments. Cell viability was determined in mononuclear cells isolated from AML patients’ peripheral blood or bone marrow after 20 h of treatment. Samples were grouped according to FLT3 genetic variation (<b>A</b>–<b>E</b>) or blast cell percentage (<b>F</b>,<b>G</b>). AML cells were treated in vitro with 100 nM PU-H71 (<b>A</b>,<b>F</b>), 100 nM S63845 (<b>B</b>,<b>G</b>), 100 nM venetoclax (<b>C</b>,<b>H</b>), PU-H71 combined with S63845 (<b>D</b>,<b>I</b>), or PU-H71 combined with venetoclax (<b>E</b>,<b>J</b>). The significance of differences in median values was calculated by a Mann–Whitney test. Significance of differences in median values was calculated by a Mann–Whitney test. Significance denoted for <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 (***); no significance denoted for <span class="html-italic">p</span> &gt; 0.05 (ns).</p>
Full article ">Figure 7
<p>CD34, CD117, and CD11b biomarkers of responses to PU-H71 and BH3 mimetic combination treatments. Cell viability was determined in mononuclear cells isolated from AML patients’ peripheral blood or bone marrow after 20 h of treatment. Samples were grouped according to CD34 (<b>A</b>–<b>E</b>), CD117 (<b>F</b>–<b>J</b>), or CD11b (<b>K</b>–<b>O</b>) expression. Cells were treated in vitro with 100 nM PU-H71 (<b>A</b>,<b>F</b>,<b>K</b>), 100 nM S63845 (<b>B</b>,<b>G</b>,<b>L</b>), 100 nM venetoclax (<b>C</b>,<b>H</b>,<b>M</b>), PU-H71 combined with S63845 (<b>D</b>,<b>I</b>,<b>N</b>), or PU-H71 combined with venetoclax (<b>E</b>,<b>J</b>,<b>O</b>). Significance of differences in median values was calculated by Mann–Whitney test. Significance denoted for <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 (***); no significance denoted for <span class="html-italic">p</span> &gt; 0.05 (ns).</p>
Full article ">Figure 8
<p>Schematic presentation of the FLT3 and c-KIT (CD117) induced signaling pathways and downstream effects. FLT3-ITD and FLT3-TKD are constitutively active growth factor receptors. C-KIT is an inducible growth factor receptor activated by stem cell factor (SCF) binding. Both tyrosine receptor kinases activate PI3K-AKT, RAS-MEK-ERK, and STAT5 leading to cell growth and proliferation via inhibition of the tumor suppressor TP53 and induction of the apoptosis regulators MCL1 and BCL2. HSP90 protein can bind and stabilize client proteins including AKT, BCL2, FLT3, JAK, MDM2, STAT5, SHP2, and BRAF. HSP90 proteins are indicated in yellow, oncogenic protein functions in red, tumor suppressor functions in green ovals, and targeted inhibitors in blue rectangles. Sharp arrows and blunt arrows indicate target induction and inhibition, respectively.</p>
Full article ">
21 pages, 3769 KiB  
Article
A New Quinone-Based Inhibitor of Mitochondrial Complex I in D-Conformation, Producing Invasion Reduction and Sensitization to Venetoclax in Breast Cancer Cells
by Matías Monroy-Cárdenas, Víctor Andrades, Cristopher Almarza, María Jesús Vera, Jorge Martínez, Rodrigo Pulgar, John Amalraj, Ramiro Araya-Maturana and Félix A. Urra
Antioxidants 2023, 12(8), 1597; https://doi.org/10.3390/antiox12081597 - 10 Aug 2023
Cited by 5 | Viewed by 1738
Abstract
Mitochondrial Complex I plays a crucial role in the proliferation, chemoresistance, and metastasis of breast cancer (BC) cells. This highlights it as an attractive target for anti-cancer drugs. Using submitochondrial particles, we identified FRV–1, an ortho-carbonyl quinone, which inhibits NADH:duroquinone activity in [...] Read more.
Mitochondrial Complex I plays a crucial role in the proliferation, chemoresistance, and metastasis of breast cancer (BC) cells. This highlights it as an attractive target for anti-cancer drugs. Using submitochondrial particles, we identified FRV–1, an ortho-carbonyl quinone, which inhibits NADH:duroquinone activity in D-active conformation and reduces the 3ADP state respiration dependent on Complex I, causing mitochondrial depolarization, ATP drop, increased superoxide levels, and metabolic remodeling towards glycolysis in BC cells. Introducing methyl groups at FRV–1 structure produced analogs that acted as electron acceptors at the Complex I level or increased the inhibitory effect of FCCP-stimulated oxygen consumption rate, which correlated with their redox potential, but increased toxicity on RMF-621 human breast fibroblasts was observed. FRV–1 was inactive in the naphthoquinone oxidoreductase 1 (NOQ1)-positive BC cell line, MCF7, but the sensitivity was recovered by dicoumarol, a NOQ1 inhibitor, suggesting that FRV–1 is a NOQ1 substrate. Importantly, FRV–1 selectively inhibited the proliferation, migration, and invasion of NQO1 negative BC cell, MDA-MB-231, in an OXPHOS- and ROS-dependent manner and sensitized it to the BH3 mimetic drug venetoclax. Overall, FRV–1 is a novel Complex I inhibitor in D-active conformation, blocking possibly the re-activation to A-state, producing selective anti-cancer effects in NQO1-negative BC cell lines. Full article
(This article belongs to the Section Health Outcomes of Antioxidants and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Synthesis of compounds.</p>
Full article ">Figure 2
<p>Effect of rotenone and FRV–1 on Active (A-state) and Deactive (D-state) Complex I activity of submitochondrial particles (SMP) from breast cancer cells. (<b>A</b>) Current model on A/D transition of Complex I, (<b>B</b>,<b>C</b>) Effect of rotenone (2.5 µM) and (<b>D</b>,<b>E</b>) FRV–1 (50 µM) on A- and D-state Complex I activity (NADH: duroquinone oxidoreductase) in SMP from breast cancer TA3/Ha cells. 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; 0.001, vs. Control (DMSO) and n.s.: not significant.</p>
Full article ">Figure 3
<p>Effects of ortho-carbonyl substituted quinones and hydroquinones on submitochondrial particles from breast cancer cells. (<b>A</b>) Chemical structures of ortho-carbonyl substituted quinones and hydroquinones. (<b>B</b>,<b>C</b>) Effect of quinones and hydroquinones on NADH:duroquinone activity. (<b>D</b>–<b>F</b>) Inhibition of NADH:duroquinone, NADH:juglone, and NADH:FeCN reductase activity by FRV–1 (100 µM), (<b>G</b>) NADH:FRV–1, and NADH:FRV–2 oxidoreductase activity. (<b>H</b>) OD spectrum of FRV–2 and FRV–4 (100 µM) in buffer assay, (<b>I</b>) OD spectrum of hydroquinone FRV–4 formation (peak: 385 nm) in buffer assay containing SMP, FRV–2, and NADH. (<b>J</b>) Quantification of FRHV–2 formation in SMP treated with FRV–2. The data shown are the mean ± SD of three independent experiments. *** <span class="html-italic">p</span> &lt; 0.001, vs. Control (DMSO) and n.s.: not significant.</p>
Full article ">Figure 4
<p>Effect of FRV–1 and FRHV–1 on tumor mitochondrial respiration. (<b>A</b>,<b>B</b>) Effect of compounds on mitochondrial respiration dependent on each respiratory complex in ADP presence (state 3ADP). (<b>C</b>) Effect on compounds on Complex I-dependent respiratory control ratio (RCR). (<b>D</b>) Effect of FRV–1 on Complex I-dependent respiration in states 4o and 3u, in the presence of 1 μM oligomycin and 0.2 μM CCCP, respectively. 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 (DMSO).</p>
Full article ">Figure 5
<p>FRV–1, but no FRHV–1, reduces the mitochondrial bioenergetics and inhibits the proliferation of TA3/Ha cells by ROS production. Effects of FRV–1 and FRHV–1 on (<b>A</b>) NAD(P)H levels, (<b>B</b>) Δψm, (<b>C</b>) ATP levels, (<b>D</b>) mitochondrial superoxide, and (<b>E</b>) intracellular ROS levels at 2 h of treatment. (<b>F</b>) Effect of FRV–1 and FRHV–1 (50 µM) on proliferation and (<b>G</b>) cell cycle distribution levels. (<b>H</b>,<b>I</b>) N-acetylcysteine (NAC) partially prevents the effect of FRV–1 on NAD(P)H and proliferation. Data are expressed as means ± 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 (DMSO), and n.s.: not significant, Rot: rotenone, Olig: oligomycin.</p>
Full article ">Figure 6
<p>FRV–1 inhibits mitochondrial respiration, triggering a metabolic shift toward glycolysis in MDA-MB-231 cells. (<b>A</b>) Effect of FRV–1 on mitochondrial respiration when injected, reducing the (<b>B</b>) basal OCR, (<b>C</b>) ATP-driven OCR, and (<b>D</b>) maximal OCR. (<b>E</b>) Effect of FRV–1 on the profile of respiration and (<b>F</b>) OCR/ECAR ratio at 4 h of treatment. (<b>G</b>) Differential effect of FRV–1 on oxidative (galactose) and glycolytic (glucose) subpopulations of MDA-MB-231 cells. (<b>H</b>) Chemical structures of methylated analogs of FRV–1 and FRV–2. (<b>I</b>) Effect of methylated analogs on mitochondrial OCR at 50 µM. (<b>J</b>) Cyclic voltagramme for all compounds (2.0 mM) at 100 mV/s, (<b>K</b>,<b>L</b>) OCR-E<sub>R1</sub> and (<b>M</b>,<b>N</b>) OCR-E<sub>R2</sub> relationships for FRV–1 and FRV–2 and analogs. 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.</p>
Full article ">Figure 7
<p>FRV–1 and analogs reduce the viability and produce selective sensitization to BH3 mimetic drug ABT–199 in MDA-MB-231 cells. (<b>A</b>) Effect of FRV–1 and analogs on the viability of breast cancer epithelial and stromal cell lines at 48 h of treatment. (<b>B</b>) Effect of dicoumarol, an NQO1 inhibitor, on the viability of MCF7 cells at 48 h of exposition. (<b>C</b>) Dependence of NOQ1 on the effect of FRV–1 in MCF7. Cells were exposed to NQO1 inhibitor dicumarol (Dic) 1 h before treatment with quinone and viability was evaluated at 48 h of treatment. (<b>D</b>–<b>F</b>) Effect of combination FRV–1 and analogs (50 µM) with ABT–199 (10 µM). 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, and n.s.: not significant.</p>
Full article ">Figure 8
<p>FRV–1 inhibits the migration and invasion in triple-negative breast cancer cells. (<b>A</b>) Oxygen consumption rate (OCR) profile of MDA-MB-231 WT (OXPHOS functional) and ρ0 cells (OXPHOS lacking), (<b>B</b>) basal, (<b>C</b>) maximal OCR (in the presence of FCCP), (<b>D</b>) extra-cellular acidification rate (ECAR) profile of MDA-MB-231 WT and ρ0 cells, (<b>E</b>) glycolysis, and (<b>F</b>) glycolytic capacity. (<b>G</b>,<b>H</b>) Effect of FRV–1 (25 µM) on the migration of MDA-MB-231 WT and ρ0 cells, (<b>I</b>) zymography for MMP activity in MDA-MB-231 WT cells treated with FRV–1 (25 µM), mitoTEMPO (1 µM) and combination and (<b>J</b>) invasion of MDA-MB-231 WT 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, and n.s.: not significant.</p>
Full article ">
20 pages, 3685 KiB  
Review
T-Cell Prolymphocytic Leukemia: Diagnosis, Pathogenesis, and Treatment
by Marc Gutierrez, Patrick Bladek, Busra Goksu, Carlos Murga-Zamalloa, Dale Bixby and Ryan Wilcox
Int. J. Mol. Sci. 2023, 24(15), 12106; https://doi.org/10.3390/ijms241512106 - 28 Jul 2023
Cited by 3 | Viewed by 4735
Abstract
T-cell prolymphocytic leukemia (T-PLL) is a rare and aggressive neoplasm of mature T-cells. Most patients with T-PLL present with lymphocytosis, anemia, thrombocytopenia, and hepatosplenomegaly. Correct identification of T-PLL is essential because treatment for this disease is distinct from that of other T-cell neoplasms. [...] Read more.
T-cell prolymphocytic leukemia (T-PLL) is a rare and aggressive neoplasm of mature T-cells. Most patients with T-PLL present with lymphocytosis, anemia, thrombocytopenia, and hepatosplenomegaly. Correct identification of T-PLL is essential because treatment for this disease is distinct from that of other T-cell neoplasms. In 2019, the T-PLL International Study Group (TPLL-ISG) established criteria for the diagnosis, staging, and assessment of response to treatment of T-PLL with the goal of harmonizing research efforts and supporting clinical decision-making. T-PLL pathogenesis is commonly driven by T-cell leukemia 1 (TCL1) overexpression and ATM loss, genetic alterations that are incorporated into the TPLL-ISG diagnostic criteria. The cooperativity between TCL1 family members and ATM is seemingly unique to T-PLL across the spectrum of T-cell neoplasms. The role of the T-cell receptor, its downstream kinases, and JAK/STAT signaling are also emerging themes in disease pathogenesis and have obvious therapeutic implications. Despite improved understanding of disease pathogenesis, alemtuzumab remains the frontline therapy in the treatment of naïve patients with indications for treatment given its high response rate. Unfortunately, the responses achieved are rarely durable, and the majority of patients are not candidates for consolidation with hematopoietic stem cell transplantation. Improved understanding of T-PLL pathogenesis has unveiled novel therapeutic vulnerabilities that may change the natural history of this lymphoproliferative neoplasm and will be the focus of this concise review. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Peripheral blood involvement by T-PLL. Representative images demonstrate that T-PLL tumor cells (Wright stain) can be variable in size and range from medium (<b>right panel</b>) to large (<b>left panel</b>), with irregular nuclear contours, condensed chromatin, and prominent nucleoli. The morphological and cytological features of T-PLL have been previously documented and described [<a href="#B2-ijms-24-12106" class="html-bibr">2</a>,<a href="#B6-ijms-24-12106" class="html-bibr">6</a>,<a href="#B7-ijms-24-12106" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>B</b>) Involvement of bone marrow with T-PLL. (<b>A</b>) Highlights of interstitial involvement by tumor cells (blue arrows). (<b>B</b>) The most common morphological pattern is composed of medium lymphocytes with round nuclear contours and central prominent nucleoli (blue arrows). (<b>C</b>) Focal aggregates of tumor cells are usually observed (dotted circles). (<b>D</b>) Small-cell morphology variant is characterized by small to medium forms with condensed chromatin (blue arrows). Occasional large “cerebriform” lymphocytes are also observed (red arrows). The morphological features characteristic of T-PLL bone marrow involvement have been previously described [<a href="#B2-ijms-24-12106" class="html-bibr">2</a>,<a href="#B6-ijms-24-12106" class="html-bibr">6</a>,<a href="#B7-ijms-24-12106" class="html-bibr">7</a>].</p>
Full article ">Figure 3
<p>Skin involvement by T-PLL (black arrows). The tumor cells are predominantly distributed in a perivascular fashion, with no epidermotropism. Characteristic histological features of dermal involvement by T-PLL have been previously described [<a href="#B2-ijms-24-12106" class="html-bibr">2</a>,<a href="#B6-ijms-24-12106" class="html-bibr">6</a>,<a href="#B7-ijms-24-12106" class="html-bibr">7</a>].</p>
Full article ">Figure 4
<p>Therapeutic pathways in T-PLL. While alemtuzumab is the current frontline therapy, an improved understanding of T-PLL pathogenesis is leading to potentially targetable vulnerabilities. These include attempts to reinstate p53 function through MDM2 inhibitors as well as exploitation of faulty double strand break repair via PARP inhibitors, inhibition of TCR activation through ITK inhibitors as well as AKT/PI3K inhibitors, JAK inhibitors which inhibit cell growth and differentiation through the JAK-STAT pathway, BH3 mimetics to promote BAX and BAK mediated cell death, epigenetic approaches including hypomethylating agents and HDAC inhibitors, and novel antibody targeting e.g., CCR7. The illustration was created using Biorender.com and is adapted from previous review articles [<a href="#B47-ijms-24-12106" class="html-bibr">47</a>,<a href="#B72-ijms-24-12106" class="html-bibr">72</a>,<a href="#B73-ijms-24-12106" class="html-bibr">73</a>].</p>
Full article ">
23 pages, 25022 KiB  
Article
MCL1 Inhibition Overcomes the Aggressiveness Features of Triple-Negative Breast Cancer MDA-MB-231 Cells
by Giovanni Pratelli, Daniela Carlisi, Diana Di Liberto, Antonietta Notaro, Michela Giuliano, Antonella D’Anneo, Marianna Lauricella, Sonia Emanuele, Giuseppe Calvaruso and Anna De Blasio
Int. J. Mol. Sci. 2023, 24(13), 11149; https://doi.org/10.3390/ijms241311149 - 6 Jul 2023
Cited by 3 | Viewed by 1677
Abstract
Triple-Negative Breast Cancer (TNBC) is a particularly aggressive subtype among breast cancers (BCs), characterized by anoikis resistance, high invasiveness, and metastatic potential as well as Epithelial–Mesenchymal Transition (EMT) and stemness features. In the last few years, our research focused on the function of [...] Read more.
Triple-Negative Breast Cancer (TNBC) is a particularly aggressive subtype among breast cancers (BCs), characterized by anoikis resistance, high invasiveness, and metastatic potential as well as Epithelial–Mesenchymal Transition (EMT) and stemness features. In the last few years, our research focused on the function of MCL1, an antiapoptotic protein frequently deregulated in TNBC. Here, we demonstrate that MCL1 inhibition by A-1210477, a specific BH3-mimetic, promotes anoikis/apoptosis in the MDA-MB-231 cell line, as shown via an increase in proapoptotic markers and caspase activation. Our evidence also shows A-1210477 effects on Focal Adhesions (FAs) impairing the integrin trim and survival signaling pathways, such as FAK, AKT, ERK, NF-κB, and GSK3β-inducing anoikis, thus suggesting a putative role of MCL1 in regulation of FA dynamics. Interestingly, in accordance with these results, we observed a reduction in migratory and invasiveness capabilities as confirmed by a decrease in metalloproteinases (MMPs) levels following A-1210477 treatment. Moreover, MCL1 inhibition promotes a reduction in EMT characteristics as demonstrated by the downregulation of Vimentin, MUC1, DNMT1, and a surprising re-expression of E-Cadherin, suggesting a possible mesenchymal-like phenotype reversion. In addition, we also observed the downregulation of stemness makers such as OCT3/4, SOX2, NANOG, as well as CD133, EpCAM, and CD49f. Our findings support the idea that MCL1 inhibition in MDA-MB-231 could be crucial to reduce anoikis resistance, aggressiveness, and metastatic potential and to minimize EMT and stemness features that distinguish TNBC. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>A-1210477 reduces cell viability of MDA-MB-231 cells in a dose- and time-dependent manner. (<b>A</b>) MDA-MB-231 cells were treated with increasing doses of A-1210477 (5–7.5–10–15 µM) for different times of incubation. Then, cell viability was analyzed via MTT assay and expressed as the percentage with respect to control cells. (<b>B</b>) Phase-contrast microscopy images of MDA-MB-231 cells showing the effects of A-1210477 treatment on cell morphology (magnification 200×). All the experiments were carried out in triplicate. Data are expressed as 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.</p>
Full article ">Figure 2
<p>Cell death effects induced by A-1210477 in MDA-MB-231 cells. (<b>A</b>) Immunoblot images show anoikis and apoptosis markers. After treatment for 72 h with 10 µM A-1210477, 20 µg of whole cell lysates were submitted to Western blot analyses, using specific antibodies as described in the <a href="#sec4-ijms-24-11149" class="html-sec">Section 4</a>. GAPDH was used as a loading control. Images are representative of at least three independent experiments. Data are expressed as mean ± SD in fold changes and are reported in the histogram. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Fluorescence microscopy images of cells stained with Hoechst 33342 dye evidencing chromatin condensation and fragmentation (magnification 200×).</p>
Full article ">Figure 3
<p>A-1210477 treatment induces changes in integrin expression, FA destruction, and kinase signaling interruption. Western blot analyses show the change in levels of peculiar subunits’ integrins involved in anoikis: (<b>A</b>) FA proteins (<b>B</b>) and survival kinases AKT and ERK (<b>C</b>). All the effects are reached after 72 h of treatment with 10 µM A-1210477. GAPDH was used as an internal control. Images are representative of at least three independent experiments. Data are expressed as mean ± SD in fold changes and are reported in the histogram. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Effects of treatment with A-1210477 on migratory and invasive capabilities of MDA-MB-231 cells. Representative images of phase contrast microscopy reporting scratch wound healing (<b>A</b>) and Transwell invasion assay (<b>B</b>). After 48 h treatment, cells that migrated to the underside of the insert were stained with Cristal Violet (magnification 200×). Graph summarizes the relative percentage of cells migrating in treated cells with respect to control cells. (<b>C</b>) MMP evaluation after A-1210477 treatment. Using specific antibodies, 20 µg of whole cell lysates were submitted to Western blot analyses. GAPDH was used as a loading control. Data represent the mean ± SD. Images are representative 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.</p>
Full article ">Figure 5
<p>A-1210477 treatment reduces stemness markers in MDA-MB-231 cells. (<b>A</b>) Cytofluorimetric analyses show cell-surface expression of stemness markers. After A-1210477 treatment, cells were incubated with appropriate conjugated antibodies (<a href="#app1-ijms-24-11149" class="html-app">Table S1</a>) and analyzed via FACS. The histogram plots report the relative expression of CD44<sup>+</sup>/CD24<sup>−/low</sup>, CD49f, and CD326-positive cells in treated cells with respect to the control. (<b>B</b>) qRT-PCR analysis of gene stemness-related <span class="html-italic">SOX2</span>, <span class="html-italic">NANOG</span>, and <span class="html-italic">OCT3/4</span>. (<b>C</b>) Stemness marker CD133 Western blot analysis. Gene and protein expression was normalized with GAPDH and expressed as fold changes. All the experiments were performed in triplicate. Data are expressed as 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.</p>
Full article ">Figure 6
<p>EMT expression markers in MDA-MB-231 cells after A-1210477 treatment. (<b>A</b>) E-Cadherin, N-Cadherin, Vimentin, MUC1, and Cytokeratin-19 Western blot analyses in MDA-MB-231 cells treated with A-1210477 with respect to control cells. (<b>B</b>) qRT-PCR of <span class="html-italic">E-Cadherin</span> in MDA-MB-231. Gene and protein expression was normalized with GAPDH and expressed as fold changes. All the experiments were performed in triplicate. Data are expressed as 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.</p>
Full article ">Figure 7
<p>E-Cadherin de novo expression mediated by GSK3β activity and DNMT1 degradation. (<b>A</b>) DNMT1 and GSK3β Western blot analysis. The blot image shows a reduction in DNMT1 levels and a decrease in phosphorylated-form GSK3β. (<b>B</b>) qRT-PCR of <span class="html-italic">DNMT1</span> displays a non-significative change in mRNA expression suggesting protein degradation. Gene and protein expression was normalized with GAPDH and expressed as fold changes. All the experiments were performed in triplicate. Data are expressed as mean ± SD. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>NF-κB and STAT3 phosphorylated forms reduction in MDA-MB-231 cells by A-1210477 treatment promotes MCL1 downregulation. (<b>A</b>) Western blot analyses of NF-κB and STAT3 phosphorylation. (<b>B</b>) qRT-PCR of <span class="html-italic">MCL1</span> expression levels in MDA-MB-231 treated cells with respect to the control cells. Gene and protein expression was normalized with GAPDH and expressed as fold changes. All the experiments were performed in triplicate. Data are expressed as mean ± SD. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>MCL1/FAK and MCL1/p130Cas interaction. Images reporting IP analysis in MDA-MB-231 treated cells with respect to the control cells show the interaction of MCL1 with FAK and p130Cas. IP was performed as described in the Materials and Methods section. In the left panel are Western blot images of the immunoprecipitated proteins (400 µg) with MCL1 antibody and detected with MCL1, FAK, and p130Cas. Mock lanes are obtained by loading beads alone. The right panel report Western blot analysis of MCL1, FAK, and p130Cas performed in input proteins (20 µg). Protein expression in input was normalized with GAPDH and expressed as fold changes. All the experiments were performed in triplicate. Data are expressed as mean ± SD. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>MCL1 Inhibition by A-1210477 reduces the aggression signatures in MDA-MB-231 cells via FA disruption.</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 1668
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 ">
Back to TopTop