Open Access

ADAM10 mediates shedding of carbonic anhydrase IX ectodomain non‑redundantly to ADAM17

  • Authors:
    • Miriam Zatovicova
    • Ivana Kajanova
    • Martina Takacova
    • Lenka Jelenska
    • Olga Sedlakova
    • Martina Labudova
    • Silvia Pastorekova
  • View Affiliations

  • Published online on: December 15, 2022     https://doi.org/10.3892/or.2022.8464
  • Article Number: 27
  • Copyright: © Zatovicova et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Carbonic anhydrase IX (CA IX) is a transmembrane enzyme participating in adaptive responses of tumors to hypoxia and acidosis. CA IX regulates pH, facilitates metabolic reprogramming, and supports migration, invasion and metastasis of cancer cells. Extracellular domain (ECD) of CA IX can be shed to medium and body fluids by a disintegrin and metalloproteinase (ADAM) 17. Here we show for the first time that CA IX ECD shedding can be also executed by ADAM10, a close relative of ADAM17, via an overlapping cleavage site in the stalk region of CA IX connecting its exofacial catalytic site with the transmembrane region. This finding is supported by biochemical evidence using recombinant human ADAM10 protein, colocalization of ADAM10 with CA IX, ectopic expression of a dominant‑negative mutant of ADAM10 and RNA interference‑mediated suppression of ADAM10. Induction of the CA IX ECD cleavage with ADAM17 and/or ADAM10 activators revealed their additive effect. Similarly, additive effect was observed with an ADAM17‑inhibiting antibody and an ADAM10‑preferential inhibitor GI254023X. These data indicated that ADAM10 is a CA IX sheddase acting on CA IX non‑redundantly to ADAM17.

Introduction

Ectodomain (ECD) shedding is an important regulatory mechanism, which controls spectrum and abundance of cell surface proteins mediating cellular responses to molecular and physiological signals (1). Concurrently, it generates soluble protein variants executing autocrine and/or paracrine signaling. Proper ECD shedding is essential for diverse biological processes, including cell fate decisions, proliferation, migration, metabolism, tissue and organ development, and its dysregulation contributes to various pathologies including cancer (2).

ECD shedding occurs via proteolytic cleavage of the extracellular portions of type I or II transmembrane proteins, or GPI-anchored proteins. It can be accomplished by various proteinases, among which a disintegrin and metalloproteinase (ADAM) family members have a prominent position (3,4). Out of 22 ADAM family members, ADAM17 and ADAM10 attract a lot of attention due to their functional implications in key signaling pathways. ADAM17, originally identified as TNFα-converting enzyme (TACE), is primarily related to EGFR signaling and represents a major sheddase for proinflammatory cytokines, whereas ADAM10 has been implicated mainly in Notch1 signaling and cleaves mostly cell adhesion molecules (5). However, ADAM17 and ADAM10 can also share several substrates (such as CD44, APP, DLL1, Notch, HB-EGF and TNFβ) suggesting that their complementary and/or compensatory activities are needed in certain cell or tissue contexts (6,7).

It was previously demonstrated by the authors that ADAM17 cleaves the ECD of carbonic anhydrase IX (CA IX) (811). CA IX is a cancer-associated, hypoxia-induced type I transmembrane protein participating in pH regulation, metabolic adaptation to hypoxia/acidosis and in cell migration-invasion-metastasis (1217). CA IX ECD consists of two structural components: a highly active enzyme domain (CA) catalyzing the reversible conversion of carbon dioxide to bicarbonate and proton, and an N-terminal proteoglycan-like region (PG). While the CA domain catalytic activity participates in the control of tumor pH (14,15,18,19), the PG region mediates non-catalytic proton extrusion across the plasma membrane and is implicated in cell adhesion (2024). A short stalk connecting the CA IX ECD to a single-pass transmembrane region contains the ADAM17 cleavage site. Deletion of 10 amino acids from the stalk completely prevents CA IX ECD shedding and reinforces the tumorigenic phenotype of CA IX-expressing cells (11).

However, neither silencing nor inhibition of ADAM17 are sufficient to fully block CA IX ECD release suggesting that an additional protease can participate in this process (8,11). As ADAM10 is structurally and functionally related to ADAM17, its possible involvement in CA IX ECD shedding was investigated. In the present study, the first experimental evidence was provided using biochemical and molecular/cell biology approaches that ADAM10 can cleave the CA IX ECD via an overlapping cleavage site with ADAM17 and it was showed that both proteinases contribute to CA IX ECD shedding in a non-redundant manner.

Materials and methods

Cell culture

Wild-type hamster ovary CHOwt cells and human cervical carcinoma C33a cells were transfected with the full-length (FL) human CA9 cDNA and its non-shed (NS) deletion mutant (del393-402 aa) using TurboFect™ transfection reagent (Thermo Fisher Scientific, Inc.) as previously described (11). The cells were routinely cultured in DMEM medium with 10% FCS (BioWhittaker; Lonza Group, Ltd.) in a humidified atmosphere of 5% CO2 at 37°C. Cell lines have been authenticated using STR profiling within the past 3 years. All experiments were performed with mycoplasma-free cells.

rhADAM10 cleavage activity assays

Catalytic activity of recombinant rhADAM10 encompassing the catalytic domain (100 ng/ml, R&D Systems, Inc.) was verified by hydrolytic processing of Fluorogenic Peptide Substrate IX Mca-K-P-L-G-L-Dpa-A-R-NH2 (10 µM; R&D Systems, Inc.) followed by monitoring of increasing fluorescence intensity at excitation and emission wavelengths of 320 and 405 nm (top read), respectively, in kinetic mode for 30, 40, 50 and 120 min at 37°C using a Synergy H4 plate reader (BioTek Instruments, Inc.). All reactions were performed in a fluorogenic buffer containing 25 mM Tris pH 9, 150 mM NaCl, 2.5 µM ZnCl2 and 0.005% Brij-35.

For the evaluation of ADAM10 cleavage of the cell membrane-bound CA IX, CHOwt cells transiently transfected with the FL human CA9 cDNA (CHOwt-FL-CA IX) and its deletion mutant NS CA IX (del393-402) (CHOwt-NS-CA IX), respectively, were treated with rhADAM10 (500 ng/ml) added to culture medium for 24 h. Thereafter, cell culture medium was harvested and CA IX ECD level was determined by ELISA.

Enzyme-linked immunosorbent assay (ELISA)

For the CA IX ECD quantification, 1×105 cells per well were seeded and cultured in a 24-well plate for 24 h. After medium exchange, the cells were incubated with either inhibitors or activators of shedding (as described below) and all supernatants were subsequently analyzed using in-house developed sandwich ELISA as previously described (8). The capture monoclonal antibody (Mab) V/10 specific for the catalytic domain of CA IX [10 µg/ml, generated in-house, (25)] was immobilized on the surface of microplate wells overnight at room temperature. After blocking and washing, diluted samples were added to the coated wells at room temperature for 2 h. The attached CA IX ECD was then allowed to react with the mixture of the PG domain-specific biotinylated Mabs [M75 and IV/18, generated and biotinylated in-house, (25)] diluted 1:7,500 (200 ng/ml) in the blocking buffer. The amount of bound detector antibodies was determined with peroxidase-conjugated streptavidin (Pierce; Thermo Fisher Scientific, Inc.) using the peroxidase substrate orthophenylene diamine (MilliporeSigma) after incubation for 1 h.

Indirect immunofluorescence

For detection of ADAM10, cells cultured on glass coverslips were incubated with anti-human ADAM10 ECD mouse Mab IgG2B Clone 163003 (R&D Systems, Inc.; cat. no. MAB1427, 1:100 in cultivation medium) for 1 h at 37°C, gently washed with PBS and fixed with methanol for 5 min at −20°C. Non-specific binding was blocked by incubation with PBS containing 1% BSA (MilliporeSigma) for 30 min at 37°C. Cells were then incubated with secondary donkey anti-mouse IgG Alexa Fluor®488-conjugated antibody (Invitrogen; Thermo Fisher Scientific, Inc.; cat. no. A-21202) diluted 1:1,000 in the blocking buffer for 1 h at 37°C. Finally, the coverslips were mounted onto slides in the Mounting Media with DAPI (cat. no. ab104139; Abcam), and analyzed by the confocal laser scanning microscope Zeiss LSM 510 Meta.

Proximity ligation assay (PLA)

PLA was performed in a humid chamber at 37°C according to the manufacturer's instructions (Olink) (26). Cells were seeded on glass coverslips and allowed to attach for 24 h. Then they were fixed with methanol, blocked with blocking buffer for 30 min and incubated with a mixture of antibodies against the ECDs of human CA IX (Abcam; rabbit, cat. no. ab15086) and ADAM10 (R&D Systems, Inc.; mouse MAB1427) each in 10 µg/ml concentration for 1 h. Following the incubation with plus and minus PLA probes for 1 h, a ligation mixture containing connector oligonucleotides and an amplification mixture containing fluorescently labeled DNA probe was added for 40 and 100 min, respectively. Samples were analyzed using a Zeiss LSM 510 Meta confocal microscope.

Internalization analyses

C33a-FL-CA IX (300,000 cells/dish) were plated on glass coverslips placed in Petri dishes of 3.5 cm diameter 24 h before the experiment. To analyze ADAM10 internalization induced by its inhibition, live cells were pre-incubated with anti-ADAM10 mouse MAB1427 from R&D Systems, Inc. (diluted 1:100 in culture medium) at 4°C for 30 min. Then the culture medium was removed, replenished by the medium containing preferential ADAM10 inhibitor GI254023X (GI; 10 µM; MilliporeSigma) in DMSO at 0.1 % (v/v) final concentration and the cells were incubated for additional 2, 4, 8 or 24 h at 37°C. Control samples were incubated without GI in medium containing DMSO at 0.1% (v/v) final concentration. After each incubation time point, parallel GI-treated and control samples were fixed for 5 min in ice-cold methanol at −20°C and stored for subsequent immunofluorescence. All samples were then subjected to indirect immunofluorescence using secondary donkey anti-mouse Alexa Fluor®488-conjugated antibody (Invitrogen; Thermo Fisher Scientific, Inc.; cat. no. A-21202) diluted 1:1,000 in blocking buffer containing 1% BSA for 1 h at 37°C. Finally, the coverslips were mounted onto slides in the Mounting Media with DAPI, and analyzed by a confocal laser scanning microscope Zeiss LSM 510 Meta.

To analyze the ADAM10 subcellular localization in response to internalization of CA IX, live cells were pre-incubated with anti-ADAM10 mouse MAB1427 from R&D Systems, Inc. (diluted 1:100 in culture medium) simultaneously with anti-CA IX humanized CA9hu-1 Mab capable of stimulating CA IX internalization [50 µg/ml; provided by Professor J. Pastorek, MABPRO, a.s. (27)] at 4°C for 30 min. Control samples were pre-incubated with anti-ADAM10 mouse MAB1427 only. Immediately after removal from 4°C, samples were transferred to 37°C to allow for internalization and then incubated for additional 2, 4, 8 and 24 h. After each incubation time point, parallel samples were fixed with methanol and stored as aforementioned. Fixed and blocked control samples that were not pre-treated with the CA9hu-1 antibody at the beginning of the experiment were post-incubated with that antibody, for 1 h at 37°C. All samples were then incubated with the mixture of secondary donkey anti-mouse Alexa Fluor®488-conjugated antibodies and goat anti-human Alexa Fluor®555-conjugated antibodies (Invitrogen; Thermo Fisher Scientific, Inc.; cat. no. A-21433) diluted 1:1,000 in the blocking buffer containing 1% BSA for 1 h at 37°C. Finally, the coverslips were mounted onto slides in Mounting Media with DAPI, and analyzed by the confocal laser scanning microscope Zeiss LSM 510 Meta.

Reverse transcription-quantitative PCR (RT-qPCR)

Total RNA was isolated using RNeasy Plus Mini Kit (Qiagen) and reverse transcription of 1 µg RNA for each sample was performed with the High-Capacity cDNA Reverse Transcription kit (Applied Biosystems; Thermo Fisher Scientific, Inc.) according to the manufacturer's instructions. qPCR was carried out using Maxima SYBR Green PCR Master mix (Thermo Fisher Scientific, Inc.) with the following gene-specific primers: ADAM10 sense, 5′-GACCACAGACTTCTCCGGAAT-3′ and antisense 5′-TGAAGGTGCTCCAACCCAAG-3′; ADAM17 sense, 5′-TGGATGAAGGAGAAGAGTGTGA-3′ and antisense, 5′-AAGATCCAAGCAAACAGTGTCAT-3′; and β-actin sense, 5′-TCCTCCCTGGAGAAGAGCTA-3′ and antisense, 5′-ACATCTGCTGGAAGGTGGAC-3′. qPCR was performed for 10 min at 95°C for initial denaturation followed by 40 cycles of 95°C for 15 sec and 60°C for 1 min. Sample CT values were normalized to β-actin. Relative expression was calculated using the 2-ΔΔCq method (28). All amplifications were performed in triplicates.

RNA interference

For siRNA-mediated silencing of de novo ADAM10 synthesis, C33a-FL-CA IX cells were seeded at 8×104 cells/cm2. One day later, cells were transfected with 1 µg of either MISSION® enzymatically-prepared small interfering RNA (esiRNA) human ADAM10 (Sigma-Aldrich; Merck KGaA; cat. no. EHU129311) or control esiRNA targeting RLUC (Sigma-Aldrich; Merck KGaA; cat. no. EHURLUC) using Attractene transfection reagent (Qiagen) according to the manufacturer's recommendations. Briefly, esiRNA was diluted in serum-free DMEM medium and Attractane Transfection Reagent in a total volume of 100 µl, and incubated for 15 min at room temperature. Following the medium exchange, the transfection reagent/esiRNA mixture was added drop-wise onto C33a-FL-CA IX cells. The day after transfection, cells were incubated with fresh cultivation medium either with or without 10 µM GI inhibitor for 24 h [as previously described (29)] and subsequently processed for ELISA analysis. Effect of ADAM10 silencing was evaluated by RT-qPCR in transfected versus control cells.

Expression of dominant-negative mutant of ADAM10 (ΔMP)

C33a-FL-CA IX cells were plated into 35-mm Petri dishes to reach ~70% monolayer density on the next day. Transient transfection was performed with 6 µg of pcDNA3-Delta (Pro-MP) ADAM10-HA plasmid encoding a ΔMP mutant lacking prodomain and metalloprotease domain, a gift from Axel Ullrich (Addgene, Inc; plasmid cat. no. 65107; http://n2t.net/addgene:65107; RRID: Addgene_65107) using TurboFect™ reagent according to the manufacturer's recommendation. Briefly, plasmid DNA (6 µg) was diluted in serum-free DMEM medium and TurboFect Transfection Reagent in a total volume of 200 µl, and incubated for 15 min at room temperature. The transfection reagent/plasmid DNA mixture was added drop-wise to each dish without medium exchange. Cells transfected with empty vector served as negative control. One day later, transfected cells were trypsinized and plated in six-well plates. Half of the transfected cells were incubated in the presence of GI inhibitor diluted in DMEM to working concentration of 10 µM. As a negative control, DMSO was added in a volume corresponding to the concentration of the inhibitor. After 24 h, proteins from transfected cells were extracted and cell culture medium was collected for ELISA.

Western blotting

Western blotting was performed as previously described (8). In brief, C33a-FL-CA IX and C33a-NS-CA IX cells were lysed in ice-cold RIPA buffer (150 mM NaCl, 1% Triton X-100, 0.05% NaDOC, 1 mM EDTA, 0.1% SDS, 50 mM Tris-HCl pH 7.4) containing inhibitors of proteases (Roche Applied Science). Cell lysates were collected and cleared by centrifugation at 10,000 × g for 15 min at 4°C. Total protein concentration was determined using the BCA protein assay reagent (Thermo Fisher Scientific, Inc.) and 50 µg/ml of total proteins mixed with Laemmli buffer were loaded per lane, separated in 10% SDS-PAGE and transferred onto PVDF membrane (Macherey-Nagel GmbH; cat. no. 741260). The membranes were treated with blocking buffer containing 5% (w/v) non-fat milk with 0.2% Nonidet P40 for 1 h and probed with the following primary antibodies: in-house generated anti-CA IX M75 Mab (hybridoma medium 1:3 in 5% non-fat dry milk with 0.2% Nonidet P40 in PBS, 1 h, room temperature); anti-ADAM17 (cat. no. 3976; 1:1,000 in 3% BSA in TBS-T buffer, overnight, 4°C), anti-ADAM10 (cat. no. 14194; 1:1,000 in 3 % BSA in TBS-T buffer, overnight, 4°C); anti-β-actin (cat. no. 3700) 1:5,000 in 3% BSA in TBS-T buffer (0,1% Tween-20), 1 h, room temperature; all from Cell Signaling Technology, Inc). After washing with 0.1% Tween-20 in PBS, the membranes were incubated for 1 h at room temperature, with the secondary antibodies: HRP-conjugated goat anti-mouse IgG (Sigma-Aldrich; Merck KGaA; cat. no. A0168) or HRP-conjugated goat anti-rabbit IgG (Bio-Rad Laboratories, Inc.; cat. no. 1721019) diluted 1:5,000 in blocking buffer for 1 h at room temperature. Protein bands were detected using enhanced chemiluminescence kit (GE Healthcare Bio-Sciences).

Activation and inhibition of CA IX ECD shedding

For the activation of shedding, treatment of cells with phorbol-12-myristate-13-acetate (PMA; MilliporeSigma) at 20 µM final concentrations, and/or ionomycin (IONO; MilliporeSigma) at final concentration of 1 µg/ml was performed for 3 h. For the inhibition of shedding, cells were treated with serum-free medium containing preferential ADAM10 inhibitor GI (1 µM), and/or ADAM17-inhbiting antibody D1(A12) (200 nM; Abcam; cat. no. ab215268) for 3 h at 37°C.

Bioinformatics

In silico analysis of the ADAM10 and ADAM17 promoters was performed using MatInspector (https://www.genomatix.de) (30,31). The promoter sequences with the overall length of 1,418 and 1,643 bp were extracted directly from the ElDorado genome database for the ADAM10 and ADAM17 genes, respectively. Transcription factors (TFs) were selected according to the core (0.75) and matrix (optimized) similarity. The accurate position of predicted binding elements was calculated according to the transcription start site. Phenotype heatmaps containing comparisons of relative transcriptional patterns of CA9, ADAM10, and ADAM17 in tumor tissue specimens from glioblastoma and colorectal carcinoma patients were generated through IST (in silico transcriptomics) online Medisapiens database (https://ist.medisapiens.com, accessed on March 10, 2021, currently available at http://tracxn.com/d/companies/medisapiens.com).

Statistical analysis

Data were analyzed using unpaired two-tailed Student's t-test or one-way ANOVA with Dunnett's multiple comparison test using GraphPad Prism 9 (GraphPad Software, Inc.). P<0.05 was considered to indicate a statistically significant difference.

Results

CA IX cleavage by recombinant human ADAM10

It was first aimed to gain biochemical evidence that CA IX is an ADAM10 substrate. Therefore, a recombinant human ADAM10 catalytic domain (rhADAM10) was used, the activity of which was proven by the cleavage of a commercial fluorogenic peptide (Fig. 1A). RhADAM10 was then shown to cleave a transiently expressed full-length (FL) CA IX from the surface of transfected wild-type CHOwt cells, that express a functional ADAM17 proteinase, but lack an endogenous ADAM10 proteinase (Fig. 1B). In line with this finding, incubation of CHOwt-FL-CA IX cells with ADAM17 activator PMA resulted in shedding of CA IX ECD, while ADAM10 activator IONO had no effect (Fig. 1C). However, CA IX ECD was shed to culture medium upon external addition of rhADAM10 (Fig. 1C).

Moreover, rhADAM10 was used to treat transiently transfected CHOwt-NS-CA IX cells expressing the NS CA IX variant generated by the deletion of 10 amino acids from the stalk region (11). As recently demonstrated, the NS CA IX variant exhibits cell surface expression similar to FL CA IX (11). Even though rhADAM10 was able to significantly increase the level of ECD shed from FL CA IX, it could not cleave the NS CA IX variant (Fig. 1B). These findings demonstrated that CA IX is a direct ADAM10 substrate and that the ADAM10 cleavage site is localized in the fragment deleted from the NS variant, which also contains the cleavage site for ADAM17 (11).

Colocalization and proximity of ADAM10 and CA IX

The role of endogenous ADAM10 in ECD shedding from the CA IX protein expressed on the surface of human cancer cells was next examined. For this purpose, stably transfected C33a cervical carcinoma cells constitutively expressing FL CA IX were employed. C33a-FL-CA IX cells express ADAM10 at the mRNA and protein levels and subcellular localization similar to control C33a-neo cells, as demonstrated by indirect immunofluorescence using the MAB1427 antibody specific for human ADAM10 (Fig. 2A and B). To further explore the spatial relationship between ADAM10 and CA IX, PLA was performed, which visualizes protein-protein interactions in situ. This in-cell co-immunoprecipitation represents an alternative to conventional co-immunoprecipitation of extracted proteins while preserving the natural context of the analyzed interactors (26). PLA was executed with the primary antibodies binding to the N-terminal regions of ADAM10 and CA IX. Results revealed a clear fluorescent signal suggesting that ADAM10 is in close proximity and can interact with CA IX (Fig. 2C). No PLA signal was observed in the control C33a-neo cells (Fig. 2D).

Co-internalization of ADAM10 and CA IX

It was next aimed to determine whether ADAM10 remains co-distributed with CA IX during internalization induced by the specific CA9hu-1 Mab directed to the CA IX catalytic domain (27). For that purpose, C33a-FL-CA IX cells were first incubated at 4°C with both anti-ADAM10 and anti-CA IX antibodies to enable their binding only to the subpopulation of ADAM10 and CA IX molecules present on the cell surface. Control samples were pre-treated only with ADAM10 antibody, which is unable to trigger internalization. The pre-bound antibodies could enter the cells together with their antigens only via active endocytosis (i.e. internalization), which was facilitated by transferring the cells to 37°C. This approach allowed us to observe the endocytosis and recycling of the complexes of the primary antibodies bound to their antigens throughout the experimental period and detect them with the corresponding secondary antibodies at different time points. In the absence of internalization-inducing CA9hu-1 antibody, both CA IX and ADAM10 displayed plasma membrane localization (Fig. 3A-D). By contrast, CA9hu-1 antibody-induced internalization of CA IX was clearly visible already after 2 h and was accompanied by internalization of ADAM10 as indicated by their overlapping fluorescence signals (Fig. 3E-G). Later on, both molecules recycled back to the plasma membrane (Fig. 3G and H).

Effect of molecular targeting of ADAM10 on CA IX ECD shedding

Based on the aforementioned observations, it was aimed to directly target ADAM10 expression in C33a-FL-CA IX cells by RNA interference and follow its impact on CA IX ECD shedding. The initial attempts to suppress ADAM10 were not fully satisfactory despite the clear downregulation of ADAM10 transcription (Fig. S1A), possibly due to the fact that the reduced de novo synthesis of ADAM10 protein remained unrecognized within the large pool of existing ADAM10 molecules. To solve this problem, an approach inspired by the study of Seifert et al (29) was chosen, where it was revealed that exposing cells to GI, a preferential ADAM10 inhibitor that binds to the protease active-site, leads to its internalization followed by lysosomal degradation of mature ADAM10 molecules. Therefore, it was assumed that inactivation linked to a reduction in the pool of existing mature ADAM10 molecules would enhance the effect of ADAM10 suppression. We first verified that GI causes the ADAM10 internalization also in C33a-FL-CA IX cells (Fig. 4). This was performed by pre-incubation of live cells with anti-ADAM10 antibody to allow for its binding only to the molecules exposed on the cell surface. Then the cells were incubated with or without GI at 37°C to enable endocytosis. Afterwards, the cells were fixed and stained with the secondary antibody. The cells without added GI inhibitor displayed ADAM10 on the plasma membrane (Fig. 4A-D). By contrast, exposure of cells to 10 µM GI at 37°C led to a reduced membrane localization of ADAM10 apparently due to its internalization (Fig. 4E-H).

Then, transient esiRNA-mediated ADAM10 silencing was performed, followed by 24 h incubation of the transfected cells with and without GI inhibitor at 10 µM concentration (Fig. 4I). Rather than using a single chemically-synthesized siRNA, it was decided to employ an esiRNA made up of a diverse pool of siRNA that all target the same mRNA sequence. Each individual siRNA has a lower concentration in the pool, leading to lower off-target effects while producing an efficient knockdown. Alternatively, transient expression of a ∆MP lacking the pro- and metalloprotease domains was accomplished (Figs. 4J and S1B) (32) again followed by the 24 h treatment with or without 10 µM GI. As recently demonstrated, the long-term effect (24 h) of 10 µM GI is limited to ADAM10 and not observed for ADAM17, which is not a primary target of this inhibitor (29). Culture media from cells subjected to esiRNA or ∆MP transfections incubated with and without GI were then analyzed for CA IX ECD by ELISA (Fig. 4I and J).

Both approaches led to similar effects on CA IX ECD shedding (Fig. 4I and J). Targeting of ADAM10 alone reduced CA IX ECD release to medium by 9% for esiRNA and 18% for ∆MP. Silencing of de novo synthesis was inferior compared with the ∆MP expression as ∆MP protein could potentially dimerize with existing ADAM10 molecules and perturb their function. Inhibition of mature ADAM10 by GI in non-transfected cells decreased CA IX ECD shedding by 15% for esiRNA and 28% for ∆MP. However, simultaneous targeting and inhibition of ADAM10 resulted in 37% (esiRNA) and 41% (∆MP) reduction of CA IX ECD release to the culture medium. These data are in line with the aforementioned biochemical and co-localization experiments and further support the view that ADAM10 is involved in the CA IX ECD shedding.

Additive effects of ADAM10 and ADAM17 activators and inhibitors on CA IX ECD shedding

Since C33a cells express both ADAM10 (Fig. 2A) and ADAM17 (11), they represent a suitable cell model to investigate possible relationship between these two ADAMs in CA IX ECD shedding. qPCR analysis of C33a cells revealed that the level of ADAM10 transcript is significantly higher (~2.5-times) than the level of ADAM17 in these cells irrespectively of whether they express FL CA IX or NS CA IX (Fig. 5A). Expression of both metalloproteinases in C33a cells was also revealed by Western blotting (Fig. 5B). Although the intensity of protein bands appeared to be similar, these two results cannot be directly compared due to the fact that the antibodies specific for ADAM10 and ADAM17, respectively, may differ in the binding properties. As revealed in the present study for ADAM10 (Figs. 2 and 3) and recently for ADAM17 (11) by indirect immunofluorescence of live cells, both metalloproteinases are localized on the cell surface in CA IX proximity where they can accomplish CA IX ECD cleavage.

C33a-FL-CA IX cells were treated either with IONO that is known to activate primarily ADAM10 or with PMA predominantly activates ADAM17 (33), or simultaneously with both activators to find out their individual and combined effect on activated CA IX ECD shedding. Medium was harvested from the cells treated with activators for 3 h and analyzed for CA IX ECD by ELISA. As revealed in Fig. 5C, each activator was able to significantly increase the level of CA IX ECD from C33a-FL-CA IX cells and moreover, they showed an additive effect when used together. No basal ECD shedding and no activation was observed in C33a-NS-CA IX cells (data not shown).

C33a-FL-CA IX cells were then treated with inhibitors (Fig. 5D), namely the ADAM17-inhibiting antibody D1(A12) that binds to a conformation-sensitive cross-domain epitope on ADAM17 molecule (34) and GI at 1 µM concentration that is fully specific for ADAM10 and not sufficient for inhibition of ADAM17 in this experimental setting (35) (Fig. S1D). D1(A12) Mab was used because it is specific only for ADAM17, whereas other known ADAM10 inhibitor, a hydroxamate-based TAPI-0, can inhibit not only ADAM17 but also MMP metalloproteinases and potentially also carbonic anhydrases (36). GI was used at 1 µM concentration, since at 10 µM concentration it could slightly reduce shedding of CA IX from the surface of live CHO cells that express only ADAM17 but not ADAM10 (Fig. S1C and D). In the internalization experiments, slight inhibition of ADAM17 alongside full inhibition of ADAM10 did not interfere with the ADAM10 endocytosis. However, in the experiments aimed at CA IX ECD shedding from C33a cells that express both sheddases, it was required to be certain that GI specifically inhibits only ADAM10. Importantly, a significant additive effect of inhibitors was observed on constitutive CA IX shedding from C33a-FL-CA IX (Fig. 5D), suggesting contribution of both ADAM17 and ADAM10.

Collectively, these data suggested that CA IX ECD shedding is mediated not only by ADAM17, as previously described (8,11), but also by ADAM10, as demonstrated in the present study.

Discussion

CA IX is a cancer-associated carbonic anhydrase isoform expressed in a broad range of tumor types, where it is primarily localized on the surface of cells exposed to chronic hypoxia and/or cells with oncogenic mutation(s) leading to activation of the hypoxia-inducible factor (HIF) pathway. The cell surface position allows CA IX to use the exofacial catalytic domain to regulate tumor pH via reversible conversion of CO2 generated by oncogenic metabolism to bicarbonate ions and protons. Bicarbonate ions are imported into the cytoplasm by bicarbonate transporters to maintain slightly alkaline intracellular pH, which is crucial for cell survival and proliferation, whereas protons remain in pericellular space and contribute to extracellular acidosis, which supports invasiveness and metastasis. The N-terminal proteoglycan-like region of the extracellular domain of CA IX mediates cell adhesion and participates in a non-catalytic export of protons in co-operation with lactate transporters [reviewed in (13,37)].

It was previously demonstrated by the authors that CA IX is a highly stable protein with a half-life close to 40 h (38). However, ~10% of the CA IX molecules undergo constitutive shedding of their ECD, which is sensitive to metalloproteinase inhibitor batimastat (8). Even though hypoxia increases the level of the shed ECD, the fraction of the soluble CA IX ECD is proportional to the hypoxia-induced expression of the cell-associated protein. Nevertheless, it was identified that CA IX ECD shedding can be increased by PMA, which activates ADAM17 through the PKC pathway, and both biochemical and molecular evidences were provided for the ADAM17-mediated CA IX shedding via the cleavage site localized in the stalk region proximal to the transmembrane region of the CA IX molecule (8,11).

In the present study, it was shown for the first time that CA IX ECD can be also cleaved by ADAM10. In addition, it was demonstrated that CA IX and ADAM10 are localized in close proximity. This is further supported by the finding that the antibody-triggered internalization of CA IX is associated with the internalization of ADAM10. Both molecules exhibit overlapping subcellular distribution not only during endocytosis, but also during recycling to the plasma membrane indicating that they remain associated throughout the intracellular path. Thus, the co-localization of ADAM10 with CA IX fulfils the spatial prerequisite for their interaction and CA IX ECD cleavage.

The data obtained with the NS CA IX variant suggested that ADAM10 can mediate CA IX ECD shedding via the cleavage site overlapping with that of ADAM17. Based on comparison of the sequence of amino acids deleted from the NS mutant with the predicted cleavage sites for dual ADAM10&17 substrates (39), it can be deduced that the cleavage site may be localized on the CA IX molecule between the positions 402*403 corresponding to PRAAE*PVQ sequence. Moreover, data obtained in the present study supported the view that ADAM10 and ADAM17 proteinases are involved in both constitutive shedding (as demonstrated by inhibitors) and activated shedding (as demonstrated by activators) of CA IX.

ADAM10 is regulated by different pathways and can cleave a spectrum of molecular targets only partially overlapping with targets of ADAM17 (5,33). Thus, ADAM10 appears to facilitate CA IX ECD shedding in additional signaling and physiological contexts, and thereby can affect responses mediated by soluble CA IX also in the absence of ADAM17. Based on previous data by the authors, soluble CA IX appears to interfere with the local functions of the cell-tethered CA IX protein that are required for tumorigenic and metastatic phenotype, including proper execution of the pH-regulatory and adhesion functions, presumably via competitive interactions with CA IX partner proteins in the plasma membrane or in the extracellular space (11).

When looking more closely at the regulation and effector functions of ADAM10 versus ADAM17, a large diversity of molecular events can be observed, that play a role in distinct expression patterns as well as activities of these two related proteinases, which execute CA IX ECD shedding (Fig. 6).

Data from the literature and from in silico analysis of ADAM10 and ADAM17 promoters using the MatInspector tool (Fig. S2) suggested involvement of the general TF SP1 through multiple binding sites in transcription of both ADAM10 and ADAM17 genes (40,41). Both genes were also shown to be transcriptionally activated by hypoxia (42,43). Typical hypoxia-response element sequences recognized by the HIF TF were found and functionally proven in ADAM17 promoter, while ADAM10 promoter appears to contain just a HIF-ancillary sequence and two ARNT (i.e. HIF-β) binding sequences. In addition, both ADAM10 and ADAM17 promoters include several binding sites for ETS1 factor, which can cooperate with other TFs, including CREB, SP1 and HIF. Notably, while ADAM10 seems to respond also to moderate hypoxia (0.5–5% of oxygen), ADAM17 is transcriptionally induced by severe hypoxia (below 0.1% O2) (42,43). Moreover, transcription of ADAM10 was shown to be induced by retinoic acid (RA) through its binding to RA receptor and retinoid receptors to RXR motifs in the ADAM10 promoter (40). ADAM10 (but not ADAM17) is also regulated by paired box 2 TF, particularly in WT1-mutated kidney cancers through multiple promoter binding sites and by XBP1 factor involved in response to DNA damage (4446). On the other hand, ADAM17 transcription is regulated by forkhead box M1 (47). Both ADAM10 and ADAM17 can be also regulated post-transcriptionally by distinct sets of miRNAs (48).

Another level of expression diversity is given by the post-translational regulation. Maturation of ADAM10 depends on tetraspanins (TSPANC8), whereas maturation of ADAM17 is driven by rhomboid chaperones iRhom1 and iRhom2 encoded by RHBDF1 and 2 genes (49,50). Moreover, ADAM10 can be inhibited by tissue inhibitor of metalloproteinase (TIMP) 1 and TIMP3, whereas ADAM17 can be inhibited by the TIMP3 and TIMP4 molecules, which are expressed in various tumor types at various levels and can locally impair the cleavage of ADAMs' substrates including CA IX (51). Finally, activation of shedding can be mediated either by release of Ca2+ ions through ADAM10 or by pathways involving PKC, EGFR/PI3K, ERK, TNFα or NF-κB signaling through ADAM17 (33).

All these circumstances create specific tumor tissue contexts, where the shedding of CA IX ECD may depend on multiple factors such as those driving the expression of CA IX, expression of ADAM10 and ADAM17, or on their inhibitors and activators (Fig. 6). This complex picture is evident from the phenotype heatmap generated by IST Medisapiens (https://ist.medisapiens.com) analysis of the RNA-seq metadata for glioblastoma and colorectal carcinoma, illustrating expression of selected genes in tumor tissues of individual patients (Fig. S3A and B). The analysis showed distinct transcription patterns of CA IX, ADAM17 and ADAM10, suggesting that in certain CA IX-expressing tumors lacking ADAM17, CA IX ECD shedding can be potentially executed by ADAM10. This renders CA IX shedding possible in a broader range of tumor tissues. Similarly, distinct expression patterns can be observed also in other tumor types.

Accumulating experimental and clinical studies demonstrate correlations of CA IX expression in cancer or stromal cells to aggressive tumor phenotype, poor patient prognosis and resistance to therapy. However, potential clinical value of CA IX ECD shedding remains unclear despite numerous efforts to exploit soluble CA IX as non-invasive tumor biomarker. While certain studies support its prognostic and/or predictive value, others do not show any significant relationship between the CA IX ECD levels and clinical parameters. This can be at least in part due to complex regulation of the CA IX ECD shedding. Uncovering the involvement of ADAM10 in the CA IX ECD cleavage sheds a new light on this intricate phenomenon and represents an important step towards its improved understanding in terms of both tumor biology and clinical applications.

Supplementary Material

Supporting Data

Acknowledgements

The authors would like to thank professor Jaromír Pastorek (MABPRO, a. s.) for generous gift of the CA9hu-1 Mab and Dr Tereza Golias (Biomedical Research Center SAS) for language editing.

Funding

The present study was supported by the Slovak Scientific Grant Agency (grant nos. VEGA 2/0074/20 and VEGA 2/0076/20), the Research & Developmental Support Agency (grant nos. APVV-15-0697 and APVV-19-0098) and the George Schwab and Leona Lauder Foundation.

Availability of data and materials

All data generated or analyzed during this study are included in this published article.

Authors' contributions

MZ, IK and MT designed and performed the experiments, analyzed/interpreted the data and contributed to the manuscript writing. LJ contributed to immunofluorescence and ELISA assays. OS performed the qPCR experiments. ML contributed to confocal microscopy. MT performed bioinformatic analyses. SP designed the study, supervised the experiments, interpreted the data, wrote and edited the manuscript. MZ, IK and MT confirm the authenticity of all raw data. All authors have read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

SP, MZ and MT are co-inventors of patents related to CA IX. The remaining authors declare that they have no competing interests.

Glossary

Abbreviations

Abbreviations:

ADAM

a disintegrin and metalloproteinase

CA IX

carbonic anhydrase IX

CHO

Chinese hamster ovary cells

ECD

ectodomain

FL

full-length

GI

GI254023X

HIF

hypoxia-inducible factor

Mab

monoclonal antibody

∆MP

dominant-negative mutant of ADAM10

NS

non-shed

PLA

proximity ligation assay

PMA

phorbol 12-myristate 13-acetate

TF

transcription factor

TIMP

tissue inhibitor of metalloproteinases

References

1 

Lichtenthaler SF, Lemberg MK and Fluhrer R: Proteolytic ectodomain shedding of membrane proteins in mammals-hardware, concepts, and recent developments. EMBO J. 37:e994562018. View Article : Google Scholar : PubMed/NCBI

2 

Hayashida K, Bartlett AH, Chen Y and Park PW: Molecular and cellular mechanisms of ectodomain shedding. Anat Rec (Hoboken). 293:925–937. 2010. View Article : Google Scholar : PubMed/NCBI

3 

Murphy G: The ADAMs: Signalling scissors in the tumour microenvironment. Nat Rev Cancer. 8:929–941. 2008. View Article : Google Scholar : PubMed/NCBI

4 

Mullooly M, McGowan PM, Crown J and Duffy MJ: The ADAMs family of proteases as targets for the treatment of cancer. Cancer Biol Ther. 17:870–880. 2016. View Article : Google Scholar : PubMed/NCBI

5 

Saftig P and Reiss K: The ‘A Disintegrin And Metalloproteases’ ADAM10 and ADAM17: Novel drug targets with therapeutic potential? Eur J Cell Biol. 90:527–535. 2011. View Article : Google Scholar : PubMed/NCBI

6 

Pruessmeyer J and Ludwig A: The good, the bad and the ugly substrates for ADAM10 and ADAM17 in brain pathology, inflammation and cancer. Semin Cell Dev Biol. 20:164–174. 2009. View Article : Google Scholar : PubMed/NCBI

7 

Vincent B and Checler F: α-Secretase in Alzheimer's disease and beyond: Mechanistic, regulation and function in the shedding of membrane proteins. Curr Alzheimer Res. 9:140–156. 2012. View Article : Google Scholar : PubMed/NCBI

8 

Zaťovičová M, Sedláková O, Švastová E, Ohradanova A, Ciampor F, Arribas J, Pastorek J and Pastorekova S: Ectodomain shedding of the hypoxia-induced carbonic anhydrase IX is a metalloprotease-dependent process regulated by TACE/ADAM17. Br J Cancer. 93:1267–1276. 2005. View Article : Google Scholar : PubMed/NCBI

9 

Zaťovičová M and Pastoreková S: Modulation of cell surface density of carbonic anhydrase IX by shedding of the ectodomain and endocytosis. Acta Virol. 57:257–264. 2013. View Article : Google Scholar : PubMed/NCBI

10 

Vidlickova I, Dequiedt F, Jelenska L, Sedlakova O, Pastorek M, Stuchlik S, Pastorek J, Zatovicova M and Pastorekova S: Apoptosis-induced ectodomain shedding of hypoxia-regulated carbonic anhydrase IX from tumor cells: A double-edged response to chemotherapy. BMC Cancer. 16:2392016. View Article : Google Scholar : PubMed/NCBI

11 

Kajanova I, Zatovicova M, Jelenska L, Sedlakova O, Barathova M, Csaderova L, Debreova M, Lukacikova L, Grossmannova K, Labudova M, et al: Impairment of carbonic anhydrase IX ectodomain cleavage reinforces tumorigenic and metastatic phenotype of cancer cells. Br J Cancer. 122:1590–1603. 2020. View Article : Google Scholar : PubMed/NCBI

12 

Russell S, Xu L, Kam Y, Abrahams D, Ordway B, Lopez AS, Bui MM, Johnson J, Epstein T, Ruiz E, et al: Proton export upregulates aerobic glycolysis. BMC Biol. 20:1632022. View Article : Google Scholar : PubMed/NCBI

13 

Pastorek J and Pastoreková S: Hypoxia-induced carbonic anhydrase IX as a target for cancer therapy: From biology to clinical use. Semin Cancer Biol. 31:52–64. 2015. View Article : Google Scholar : PubMed/NCBI

14 

Švastová E, Hulíková A, Rafajová M, Zat'ovicová M, Gibadulinová A, Casini A, Cecchi A, Scozzafava A, Supuran CT, Pastorek J and Pastoreková S: Hypoxia activates the capacity of tumor-associated carbonic anhydrase IX to acidify extracellular pH. FEBS Lett. 577:439–445. 2004. View Article : Google Scholar : PubMed/NCBI

15 

Švastová E, Witarski W, Csaderová L, Kosik I, Skvarkova L, Hulikova A, Zatovicova M, Barathova M, Kopacek J, Pastorek J and Pastorekova S: Carbonic anhydrase IX interacts with bicarbonate transporters in lamellipodia and increases cell migration via its catalytic domain. J Biol Chem. 287:3392–3402. 2012. View Article : Google Scholar : PubMed/NCBI

16 

Benej M, Svastova E, Banova R, Kopacek J, Gibadulinova A, Kery M, Arena S, Scaloni A, Vitale M, Zambrano N, et al: CA IX stabilizes intracellular pH to maintain metabolic reprogramming and proliferation in hypoxia. Front Oncol. 10:14622020. View Article : Google Scholar : PubMed/NCBI

17 

Gibadulinova A, Bullova P, Strnad H, Pohlodek K, Jurkovicova D, Takacova M, Pastorekova S and Svastova E: CAIX-mediated control of LIN28/let-7 axis contributes to metabolic adaptation of breast cancer cells to hypoxia. Int J Mol Sci. 21:42992020. View Article : Google Scholar : PubMed/NCBI

18 

Swietach P, Patiar S, Supuran CT, Harris AL and Vaughan-Jones RD: The role of carbonic anhydrase 9 in regulating extracellular and intracellular pH in three-dimensional tumor cell growths. J Biol Chem. 284:20299–20310. 2009. View Article : Google Scholar : PubMed/NCBI

19 

Chiche J, Ilc K, Laferriere J, Trottier E, Dayan F, Mazure NM, Brahimi-Horn MC and Pouysségur J: Hypoxia-inducible carbonic anhydrase IX and XII promote tumor cell growth by counteracting acidosis through the regulation of the intracellular pH. Cancer Res. 69:358–368. 2009. View Article : Google Scholar : PubMed/NCBI

20 

Závada J, Závadová Z, Pastorek J, Biesová Z, Jezek J and Velek J: Human tumour-associated cell adhesion protein MN/CA IX: Identification of M75 epitope and of the region mediating cell adhesion. Br J Cancer. 82:1808–1813. 2000. View Article : Google Scholar : PubMed/NCBI

21 

Jamali S, Klier M, Ames S, Barros LF, McKenna R, Deitmer JW and Becker HM: Hypoxia-induced carbonic anhydrase IX facilitates lactate flux in human breast cancer cells by non-catalytic function. Sci Rep. 5:136052015. View Article : Google Scholar : PubMed/NCBI

22 

Ames S, Pastorekova S and Becker HM: The proteoglycan-like domain of carbonic anhydrase IX mediates non-catalytic facilitation of lactate transport in cancer cells. Oncotarget. 9:27940–27957. 2018. View Article : Google Scholar : PubMed/NCBI

23 

Csaderová L, Debreová M, Radvák P, Stano M, Vrestiakova M, Kopacek J, Pastorekova S and Svastova E: The effect of carbonic anhydrase IX on focal contacts during cell spreading and migration. Front Physiol. 4:1–12. 2013. View Article : Google Scholar : PubMed/NCBI

24 

Radvák P, Repic M, Švastová E, Takacova M, Csaderova L, Strnad H, Pastorek J, Pastorekova S and Kopacek J: Suppression of carbonic anhydrase IX leads to aberrant focal adhesion and decreased invasion of tumor cells. Oncol Rep. 29:1147–1153. 2013. View Article : Google Scholar : PubMed/NCBI

25 

Zaťovičová M, Tarábková K, Švastová E, Gibadulinová A, Mucha V, Jakubícková L, Biesová Z, Rafajová M, Gut MO, Parkkila S, et al: Monoclonal antibodies generated in carbonic anhydrase IX-deficient mice recognize different domains of tumour-associated hypoxia-induced carbonic anhydrase IX. J Immunol Methods. 282:117–134. 2003. View Article : Google Scholar : PubMed/NCBI

26 

Söderberg O, Gullberg M, Jarvius M, Ridderstråle K, Leuchowius KH, Jarvius J, Wester K, Hydbring P, Bahram F, Larsson LG and Landegren U: Direct observation of individual endogenous protein complexes in situ by proximity ligation. Nat Methods. 3:995–1000. 2006. View Article : Google Scholar : PubMed/NCBI

27 

Zatovicova M, Kajanova I, Barathova M, Takacova M, Labudova M, Csaderova L, Jelenska L, Svastova E, Pastorekova S, Harris AL and Pastorek J: Novel humanized monoclonal antibodies for targeting hypoxic human tumors via two distinct extracellular domains of carbonic anhydrase IX. Cancer Metab. 10:32022. View Article : Google Scholar : PubMed/NCBI

28 

Livak KJ and Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(−Delta Delta C(T)) method. Methods. 25:402–408. 2001. View Article : Google Scholar : PubMed/NCBI

29 

Seifert A, Düsterhöft S, Wozniak J, Koo CZ, Tomlinson MG, Nuti E, Rossello A, Cuffaro D, Yildiz D and Ludwig A: The metalloproteinase ADAM10 requires its activity to sustain surface expression. Cell Mol Life Sci. 78:715–732. 2021. View Article : Google Scholar : PubMed/NCBI

30 

Cartharius K, Frech K, Grote K, Klocke B, Haltmeier M, Klingenhoff A, Frisch M, Bayerlein M and Werner T: MatInspector and beyond: Promoter analysis based on transcription factor binding sites. Bioinformatics. 21:2933–2942. 2005. View Article : Google Scholar : PubMed/NCBI

31 

Quandt K, Frech K, Karas H, Wingender E and Werner T: Matlnd and matlnspector: New fast and versatile tools for detection of consensus matches in nucleotide sequence data. Nucleic Acids Res. 23:4878–4884. 1995. View Article : Google Scholar : PubMed/NCBI

32 

Gschwind A, Hart S, Fischer OM and Ullrich A: TACE cleavage of proamphiregulin regulates GPCR-induced proliferation and motility of cancer cells. EMBO J. 22:2411–2421. 2003. View Article : Google Scholar : PubMed/NCBI

33 

Le Gall S, Bobé P, Reiss K, Horiuchi K, Niu XD, Lundell D, Gibb DR, Conrad D, Saftig P and Blobel CP: ADAMs 10 and 17 represent differentially regulated components of a general shedding machinery for membrane proteins such as transforming growth factor α, L-selectin, and tumor necrosis factor alpha. Mol Biol Cell. 20:1785–1794. 2009. View Article : Google Scholar : PubMed/NCBI

34 

Tape CJ, Willems SH, Dombernowsky SL, Stanley PL, Fogarasi M, Ouwehand W, McCafferty J and Murphy G: Cross-domain inhibition of TACE ectodomain. Proc Natl Acad Sci U S A. 108:5578–5583. 2011. View Article : Google Scholar : PubMed/NCBI

35 

Ludwig A, Hundhausen C, Lambert M, Broadway N, Andrews RC, Bickett DM, Leesnitzer MA and Becherer JD: Metalloproteinase inhibitors for the disintegrin-like metalloproteinases ADAM10 and ADAM17 that differentially block constitutive and phorbol ester-inducible shedding of cell surface molecules. Comb Chem High Throughput Screen. 8:161–171. 2005. View Article : Google Scholar : PubMed/NCBI

36 

Supuran CT: How many carbonic anhydrase inhibition mechanisms exist? J Enzyme Inhib Med Chem. 31:345–360. 2016. View Article : Google Scholar : PubMed/NCBI

37 

Becker HM: Carbonic anhydrase IX and acid transport in cancer. Br J Cancer. 122:157–167. 2020. View Article : Google Scholar : PubMed/NCBI

38 

Rafajová M, Zatovicová M, Kettmann R, Pastorek J and Pastoreková S: Induction by hypoxia combined with low glucose or low bicarbonate and high posttranslational stability upon reoxygenation contribute to carbonic anhydrase IX expression in cancer cells. Int J Oncol. 24:995–1004. 2004.PubMed/NCBI

39 

Tucher J, Linke D, Koudelka T, Cassidy L, Tredup C, Wichert R, Pietrzik C, Becker-Pauly C and Tholey A: LC-MS based cleavage site profiling of the proteases ADAM10 and ADAM17 using proteome-derived peptide libraries. J Proteome Res. 13:2205–2214. 2014. View Article : Google Scholar : PubMed/NCBI

40 

Prinzen C, Müller U, Enders K, Fahrenholz F and Postina R: Genomic structure and functional characterization of the human ADAM10 promoter. FASEB J. 11:1522–1524. 2005. View Article : Google Scholar : PubMed/NCBI

41 

Szalad A, Katakowski M, Zheng X, Jiang F and Chopp M: Transcription factor Sp1 induces ADAM17 and contributes to tumor cell invasiveness under hypoxia. J Exp Clin Cancer Res. 28:1292009. View Article : Google Scholar : PubMed/NCBI

42 

Barsoum IB, Hamilton TK, Li X, Cotechini T, Miles EA, Siemens DR and Graham CH: Hypoxia induces escape from innate immunity in cancer cells via increased expression of ADAM10: Role of nitric oxide. Cancer Res. 71:7433–7441. 2011. View Article : Google Scholar : PubMed/NCBI

43 

Rzymski T, Petry A, Kračun D, Rieß F, Pike L, Harris AL and Görlach A: The unfolded protein response controls induction and activation of ADAM17/TACE by severe hypoxia and ER stress. Oncogene. 31:3621–3634. 2012. View Article : Google Scholar : PubMed/NCBI

44 

Lee SB, Doberstein K, Baumgarten P, Wieland A, Ungerer C, Bürger C, Hardt K, Boehncke WH, Pfeilschifter J, Mihic-Probst D, et al: PAX2 regulates ADAM10 expression and mediates anchorage-independent cell growth of melanoma cells. PLoS One. 6:e223122011. View Article : Google Scholar : PubMed/NCBI

45 

Doberstein K, Pfeilschifter J and Gutwein P: The transcription factor PAX2 regulates ADAM10 expression in renal cell carcinoma. Carcinogenesis. 32:1713–1723. 2011. View Article : Google Scholar : PubMed/NCBI

46 

Reinhardt S, Schuck F, Grösgen S, Riemenschneider M, Hartmann T, Postina R, Grimm M and Endres K: Unfolded protein response signaling by transcription factor XBP-1 regulates ADAM10 and is affected in Alzheimer's disease. FASEB J. 28:978–997. 2014. View Article : Google Scholar : PubMed/NCBI

47 

Kim IM, Ramakrishna S, Gusarova GA, Yoder HM, Costa RH and Kalinichenko VV: The Forkhead Box m1 transcription factor is essential for embryonic development of pulmonary vasculature. J Biol Chem. 280:22278–22286. 2005. View Article : Google Scholar : PubMed/NCBI

48 

Vincent B: Regulation of the α-secretase ADAM10 at transcriptional, translational and post-translational levels. Brain Res Bull. 126:154–169. 2016. View Article : Google Scholar : PubMed/NCBI

49 

Matthews AL, Noy PJ, Reyat JS and Tomlinson MG: Regulation of A disintegrin and metalloproteinase (ADAM) family sheddases ADAM10 and ADAM17: The emerging role of tetraspanins and rhomboids. Platelets. 28:333–341. 2017. View Article : Google Scholar : PubMed/NCBI

50 

Düsterhöft S, Babendreyer A, Giese AA, Flasshove C and Ludwig A: Status update on iRhom and ADAM17: It's still complicated. Biochim Biophys Acta Mol Cell Res. 1866:1567–1583. 2019. View Article : Google Scholar : PubMed/NCBI

51 

Jackson HW, Defamie V, Waterhouse P and Khokha R: TIMPs: Versatile extracellular regulators in cancer. Nat Rev Cancer. 17:38–53. 2017. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

February-2023
Volume 49 Issue 2

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Zatovicova M, Kajanova I, Takacova M, Jelenska L, Sedlakova O, Labudova M and Pastorekova S: ADAM10 mediates shedding of carbonic anhydrase IX ectodomain non‑redundantly to ADAM17. Oncol Rep 49: 27, 2023
APA
Zatovicova, M., Kajanova, I., Takacova, M., Jelenska, L., Sedlakova, O., Labudova, M., & Pastorekova, S. (2023). ADAM10 mediates shedding of carbonic anhydrase IX ectodomain non‑redundantly to ADAM17. Oncology Reports, 49, 27. https://doi.org/10.3892/or.2022.8464
MLA
Zatovicova, M., Kajanova, I., Takacova, M., Jelenska, L., Sedlakova, O., Labudova, M., Pastorekova, S."ADAM10 mediates shedding of carbonic anhydrase IX ectodomain non‑redundantly to ADAM17". Oncology Reports 49.2 (2023): 27.
Chicago
Zatovicova, M., Kajanova, I., Takacova, M., Jelenska, L., Sedlakova, O., Labudova, M., Pastorekova, S."ADAM10 mediates shedding of carbonic anhydrase IX ectodomain non‑redundantly to ADAM17". Oncology Reports 49, no. 2 (2023): 27. https://doi.org/10.3892/or.2022.8464