Next Article in Journal
Targeting Lactylation: From Metabolic Reprogramming to Precision Therapeutics in Liver Diseases
Previous Article in Journal
The Role of Senolytics in Osteoporosis
Previous Article in Special Issue
Targeting APC/C Ubiquitin E3-Ligase Activation with Pyrimidinethylcarbamate Apcin Analogues for the Treatment of Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cell Settling, Migration, and Stochastic Cancer Gene Expression Suggest Potassium Membrane Flux May Initiate pH Reversal

Brain Health Research Institute, Kent State University, Kent, OH 44242, USA
Biomolecules 2025, 15(8), 1177; https://doi.org/10.3390/biom15081177 (registering DOI)
Submission received: 30 June 2025 / Revised: 3 August 2025 / Accepted: 8 August 2025 / Published: 16 August 2025
(This article belongs to the Collection Feature Papers in Chemical Biology)

Abstract

Attraction of glioblastoma cells to potassium was suspected when glioblastoma cells clustered around dying cells and migrated towards serum (high [K+]) and increased potassium. Potassium channel proteins (KCN family, 90 members) mediating alterations in the transmembrane flux may provide K+ that releases H+ bound to inner membranes in cancer cells for cytosolic proton transfer, possibly conformational in water (Grotthuss), to extrusion sites. Cell settling and migration assay results led to collecting 70 studies, unbiased by the authors for inclusion of KCN genes, that detected KCN differentially expressed genes (DEGs). Of 53 KCN DEGs found among 29 malignancies, 62.3% encoded H+-sensitive proteins. KCN DEGs encoding H+-sensitive proteins were more prevalent in 50 studies involving one or more categories (seven oncogenes and histone/DNA modifiers) versus those with none; p = 0.0325. Pertinent genes for lactate outflow, etc., had relatively normal levels of expression. Brain tumors in REMBRANDT (database) showed altered expression of KCN genes encoding H+-sensitive proteins in glioblastomas versus less invasive oligodendrogliomas of patients on anti-seizure medications, with less KCNJ16/Kir5.1; p = 5.32 × 10−8 in glioblastomas. Altered H+-sensitive potassium flux via the KCN family, downstream of oncogenes and histone/DNA modifiers, putatively incites proton transfers for H+ release during pH reversal (pHi > pHe) in cancer.

1. Introduction

1.1. General

Potassium flux maintains cell membrane potential, assists volume changes, and permits neurons to fire action potentials, as the potassium they release is buffered by glial cells. Potassium is proposed to also exchange with fixed H+ on inner membranes to incite proton transfer and extrusion for tumor cell survival and invasion. Reversal of the pH gradient, i.e., intracellular pH (pHi) is greater than extracellular pH (pHe), may occur. K+/H+ cation exchange may explain the attraction of glioblastoma cells (highest-grade glial tumor) to axons in CNS white-matter tracts and peripheral nerves [1,2]. Cell morphology and viability in an experimental microaqueous environment suggested glioblastoma cell attraction to K+. Viable cells tightly clustered around single dying cells, putatively releasing cytoplasmic K+. Also, glioblastoma cells migrated towards fetal bovine serum with levels of 5–15 mM K+ [3]. The in vitro clustering behavior reflected how glioblastoma cells typically rim necrotic foci (pseudopalisading), often being band-like in vivo [4,5,6]. Pseudopalisading cells have a migratory phenotype [6,7]. Related pH features of benign astrocytic (glial) cells include (i) an increase in pHi with neuronal stimulation, possibly triggered by membrane K+ redistribution during depolarization [8,9]. Intracellular alkalinization of glial cells has been attributed to their uptake of potassium released by repetitive neuronal firing in vivo [10]. Also, (ii) glycolysis (inhibited by lactic acid) is stimulated by changes in intracellular [H+] rather than [HCO3-] in astrocytes [11]. Cytoplasmic alkalinization occurring via efflux of protons, resulting in extracellular acidosis, has been attributed to ion transporters, pumps, exchangers, etc. [12,13,14,15,16,17,18,19]. Redistribution of K+ has been suggested as contributing to tumor cell alkalinization [12]. Also, pH-sensitive K+ channels may be “sensors”, as described in pH reversal [13,14,16,17]. This study proposes that potassium flux plays an active role.
The KCN (potassium channel) family became the focus of a stochastic search for evidence of anomalous K+ flux based on its large size, complex family interactions, extensive non-family regulation, functional reliance on biological milieus, and a wide chromosomal distribution that increases their odds of genetic damage. Retrieval of data unbiased for KCN gene expression in tumors on differentially expressed genes (DEGs) of KCN family members and associated non-KCN DEGs in 70 studies of many cancer types provides a stochastic view of anomalous potassium flux. Also, a brain tumor database, REMBRANDT, and genome analysis of six glioblastoma cell lines were queried for KCN genomic data [20,21,22]. The detected H+ sensitivity of a large proportion of KCN family members and high likelihood of K+ as a replacement cation for H+ suggest that potassium ion flux across plasma membranes or accumulation near them plays a role in pH reversal in malignancy. Deregulated, inward channel flux or loss of diffusely distributed outward K+ flow (necessitating focally concentrated outward flow), leading to accumulations of K+ ions near H+ bound to anions of inner cell membranes, potentially leads to K+/H+ exchange that releases H+ as an initial step in pH reversal.

1.2. Cytosolic Proton Transfer Related to K+ and Hydration

Potassium replacement of inner-membrane-bound H+ should augment hydronium ion (H3O+) accumulation seen at water–lipid interfaces [23]. This putatively sets the stage for pH reversal. The mobility of unbound H+ (H3O+ at 36.2 × 10−4 cm2 V−1s−1) is high compared to K+ and Na+ [24]. In immortalized embryonic kidney cells, H+ diffusion through the cytosol occurred 60-fold faster than when limited to the inner plasma membrane, i.e., 0.6 μms−1 versus 0.01 μms−1 [25]. In cardiac myocytes, the cytoplasmic H+ distribution speed, 0.7 μms−1, was comparable based on diffusion using HCl in KCl solutions with membrane acid transport inhibited [26]. Also, dissociation of HCl occurred in a microaqueous environment, with the acid’s protons transferred to a neighboring water molecule and then another in clusters of HCl(H2O)n, where n = 1 to 5 and 7. With addition of a fifth and seventh water molecule, the covalent H-Cl bond transitioned to the contact ion pair, H3O+Cl, plus free hydrogens. HCl(H2O)5 and HCl(H2O)7 clusters were spontaneously dissociated structures with contact ionic bonds [27]. Thus, H3O+ may be distributed to the cytosol after putative K+/H+ exchange on inner membranes and after dissociation of acids in cytosols, possibly microaqueous.
Cytosolic proton transfer has been attributed to taurine, carnitine, carnosine, or homocarnosine behaving as mobile buffers [26]. However, proton transfers from H3O+ may also occur by a rapid mechanism consistent with behavior described by C.J.T. Grotthuss in 1806 [28]. Proton transfers occur via H-bonded complexes, H9O4+ and H5O2+, or by wires of protonated water molecules linked by interconversions, consistent with behavior of salts and water that Grotthuss observed between electrodes. With these interconversions, a proton that enters a water matrix is not the same proton that exits elsewhere [24,28,29,30,31,32,33,34]. Conformational changes for proton transfer in water occur as a rapid falling-domino effect for wide-reaching delivery of cytoplasmic protons to membrane extruders. Bending of hydrogen bonds to accommodate acceptance of a new proton quickly ripples through a properly aligned water matrix to release another proton distally while snapping back to normal to be ready for the next proton that enters. Hydrogen-bonded water chains mediate proton transfer much faster than diffusion. Rapid transfer of H+ in water is also known as a “proton jump” [35]. H+ can be transferred to cell exit sites, such as exchangers swapping out H+, monocarboxylate transporters venting lactate that requires accompanying protons, exiting through gap junctions, etc. Although the Grotthuss mechanism may not mediate extracellular tumor acidity, it was not ruled out as occurring intracellularly [36]. The Grotthuss mechanism describes an “everywhere, all at once” strategy for delivery of protons to cell exit sites. Therefore, inner-membrane proton exchange with K+ is probably highly regulated in normal cells, consistent with K+ flux via the large, complex KCN family. Loss of regulation, reflected by differential expressions of KCN genes, allows assessment of potential ionic dynamics related to pH reversal in malignant cells. Genetic damage of KCN family members or altered regulation by oncogenes, modified histones/DNA, etc., putatively plays a role in pH reversal to aid survival of glycolytic cells producing lactate that is not removed via glycogen formation.
Also, loose aggregation of water molecules with potassium is important. Lack of water due to poor circulation, cell crowding, and pressure from fibrotic, stiff surroundings, especially during invasion, produces microaqueous conditions within cancer cells. However, regional hydration may occur by release of water molecules from K+-7H2O to join protonated water structures close to cell membranes, but this is unknown.
Volume changes related to ions occur in cells during division and migration. Protrusion of pseudopodia (increased volume in cell projections) and shrinkage of their retracted rear cell bodies occur with oscillations of cell membrane potential and osmotic changes via KCl uptake and release through potassium and chloride channels, with aquaporin participation and possibly K+ refilling/recycling [37,38,39,40,41,42]. Thus, cell migration produces ion and hydration dynamics that may influence proton transfer at the leading edge of migrating cells. Pseudopodia are glycolytic, based on proteomic analysis [43]. Also, enhanced proton release leading to extracellular acidosis aids tumor invasion via activation of metalloproteases for matrix degradation [44,45,46,47,48,49,50]. Tumor cells’ escape from poorly perfused, microaqueous tissue may be aided by K+/H+ exchange initiating pH reversal.

1.3. Need for Proton Extrusion in Cancer

Metabolism of cancer cells produces excess protons [15,51]. In addition to the acids produced by metabolic pathways, free radicals generated are also acidic [52,53,54]. Extrusion of protons avoids cytoplasmic acidosis, so this not only protects glycolysis but also prevents cytochrome c-mediated activation of caspase-3 to stop apoptosis [55], protects gap junctions [56,57,58,59], and activates ATP citrate lyase (ACLY), optimal pH of 8.5–8.7 [60,61]. ACLY’s removal of citrate (inhibitor of glycolysis) is protective in glioblastomas [62]. Also, by generating acetyl-CoA from citrate, nuclear ACLY promotes histone lysine acetylation, so that its positive charge is removed to allow loosening of chromatin that drives expression of genes for cell adhesion, migration, and integrin signaling for cancer cell invasion [63,64]. Thus, proton removal from the cytoplasm favors tumor cell fitness as well as enabling migration to escape adverse conditions [64,65].

1.4. Complexity of K+ Flux

The direction and strength of K+-channel flow occurs largely via the KCN family, encoded by 90 genes [66,67]. Also, numerous non-KCN regulators are present in biological milieus, such as polyamines, voltage, H+, Ca2+, Mg2+, Zn2+, Na+, ATP, miR-133b, reactive oxygen species (ROS), reactive nitrogen and sulfur species, nitric oxide, hydrogen sulfide, hypoxia, hyperglycemic glucose, nucleotides, protein kinases A, C, and G, guanosine-triphosphate-binding proteins (G-proteins), regulator of G protein signaling (RGS) proteins, phosphatidyl inositol 4,5 bisphosphate, estrogen, testosterone, cholesterol, sulfonylurea, β-hydroxybutyrate, hypoxia inducible factor-1α, 14-3-3 protein, β-coat protein, dipeptidyl aminopeptidase-like proteins, GABA, NAD(+), NADP(+), NADH, NADPH, glutathione, glutathione disulfide, mechanical stimulation, cell cycle progression, products of LINGO1,2, and BKγ1-4, encoded by LRRC26,52,55,38 [68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90]. Also, 19% of proteins targeted by small-molecule drugs used to treat numerous diseases are ion channels [91,92,93]. ROS are generated via mitochondrial respiratory chain electron transport redox reactions and the NOX family of NADPH oxidases [71,84,94,95,96,97,98]. Colon carcinoma cells with acid-resistant growth showed upregulated expression of NOX1/NADPH oxidase 1 [99]. Interestingly, isolated respiring rat brain mitochondria increased ROS production when their medium was alkalinized [100].
The widespread chromosomal distribution of KCN genes and non-family regulators suggests failsafe redundancies exist for genomic damage to maintain levels of K+ for survival, without being optimal for differentiation, in dividing tumor cells. Given that monocarboxylate transporters, MCT1 and MCT4, transport negatively charged lactate molecules coupled one-for-one with protons across cell membranes [101] and protons can also be extruded via Na+/H+ exchanger family members, ATP driven pumps, etc. [12,13,14,15,16,17,18,19,102], there are pertinent genes among the non-KCN DEGs that are also of stochastic interest.
Mediators of potassium flux influencing the relationship of K+ and H+ at cell membranes should be clarified. K+ accumulation due to a lack of normal diffuse K+ efflux or buildup of K+ around large outwardly conducting potassium channels, could putatively exchange with bound protons on inner membranes or it could happen with deregulated inwardly channeled K+ flux (possibly with K+ being recycled for refilling as needed if efflux becomes excessive). These events with K+ could occur without affecting the functions of key structural proteins that would be seen if Na+ was used [103]. Although both potassium and sodium are available to replace membrane-bound H+, there should be a preference for potassium due to its weaker binding of water, so that less energy is required to displace water from potassium [104,105]. K+ ions permeate their membrane channels in a desolvated (‘naked’) state rapidly, so at this point they release their associated water molecules. There is exclusion of smaller Na+ ions, which have stronger water binding [106,107,108]. Also, compared to Na+, K+ is better at maintaining solubility of cell components containing carboxylates and phosphatesto help explain its higher concentration in the cytoplasm [105,109]. Numerous KCN subfamilies may be involved with anomalous potassium flux. Deregulated diffuse, passive K+ entry via the KCNJ/Kir (inwardly rectifying, also bidirectional, flux) or disrupted K+ outflow via numerous Kv (voltage dependent) KCN subfamily and KCNK/K2P subfamily (tandem two-pore “leak”) members are important. Also, Ca2+ activated potassium channel family members that mediate outward flow, KCNMA1/BK, and beta subunits that regulate it, KCNN/SKs, and KCNN4/IK, with BK and IK providing large and medium flux, respectively, likely play important roles. Higher flow rates may provide compensatory localized K+ release and are likely affected by release of calcium from cancer cells. However, all types of K+ transits may be involved, and a role for Na+ in releasing H+ cannot be ruled out given the high levels of Na+ in some tumors [110,111,112]. A few K+ channels are influenced by Na+.
Loss of normal cellular responses to pH via loss or gain of KCN channels that function with H+ sensitivity may alter cancer cells. Glioblastoma cell behavior suggests an opportunistic response to sources of K+, thus achieving levels of K+ ions at innermembrane surfaces that promote K+/H+ exchanges to initiate pH reversal for cell survival and invasion. These sources include (i) neurons and axons firing action potentials where K+ is released to maintain electroneutrality, (ii) dying cells, including red cells (K+, 80–120 mmol/L) and necrotic tissue releasing cytoplasmic K+ [113,114,115,116], and (iii) damaged blood vessels leaking plasma (potassium, 3.8–5.0 mmol/L [117]). Deregulation via loss of H+-sensitive KCNK1 (among a small group of genes identified) was advantageous for breast cancer cell survival of induced necrosis [118].

1.5. Biological Milieu Constituents to Consider in Studying K+ Flux

Relevant details of milieu constituents in studying KCN channels are listed. Serum includes calcium (2.30–2.74 mmol/L), magnesium (0.74–1.23 mmol/L), phospholipids (1.50–3.80 g/L), sodium (136–142 mmol/L, adult plasma value) [117], and polyamines (0.1–1.0 μM, fetal bovine serum [3]). Magnesium and polyamines (positively charged) are important gating molecules of inwardly rectifying KCNJ/Kir-family channel members [119]. Removing fetal bovine serum from fibroblasts resulted in decreased intracellular potassium [120]. Also, fetal bovine serum lost the ability to depolarize the resting membrane potential of airway smooth muscle cells after dialysis against distilled water but its freeze-dried dialysate retained this ability after reconstitution [121]. Elevated levels of polyamines occur in cerebrospinal fluid of patients with brain tumors [122] and in tumor-associated myeloid cells infiltrating glioblastomas [123]. Also, polyamines are formed when L-arginine, released from apoptotic bodies, is metabolized by Arg1 [124]. Other modifiers within necrotic tissue include hemoglobin breakdown products [125,126]. Tumor-associated macrophages are influenced to promote malignancy (increased expression of ARG1, VEGF, TGFB, etc.) by lactate, presumably shuttled from cells of bladder and gastric cancers, and others [127,128,129].
As mentioned earlier, many non-CNS tumors exhibit perineural invasion [130,131,132,133]. There may be reciprocal gain of potassium by tumor cells from nerves and axons as the tumor cells provide lactic acid back to them as an energy substrate for a mutually beneficial interchange. Although reports of specific sources of K+ affecting behavior of cells in vivo have not been found, opportunistic sources of potassium may be important for dedifferentiated cancer cells.

1.6. Findings Highlighted in This Study

Fetal bovine serum (with and without hepatocyte growth factor (HGF)) promoted glioblastoma cell migratory behavior, including protrusion of pseudopodia through 3 μm filter pores, migration of whole cells through 8 μm filter pores, and invasion through Matrigel and rat brain tissue slices, often under glycolytic conditions [43,62,134,135,136]. Glioblastoma cell clustering around dying cells and their migration to serum, with compelling background information on astrocytes, suggested involvement of anomalous potassium flux in tumor cell behavior. Cell migration of multiple glioblastoma cell lines to serum among the eight tested in this study, and the response in a separate assay to a positive gradient of potassium, suggested abnormal membrane flux as supporting tumor cells. The large number of K+-channel candidates for loss or gains in expression under potential influence of numerous modulators led to stochastic analysis of published studies for unbiased data comprised of incidentally detected KCN DEGs and accompanying non-KCN DEGs in genomic landscapes of cancer scenarios and databases. This permitted a search for pertinent functional similarities among KCN DEGs. Meta-analysis of 70 malignant (or related models’) expression studies found that the studies fell into eight categories of seven oncogenes and histone/DNA modifiers. These categories were determined by the authors of the expression (E) studies, E1–E70, if related DEGs were also detected in any of the 70 studies. The studies included 29 types of malignancy (breast, lung, brain, etc.) that incidentally detected KCN DEGs; KCNMA1 and KCNN4 were the most frequent. The encoded products of the 53 detected KCN DEGs are predominantly H+-sensitive (62.3%), and 52 (74.3%) of the 70 studies identified at least one H+-sensitive KCN DEG. There was a significant propensity for KCN DEGs to encode H+-sensitive family members in studies among the eight categories of seven oncogene and histone/DNA modifier-related scenarios compared to 20 studies in “no category”. A concurrent search of the National Institutes of Health Repository of Molecular Brain Neoplasia Data (REMBRANDT) database for expression levels of KCN genes in invasive glial tumors (glioblastomas versus oligodendrogliomas that are less invasive) of patients on anti-seizure medications showed a significant decrease in KCNJ16 and trends in three other KCNs also encoding H+-sensitive proteins in glioblastomas. Correlation with a genomic study of six glioblastoma cell lines [22] confirmed relatively decreased RNA expression of KCNJ16 among KCN genes and provided additional data. Altered expression of KCN family members, possibly compensatory, and non-KCN factors may provide quantities of K+ sufficient for preservation of cell membrane integrity and K+/H+ exchange on inner membranes to release bound H+ for pH reversal and invasion. The aim of this study was to search for potassium’s functional relationships in cancer.

2. Materials and Methods

2.1. Materials and Cell Culture

Chemicals and reagents were from Sigma-Aldrich (RRID:SCR_008988), St. Louis, MO, USA, unless otherwise stated. Anti-Met was from Cell Signaling Technology, Beverly, MA, USA, Cat# 4560, RRID:AB_2143887), 1:1000 dilution, for the beta Met subunit (145 kDa). Diff Quik (Allegiance, McGaw Park, IL, originally Harleco, now Sigma-Aldrich’s Hemacolor Solutions I, II, III [137]. A172 (RRID:CVCL_0131), C6 (TKG Cat# TKG 0242, RRID:CVCL_0194), F98 (ATCC Cat# CRL-2397, RRID:CVCL_3510), LN18 (RRID:CVCL_0392), LN229 (RRID:CVCL_0393), T98G (RRID:CVCL_0556), U87 (RRID:CVCL_0022), and U373 (RRID:CVCL_2219) glioblastoma/astrocytoma cell lines (human (A172, LN18, LN229, U87, U373) and rat (F98 and C6)) were obtained from the American Type Culture Collection, Manassas, VA, USA, RRID:SCR_021346. No cell lines were generated. Cells were maintained in Minimal Essential Media (MEM) with Eagle [138] salts (Cellgro, Mediatech, Herndon, VA, USA) with 10% fetal bovine serum, FBS, (Invitrogen, Carlsbad, CA, USA).

2.2. Statistics

Pearson product-moment correlation, R, (p values for true correlation > 0), Student’s t-Test, paired and unpaired, and Fisher’s exact test for count data were used in Excel Statistics (RRID:SCR_017294) or R Project for Statistical Computing, v4.3.1 (RRID:SCR_001905).

2.3. Databases

For gene information, the National Institutes of Health, https://www.ncbi.nlm.nih.gov, Online Mendelian Inheritance in Man (OMIM) (RRID:SCR_006437), https://omim.gov, GeneCards (RRID:SCR_002773) www.genecards.org (including protein interactions), and Ensembl https://www.ensembl.org/ (RRID:SCR_002344) databases were used. The National Library of Medicine’s PubMed database, https://pubmed.ncbi.nlm.nih.gov/, was searched for numbers of publications, using word searches for subjects of interest. OMIM was searched for KCN genes expressed in the nervous system. The Repository for Molecular Brain Neoplasia Data (REMBRANDT) (RRID:SCR_004704), NIH, included in the Georgetown Database of Cancer (G-DOC), https://sites.google.com/georgetown.edu/g-doc/home or the G-DOC Hub platform, https://gdochub.georgetown.edu/, a project of the Georgetown Lombardi Comprehensive Cancer Center, Washington, D.C., was used to provide translational research tools [20,21]. It was queried for gene expression of KCN genes, selected cancer genes, and housekeeping genes (HKGs). Medians of gene reporter probes in Adobe Flash (RRID:SCR_017258) were available through 31 December 2020. The genomic results of six glioblastoma cell lines were obtained from the supplemental materials of Patil et al. [22].

2.4. Cell Settling Assays

Human glioblastoma cells (U87, LN229, A172, T98G) maintained in culture conditions were detached with trypsin on the day of assay, and allowed to recover for 4 h in MEM with 10% FBS at 37 °C in a 5% CO2 incubator. For each assay, a Cell Settling Chamber Kit (Neuro Probe, Gaithersburg, MD, USA), with 8 wells, 7 mm diameter, was assembled. A provided glass slide, 25 × 75 mm, was placed in the recessed bottom plate with pre-punched (for wells) blotting paper, pre-moistened with PBS, a gasket, and the upper plate stacked on top. Thumbnuts secured the top plate. Centrifuged (<500 rpm, 5 min) cells were resuspended with gentle pipetting; 500,000 per ml in culture media with 10% FBS. Cell suspensions were added, at 50 μl per well. After 30 min, with media visibly absorbed by blotting paper, the chamber was disassembled, and the slide was removed and stained with Diff Quik followed by rinses in tap water. At least two replicate wells per cell line were prepared. The stained cells were examined microscopically. The cell vacuoles in U87 cells (single and clustered) were counted in settled monolayers. When the vacuoles collectively had a diameter of at least one third of an average nuclear diameter in the same 100× microscopic field, the cell was positive for vacuolization (surrogate of degeneration). The cells were analyzed in a midline sweep of 100× microscopic fields across each well. Pearson product-moment correlation coefficient, R, determinations were performed for the percentage of cells vacuolated versus the number of cells in each cluster. Morphology, ,including cell clustering, was noted. An Olympus BH2 microscope (Olympus Corp., Lake Success, NY, USA) with an Olympus FV II digital camera, using the Olympus Microsuite Five Software for Imaging Applications (Soft Imaging System, Lakewood, CO, USA), was used to capture black and white images. An American Optical Spencer Microscope (Buffalo, NY, USA, now Leica Microsystems, Inc. (Deerfield, IL (RRID:SCR_008960)) with an OptikamB9 Digital Camera 10 MegaPixels (Optika, www.optikamicroscopes.com, Via Rigla 32, 24010 Ponteranica (BG), Italy) with Adobe Photoshop v5.1 and Adobe Photoshop Elements 2024 for Windows (Adobe Systems Incorporated, San Jose, CA, USA (RRID:SCR_014199)) was used to generate color images.

2.5. Glioblastoma Cell Migration Assays

Glioblastoma cell migration assays have been described previously [62,135,136]. The A172, C6, F98, LN18, LN229, T98G, U87, and U373 glioblastoma/astrocytoma cell lines were maintained in culture conditions with half volumes of media changed the day before each assay. Cells were detached with trypsin on the day of the assay and allowed to recover in media with 10% FBS for 2 h at 37 °C in a 5% CO2 incubator. Centrifuged (<500 rpm, 5 min) cells were resuspended 2 million/mL with gentle pipetting in media for migration. Unless otherwise stated, it was MEM (salts include 5.4 mM KCl [138] but it is 5.3 mM according to Thermo Fisher) with 0.1% bovine serum albumin (BSA). Chemoattractant/chemokinetic stimulants, hepatocyte growth factor (HGF), and/or FBS in MEM or PBS media were added to the bottom wells (29–30 μL total volume) as indicated. Then, 48-well modified Boyden chambers (Neuro Probe) were assembled with porous (8 μm, 1000 pores/mm2), air-dried polycarbonate filters, 7 μm thick (Neuro Probe), coated with 0.01% porcine gelatin (solution of 1 mL 10% acetic acid, 1 mL 1% gelatin (from 10% refrigerated stock), and 98 mL H2O). One corner of each filter was cut for orientation before assembly. Cell suspensions, 60 μL, were added to each upper well. The chamber was incubated in a large, covered petri dish with a small amount of water at 37 °C for 5 h in a 5% CO2 incubator. Upon assay completion, the filter was removed and attached to a large clip and stained with Diff Quik, rinsed in 3 beakers of water, and placed on a glass slide (2” × 3”), with migrated cells on the bottom side of the filter. Unmigrated cells were wiped from the upper filter surface with cotton swabs or Kimwipes with only stained migrated cells remaining, trapped between the filter and slide. There were 4 replicates per data point in most assays. Slides with stained migrated cells on the filters were digitized (HP OfficeJet Plus 8715 printer/scanner (Hewlett Packard (formerly RRID:SCR_011873), Spring, TX, USA). Densitometry (UN-SCAN-IT Gel Analysis Software, 7.1 (RRID:SCR_017291) Silk Scientific, Inc., Provo, UT, USA) with correction for background filter staining was used to quantify cell migration. An Epson Perfection 2450 PHOTO (Epson America, Long Beach, CA, USA) was used for obtaining some images of the assays. The first two rows of migrated cell pellets, digitized, were used to calculate HGF cell migration, i.e., second rows of cells stimulated by HGF divided by the first rows of unstimulated cell migration (background), stated as times (X) background cell migration for each cell line. Additional migration assays of U87 cells evaluated motogenic/chemokinetic properties of FBS and HGF by placing them in either the bottom or both top and bottom wells. Separately, the use of MEM versus Dulbecco’s PBS in the bottom wells in a dose curve of FBS (including no FBS for MEM versus PBS) was evaluated with Migration media (Dulbecco’s PBS (contains 2.7 mM KCl, 1.8 mM KH2PO4, 137.9 mM NaCl, 10 mM Na2HPO4) with added 0.210 mM MgCl2, 0.243 mM CaCl2, 5 mM pyruvate, 5.551 mM glucose, and 0.1% bovine serum albumin) in all top wells.

2.6. Immunoblot of Met, Receptor for HGF

One-dimensional gel electrophoresis was performed on lysates from 8 cell lines, equalized for protein (10 μg per lane), in 10% polyacrylamide gels under reducing conditions. The protein standards (MagicMark, MultiMark, Invitrogen (now Thermo Fisher Scientific, Waltham, MA, USA (RRID:SCR_008452)) were 8 μg per lane. The gel was stained with Coomassie blue (Novex Colloidal Blue Stain Kit, Invitrogen, now Thermo Fisher Scientific (RRID:SCR_008452)), photographed, and destained with water. The gel contents were transferred to polyvinylidene difluoride membranes (Invitrogen, now Thermo Fisher Scientific (RRID:SCR_008452)), blocked (Detector Block, Protein Detector Western Blot Kit Lumi-GLO System, Kirkegaard & Perry Laboratories, Gaithersburg, MD, USA, now SeraCare Lifesciences:(RRID:SCR_004535)), and reacted with anti-Met (1:1000), reactive 145 kDa band. The secondary antibody 1:1000 (Kirkegaard & Perry Laboratories, now SeraCare Lifesciences:(RRID:SCR_004535)) was horseradish peroxidase-labeled anti-mouse. Immunoreactive bands were visualized via a luminol-based solution that produced chemiluminescence, and the blot was scanned (Epson Perfection 2450 PHOTO) for densitometry using UN-SCAN-It gel, v5.1.

2.7. Differentially Expressed Genes of Potassium Channels

Potassium channel (KCN) family members were found as differentially expressed genes (DEGs) in genomic landscapes described in 70 cancer studies (1301 authors, 15 journals, 2007–2024). KCN DEGs were detected with non-KCN DEGs in at least one figure (heatmap, Venn diagram, volcano plot, etc.) or table (none from Supplementary Materials). The authors had no initial focus on KCN genes. All issues of Cancer Research (2014–early 2024), Journal of Molecular Diagnostics (2020–2024), and Molecular Cancer Research (2023) were searched visually for papers reporting KCN DEGs. Also, “heatmap” or “heat map” and notation of “open access” were used for cancer articles in archives of Nature Communications (2015–2023) and Cancer Cell (2011–early 2023) to select papers for visual detection of KCN DEGs. Fifty-nine of the 70 studies were systematically selected and others were incidental (files and online listserv notifications). The studies included malignancies (or models) of breast, lung, gastrointestinal, brain, bone marrow, etc. Categories (determined after all E1–E70 studies were selected) were according to descriptions by the authors if there was also a relevant DEG reported in at least one of the E1–E70 studies.

2.8. Expression of Potassium Related Genes in Glial Tumors

REMBRANDT (G-DOC) contained tumor expression data on 67 oligodendroglioma (Oligo) and 220 glioblastoma (GBM) patients, including subgroups treated with anti-seizure (AS) medications, i.e., 24 Oligo-AS and 75 GBM-AS. Gene expressions were determined by reporter probes on microarrays, with median values provided. Differences in median values for multiple probes (when available) were found with t-tests, paired via reporter probes for the two comparison groups when possible, and otherwise, unpaired t-tests were performed. Housekeeping genes (HKGs) that lacked relevance for glial tumors were analyzed for comparisons. Additionally, Oligo-minus-GBM (OMG) differences were calculated for each probe. OMG differences for each gene of interest were compared to OMG differences for HKGs (B2M, PPIA, RRLP0, and YWHAZ). In OMG comparisons, the t-tests were for unpaired samples with unequal variance. Published studies [22,139] were used for confirmation.

3. Results

3.1. Settled Glioblastoma Cells with Fluid Removed

At the end of each cell settling assay, the extracellular fluid had been wicked away. Cell groups were viewed macroscopically; see Appendix A (Figure A1). For each cell line (A172, LN229, T98G, and U87), a striking feature was clustering of robustly viable tumor cells around single degenerating cells (same lineage) in low numbers. Tight clustering of viable cells suggested an opportunistic attraction to dying single cells, possibly their released contents. Some viable cells clustered around single dying cells stained basophilic (Figure 1A,B). Narrow linear spaces with a few thin strands connecting cells were seen in black and white photographs. Cells, single and in groups, without visibly dying centralized cells were also seen. Cytoplasmic vacuolization was most prominent in U87 cells (Figure 1A–C). Vacuole numbers were highest in single cells and decreased as cell clusters increased in size. Representative single and grouped U87 cells, layered together with white strokes around each layer for delineation, are shown in Figure 1C. For counts of vacuolated cells in two separate wells, 51% and 72% of single cells were vacuolated, whereas vacuolated cells accounted for only 11% and 18% in cell groups of greater than twenty cells, first and second wells, respectively. Pearson regressions of cells in clusters with vacuoles had significant R2 values of 0.266 (p = 0.0410) and 0.544 (p = 0.0062), first and second wells, respectively, as shown in Figure 1D. Cells in clusters exhibited less vacuolization/degeneration, possibly due to sharing cytoplasmic contents and contributions of contents from dying cells.

3.2. Cell Migration with HGF and Serum Chemoattraction

Earlier studies of U87 and LN229 glioblastoma cell lines and a clinical specimen’s cells reported chemoattraction to HGF and serum [43,62,134,135,136]. Testing on additional glioblastoma cell lines for migration to HGF and FBS, comparing levels of HGF’s receptor, Met, among them, comparing MEM (provided a positive K+ gradient) versus Dulbecco’s PBS (no K+ gradient), and more extensive testing of chemokinetic effects, are reported here. Boyden assays (48 wells), three per cell line, for two rat astrocytoma cell lines (C6 and F98) and five human glioblastoma cell lines (T98G, LN18, U87, LN229, and U373), showed cell migration to HGF alone as a chemoattractant (row 2), HGF and FBS together as chemoattractants (rows 3–9), and to FBS alone as a chemoattractant in a dose curve (rows 10–12), compared to background (unstimulated) cell migration (row 1). The concentrations of chemoattractants are indicated in the left columns. With the same amount of HGF present, adding increased amounts of FBS increased cell migration in seven of eight cell lines (Figure 2A).
Images of FBS alone as the chemoattractant in a dose curve are shown as migrated cell pellets with densitometry results in bar charts. All cell lines, except A172, showed significant migration to FBS alone. The U87 and U373 results showed impressive stepwise increases in chemotactic migration, with p values < 0.0005 and 0.000005, respectively. U373 chemotaxis to 0.5% FBS, 3.8 times background migration, was highest for all cell lines. A172 cells showed a trend for increased migration to 0.5% FBS; p = 0.0638. Some cell lines showed flattening of responses as levels of FBS increased (Figure 2B). All assays (three per cell line) are shown in Appendix A for C6, F98, T98G, and LN18 (Figure A2) and U87, LN229, A172, and U373 (Figure A3). Cell migration specifically for HGF in each cell line is stated as a multiple (X) of background cell migration (second row of assays’ cell pellet densitometry results divided by corresponding first row of background cell migration); see bottom of Figure 2C. HGF cell migration from highest to lowest for all three assays of each cell line was T98G (2.11X), U373 (1.93X), LN229 (1.52X), U87 (1.49X), C6 (1.33X), A172 (1.30X), LN18 (1.15X), and F98 (1.13X). Twelve of 92 data points showing the migrated HGF cell pellets’ densitometry being slightly less than 1X background cell migration were rounded to 1X. A few replicates did not show visible cell pellets. The immunoblot shows reactivity for HGF’s receptor, Met (β subunit, 145 kDa). Met reactivity with U87 and LN229 [134] is shown here with six other cell lines; there is no correlation with the cell motility results shown below (Figure 2C). The stained gel image showing equal protein loading for cell line lysates in the immunoblot is in Appendix A (Figure A4A).
Another assay used to evaluate MEM for migration studies included gradients of K+ and Cl in migration assays of U87 cells. Both ions are redistributed with volume changes during cell migration. In the absence of FBS, Dulbecco’s PBS in the bottom wells provided K+ as 2.7 mM KCl + 1.8 mM KH2PO4 (4.5 mM K+) and Cl as 2.7 KCl + 137.9 mM NaCl (140.6 mM Cl). MEM in the lower wells provided K+ as 5.3 mM KCl (5.3 mM K+) and Cl as 5.3 mM KCl + 1.8 mM CaCl2-2H2O (8.9 mM Cl). See Methods for a minor issue with KCl molarity in MEM. For all upper wells without FBS, “Migration” media (includes Dulbecco’s PBS) provided K+ as 2.7 mM KCl + 1.8 mM KH2PO4 (4.5 mM K+) and Cl as 2.7 KCl + 137.9 mM NaCl + 0.21 mM MgCl2 + 0.243 mM CaCl2 (141.5 mM Cl). The filter pores between upper and lower wells were 8 × 7 μm (with cells obscuring openings). Potassium did not generate any gradient for cells migrating from upper to lower wells when Dulbecco’s PBS was used in the bottom wells and there was almost no gradient for chloride. However, when MEM was used in the lower wells, a reverse gradient for Cl was generated. With MEM in the lower wells, there was a positive potassium gradient for migrating cells (5.3 mM in lower versus 4.5 mM in upper wells). The positive gradient for K+ may have contributed to significantly increased migration, p = 0.0144, without FBS present. MEM as a medium in the lower wells was also associated with significantly increased cell migration when 0.1 and 1% FBS was present, p = 0.0136 and p = 0.000442, respectively, compared to using Dulbecco’s PBS in the lower wells (Figure 2D). MEM contains 5.6 mM dextrose, 1 mM pyruvate, and 2 mM l-glutamine, whereas the “Migration” media contained 5.551 mM glucose, 5 mM pyruvate, and 0.1% bovine serum albumin as ingredients not already mentioned. Therefore, MEM as a source of fuel in the bottom wells was probably not a significant differentiating factor in the results. MEM also contains 17.856 mM sodium bicarbonate, 0.8 mM magnesium sulfate (more Mg2+ than “Migration” media (0.21 mM MgCl2)), less sodium than Dulbecco’s PBS and “Migration” media, and amino acids, vitamins, etc. An assay image is shown in Appendix A (Figure A4B). These results support the role of K+ in enhancing cancer cell migration.
Earlier studies had noted chemokinesis with HGF and 0.1% FBS in U87 cell assays [136] and chemokinesis with 5% FBS alone in LN229 cell invasion assays [135]. The dose curve for FBS includes additional data points, as shown in Figure 2E. The chemoattractant strength of HGF alone (row 2), and with 0.1% and 1% FBS (rows 5 and 8, respectively) was enhanced by both FBS concentrations in U87 cells. The chemokinetic strength of HGF alone (row 3), with 0.1% and 1% FBS (rows 6 and 9, respectively), was also enhanced by FBS. Chemotaxis and chemokinesis with 0, 0.1, and 1% FBS demonstrated strong enhancements provided by FBS versus HGF alone in either type of U87 cell migration. The combination of HGF, 2.5 ng/mL, and 1% FBS for chemoattraction significantly, p = 0.012, outperformed their chemokinetic effects on cell migration (Figure 2E). A scanned assay image is shown in Appendix A (Figure A4C).
Variable background migration (Figure 2A,B) and assay-to-assay variation for each cell line, as shown in files in Appendix A (Figure A2 and Figure A3), are routine for cell migration assays.

3.3. KCN Genes Differentially Expressed in 70 Studies (E1–E70)

3.3.1. General Features of Detected Differentially Expressed KCN Genes

Availability of K+ for glioblastoma cell fitness and migration led to consideration of KCN genes being altered in cancer cells. Genomic landscapes, i.e., heat maps (46 studies), Venn diagrams, volcano plots, tables, etc., in the 70 cancer studies included one or more of 53 KCN differentially expressed genes (DEGs) with co-expressed non-KCN DEGs. Selection criteria (mostly p < 0.05 differences, chance of false discovery, etc.) were set by the authors, who incidentally detected KCN DEGs among groups of DEGs, as listed in Appendix B (Table A1, sample view; entire list is in Supplementary Materials: Table S1: FullTblB1DEGalphaAZ). Parameters were applied for a particular type of malignancy, without or with perturbations, such as tumor treatment, transcription factor, over-expression of oncogene(s), survival, disease progression, primary versus metastasis, recurrence, etc. The 53 KCN DEGs are from 17 chromosomes (1–3, 5–12, 17–21, and X). Many encode membrane pore subunits with selectivity for K+ and some provide modulatory/regulatory effects. Subfamilies, Kv (voltage dependent, KCNA, KCNQ, KCNH, etc.), Kir (inwardly rectifying, KCNJ), K2P (two pore domain, KCNK), and KCa (calcium dependent, KCNM, KCNN, etc.) were found in 32 (46%), 10 (14%), 14 (20%), and 29 (41%), respectively, of the E1–E70 studies (some with multiple subfamily members).

3.3.2. Proton (H+) Sensitivity of Detected KCN DEGs

In keeping with K+ and H+ being monovalent cation relatives, the H+ sensitivity of KCN DEG-encoded proteins is relevant. At least 33 (62.3%) of the 53 KCN DEGs encode gene products that are H+-sensitive (130 publications in Appendix B, Table A2). Fifty-two (74.8%) of the 70 studies contained at least one KCN DEG identified as H+-sensitive. One, two, three, and six H+-sensitive KCN DEG genes were found in 36, 14, 1, and 1 of the studies, respectively. Genes for H+-sensitive proteins in Kv, Kir, K2P, and KCa subfamilies were found as DEGs in 17 (24%), 9 (13%), 13 (19%), and 21 (30%) of the 70 studies, respectively. The most frequently detected DEGs, KCNN4 and KCNMA1, present in nine and six studies, respectively, encode H+-sensitive proteins (red H+ above columns) and are also Ca2+-sensitive (Figure 3A).

3.3.3. Coding and Multiplicity of KCN DEGs and Non-KCN DEGs

Four of the 70 studies identified only non-coding KCN DEGs: KCNQ1-OT1 (E29, E33, and E66) and KCNQ5-IT1 (E30). Study E27 identified KCNMB2-AS1 and KCNE4 as DEGs. Multiple (two to nine) KCN DEGs were found in 18 (27.3%) of 66 studies with protein-coding KCN DEGs. Thirty-eight (71.7%) of 53 KCN DEGs were detected with other coding KCN DEGs in 66 of the studies. There were 102 instances of 53 KCN DEGs in 3239 DEG instances (2490 without repeats). The average numbers of KCN DEGs and all DEGs per study were 1.53 +/− 0.33 (95% CI) and 45.91+/− 7.92 (95% CI), respectively. The correlation coefficient for the number of KCN DEGs detected and size of the DEG groups was 0.379. Fifty-two (78.8%) of 66 studies with coding KCN DEGs detected at least one H+-sensitive KCN DEG (Figure 3A).

3.3.4. Malignancies in E1–E70

The three most frequent malignancies in E1–E70 were lung cancer, breast cancer, and glioblastoma, with detected genes encoding H+-sensitive products bolded as follows. Lung cancer studies included E3 (KCNK1), E8 (KCNK1, KCNN4), E12 (KCNH8, KCNMB2), E23 (KCNU1), E30 (KCNQ5-IT1), E35 (KCNK1), E43 (KCNK6), E49 (KCNN4), 2), E54 (KCNJ13), E63 (KCNK12), and E67 (KCNMB2). Breast cancer studies included E1 (KCNMB1), E4 (KCNN3), E9 (KCNH1, KCNK15), E15 (KCNT2), E22 (KCNK5, KCNK15), E33 (KCNQ1-OT1), E38 (KCNMA1), E48 (KCNJ9), and E63 (KCNK12). Glioblastoma studies included E5 (KCNJ10, KCNN3), E21 (KCNAB2), E26 (KCNE2, KCNJ13), E27 (KCNE4, KCNMB2-AS1), E31 (KCNJ10), E50 (KCND3), and E51 (KCNIP1, KCNJ16, KCNN2). H+-sensitive coding KCN DEGs were in 86%, 73%, and 44% of 7 glioblastoma, 11 lung cancer, and 9 breast cancer studies, respectively. All three melanoma studies detected H+-sensitive KCN DEGs. Numerous types of malignancy detected H+-sensitive KCN DEGs; see Appendix B (Figure A5A). Informative KCN protein abbreviations are listed in Figure A5B. E1–E70 citations for E1–E70 are given in Section 3.3.6.

3.3.5. Stochastic View of K+-Related Dynamics in E1–E70

A stochastic view (percentages) of K+-related gene expression in cancer cells is presented based on incidences of KCN and related non-KCN DEGs. Images of the most frequent members of the KCN subfamilies with pertinent non-KCN proteins are consistent with literature representations [140,141,142,143,144,145,146,147]. The relatively small number of 70 studies with differential expressions detected for genes, shown in Figure 3B, that encode supporting systems, i.e., Na/K-ATPase (9% for five genes), monocarboxylate transporter complex (5% for three genes), lactate dehydrogenases (3% for one gene), taurine transporters (2% for two genes), and 0% for genes encoding functional analogues not listed, support the implication of abnormal K+ flux in cancers with KCN DEGs forcing K+/H+ exchange to initiate pH changes. Rapid cytosolic H+ transport (putatively Grotthuss water wires) to exit sites for proton release with lactate completes the pH reversal outcome. Pertinent DEGs lacking in E1–E70 include glycogen synthases that could mediate anomalous storage of lactate without extrusion, since its storage as glycogen occurs in humans, etc. [148,149,150,151]. Accumulations of K+ may happen due to genetic loss or downregulation of KCN proteins mediating outward flux or increased inward flux due to gains or losses in copy number and/or expression of KCN DEGs in E1–E70. Movement of K+-7H2O within cells may bring water to inner membranes, possibly promoting Grotthuss conditions. Putative K+-related changes in cells that maintain relatively normal expression levels of pertinent non-KCN genes, as seen in most of the KCN DEG studies, support the processes shown in Figure 3B.

3.3.6. Categories Among Studies with E1–E70 Citations

Eight categories describe 50 of the 70 studies (Figure 4A). These were designated by the authors and accepted if there were relevant DEGs in any of the E1–E70 studies. Some studies were in multiple categories and 20 studies were in no category.
1. Of 18 studies in the first category of histone/chromatin/epigenetic (His/Chr/Epi), 10 had category-specific genes described by the authors, including E2 (N6AMT1, previously KMT9), E13 (MSH2, BET), E32 (H3K27me3, EZH2), E41 (HDAC8), E46 (KDM2B), E49 (KMT2D), and E53 (H3K27). Also, E27 and E29 focused on chromatin access and imprinting, respectively. A total of 21 imprinted DEGs (AXL, DLK1, GDAP1L1, GNG7, GRB10, IGF2, KCNQ1, KCNQOT1, LRRTM1, MEG3, NDN, NNAT, NTM, PAX8, PEG3, PEG10, PHLDA2, PLAGL1, RHOBTB3, RTL1, and TFPI2) accounted for 32 instances of imprinted genes among 3239 DEG instances (repeats included) for a rate of 0.99% in E1–E70. Detection of 0.99% imprinted genes is expected from the genome (approximately 200, i.e., <1% of all genes) [152]. Therefore, imprinted genes were not used as an inclusion criterion. Eight additional studies detected His/Chr/Epi-related DEGs as follows: E4 (DNMT3B), E30 (KDMA-AS1), E31 (H2A.X, previously H2AFX), E37 (H2AC18), E43 (H2AC18, H2BC5, H2BC12), E45 (NUPR1), E58 (DNMT3B), and E66 (BDR4, H1-4 (previously HIST1HIE)). The His/Chr/Epi category includes the following studies (with citations): E2 [153], E4 [154], E13 [155], E27 [156], E29 [157], E30 [158], E31 [159], E32 [160], E37 [161], E41 [162], E43 [163], E45 [164], E46 [165], E49 [166], E53 [167], E58 [168], E59 [169], and E66 [170] (Figure 4A, left histogram).
2. In the second category there were nine studies on TP53 according to the authors, in E6, E8, E11, E49, E49, E60, E63, E64, and E67. Additionally, three studies found TP53 and related genes as DEGs as follows: E18 (MDM2), E43 (TP53INP1A), and E50 (TP53). The twelve studies in this category are E6 [171], E8 [172], E11 [173], E18 [174], E43 [163], E49 [166], E50 [175], E60 [176], E63 [177], E64 [178], E66 [170], and E67 [179] (Figure 4A, left histogram).
3. In the third category there were seven studies on epithelial mesenchymal transition (EMT)/TGFb/mesenchymal tumor (EMT/TGFb/Mes) according to the authors, with specific genes CDH1,2, VIM, and TGFb1,2,3 in studies as follows: E16 (CDH1,2, TGFb1, VIM), E24 (TGFb1), E31, E38, E39, E43 (CDH2), and E46. Three additional studies also found these genes: E21 (TGFb1), E66 (VIM), and E70 (CDH2, TGFb2, TGFb3). The ten studies in this category are as follows: E16 [180], E21 [181], E24 [182], E31 [159], E38 [183], E39 [184], E43 [163], E46 [165], E66 [170], and E70 [185] (Figure 4A, left histogram).
4. In the fourth category there were five studies on EGFR and related genes according to the authors, with corresponding DEGs in the studies as listed: E9, E48, E50 (EGFR, ERBB2), E51 (EGFR), and E52. Additionally, four studies found related DEGs as follows: E40 (NRG1), E54 (NRG2), E59 (NRG1), and E60 (ERBB4). The nine studies in this category are E9 [186], E40 [187], E48 [188], E50 [175], E51 [189], E52 [190], E54 [191], E59 [169], and E60 [176] (Figure 4A, left histogram).
5. In the fifth category there were five studies on MYC according to the authors, with MYC as a DEG in two of these studies: E14 (MYC), E30, E31, E55, and E61 (MYC). Additionally, three studies had MYC or related genes: E12 (MYCL), E17 (MYCN), and E35 (MYC). The eight studies in this category are E12 [192], E14 [193], E17 [194], E30 [158], E31 [159], E35 [195], E55 [196], and E61 [197] (Figure 4A, left histogram).
6. The sixth category was comprised of six studies on RAS per the authors’ interest, with KRAS found as a DEG in two of the studies: E8, E43, E45, E49, E64 (KRAS), and E69 (KRAS). The six studies in this category are E8 [172], E43 [163], E45 [164], E49 [166], E64 [178], and E69 [198] (Figure 4A, left histogram).
7. In the seventh category the authors focused on BCL2 in E47 but without listing it as a DEG. However, other studies identified BCL2 as a DEG, and one also identified BCL2L1, as listed: E12 (BCL2), E22 (BCL2), E35 (BCL2), and E70 (BCL2, BCL2L1). The five studies in this category are E12 [192], E22 [199], E35 [195], E47 [200], and E70 [185] (Figure 4A, left histogram).
8. In the eighth category, the authors of one study focused on MET and found it as a DEG, E23 (MET). Three studies also identified MET as a DEG, E59 (MET), E62 (MET), and E67 (MET). The four studies in this category are E23 [201], E59 [169], E62 [202], and E67 [179] (Figure 4A, left histogram).
All the studies in the eight categories are shaded yellow in Figure 4.
9. No category. Twenty studies were not in any of the eight categories: E1 [203], E3 [204], E5 [205], E7 [206], E10 [207], E15 [208], E19 [209], E20 [210], E25 [211], E26 [212], E28 [213], E33 [214], E34 [215], E36 [216], E42 [217], E44 [218], E56 [219], E57 [220], E65 [221], and E68 [222] (Figure 4A, right histogram, no yellow shading).

3.3.7. H+ Sensitivity in E1–E70 Studies Within Eight Categories

The histogram in Figure 4 (right side) shows studies collectively in the eight categories versus those not in any category, with studies that detected H+-sensitive KCN DEGs underlined with red. The difference in detection of H+-sensitive KCN DEGs was significant between the two groups: p = 0.0325, Fisher’s exact test for count data. All eight categories were among those shared in the studies (Figure 4A, right histogram, yellow shading). The citations for E1–E70 are listed in Section 3.3.6.

3.3.8. PubMed Citations for “Proton”, “Potassium”, Etc., in Categories

Word searches of PubMed for oncogene and histone publications related to those for H+, K+, BK (KCNMA1), and Cl (contrasting ion), without and with “cancer”, were performed to further evaluate the categories. On 12 May 2025, searches for each category of TP53, MYC, TGFbeta, BCL2, EGFR, histone, RAS, and MET were performed separately for “proton”, “potassium”, and “chloride”. For all eight categories, the total citations for “proton”, “potassium” and “chloride” were 1565, 1418, and 1261, respectively, and the Pearson’s product-moment correlations for “proton” with “potassium” and “chloride” were R = 0.891 (p = 0.00148) and 0.599 (p = 0.0584), respectively. Pearson’s product-moment correlation for “potassium” with “chloride” was 0.689 (p = 0.0295). The R2 for proton with potassium was 0.79. Later searches on June 7 2025for “BK” and “potassium” revealed “BK” citations were about one-third of those for “potassium” in seven categories (TP53 was much greater); they are listed below the x-axis in Figure 4B. When searches also included “cancer” as a third word, the numbers of citations for “proton”, “potassium”, “BK”, and “chloride” on 14 May 2025, were 1061, 577, 513, and 952, respectively. However, when MET was left out, the numbers of citations for “proton”, “potassium”, “BK”, and “chloride” on 14 May 2025, were 728, 496, 439, and 879, respectively, and the Pearson’s product-moment correlations for “proton” with “potassium” and “chloride” were improved, R = 0.860 (p = 0.00649) and 0.462 (p = 0.4623), respectively, using only seven categories.. The R2 for “proton” with “potassium” for seven categories after MET removal was 0.74. The Pearson’s product-moment correlation across the seven categories for “proton” with “BK” was R = 0.901 (p = 0.00283). The R2 for the correlation of “proton” with “BK” across the seven categories was 0.81. It is unknown why citations in the MET category no longer resulted in the correlation seen in previous searches without “cancer” for “proton” and “potassium”, but the possibilities include unpublished negative studies. The answer is beyond the scope of this study. However, see Figure 2C for negative results for Met in this study. Also, compared to searches without “cancer”, as shown in Figure 4B, “BK” citations appeared to constitute an even larger proportion of the “potassium” searches when “cancer” was included (Figure 4C). Numerous citations were shared pairwise in two-word searches (not restricted to cancer) among the eight categories of seven oncogenes and histone (Figure 4D). Thus, studies involving the eight categories may often describe them collectively. Citation searches for KCN proteins other than BK yielded relatively low numbers.

3.3.9. Non-KCN DEG Repeats in E1–E70

Non-KCN DEGs found in multiple (five to seven) studies were often detected with KCNN4 and some with KCNMA1. Among non-KCN repeat DEGs, ITGA2/Integrin Alpha-2 was most frequent (seven studies). Six also included H+-sensitive coding KCN DEGs (bolded): E8 (KCNK1, -N4), E15, E39 (KCNMA1), E41 (KCNMA1, -N4), E49 (KCNN4), E64 (KCNN4), and E69 (KCNN4). KCNN4 was the most frequently found DEG, nine times in total and in five of seven studies with ITGA2. Studies of co-expressed KCNN4, ITGA2, and PLAUR, E8 [172], E41 [162], E49 [166], and E69 [198], included two lung cancers, one melanoma, and one pancreatic cancer, in the categories of His/Chr/Epi, TP53, and RAS (Figure 5A, left Venn diagram). HMGA2/High Mobility Group AT-hook 2, LIF/Leukemia Inhibitory Factor, and PLAUR/Plasminogen Activator, Urokinase Receptor (UPAR) were DEGs found in six studies each. These studies also frequently detected KCNN4 as a DEG.
One study detected KCNN4, LIF, and PLAU (Figure 5A, right Venn diagram). Two studies detected KCNN4, ITGA2, and PLAU (not shown). The PLAUR and PLAU DEGs were found together in 1 study among 10 studies that detected either of them. E41 [162] detected KCNMA1, KCNN4, ITGA2, and PLAUR (not shown).
Searches of PubMed for shared citations for “BK” with “PLAUR” + “PLAU” (combined separate searches) and “ITGA2” in three-word searches including “cancer” found significant Pearson’s product-moment correlations of R = 0.742 (p = 0.0175) and 0.718 (p = 0.0224), respectively, and R2 = 0.55 and 0.52, respectively, across the eight categories (Figure 4B).
Also, some of the non-KCN DEGs found in E1–E70 are shared with studies of genes encoding circulating proteins associated with cancer risk. The shared DEGs from the E1–E70 studies, with their frequencies, are FGFR3 (5X), GDF15 (4X), IGFBP3 (5X), IGFBP7 (4X), KRT19 (4X), and PLAUR (6X) [223,224].

3.3.10. Non-KCN DEG Superfamilies in E1–E70

Nine superfamilies of non-KCN DEGs were detected in E1–E70, with at least 15 occurrences, including 74X for SLC (Solute Carrier), 32X for COL (Collagen), 27X for ITG (Integrin), 22X for TMEM (Transmembrane), 21X for CD (Cluster of Differentiation), 19X for GPR (G-Protein-coupled Receptor), 18X for SOX (Sry-related HMG bOX), 15X for SERPIN (Serine Proteinase Inhibitors), and 15X for TNFRSF (Tumor Necrosis Factor Receptor Super Family).
The largest numbers of repeat DEG occurrences from superfamilies were ITGA2 (previously mentioned, 7X), SOX2 (5X), and SLC1A2/Excitatory Amino Acid Transporter 2 (EAAT2) (4X). EAAT2 clears glutamate from extracellular space at synapses [225,226,227]. Six Regulators of G protein Signaling (RGS) family members were found in seven studies. Eighteen (25.7%) of E1–E70 had either GPR- or RGS-family member DEGs. Among GPR DEG proteins, none are known as being H+-sensitive; see Appendix B (Text A1).

3.3.11. KCN H+ Sensitivity in E1–E70 and Cancer Reviews

Ten reviews describing KCN family members in malignancy were compared with detection of KCN members in E1–E70. The numbers of KCN genes described in the reviews are as follows: 7 for glioblastomas [228], 3 in a report for a cancer colloquium [229], 23 for 18 cancer types [230], 13 for 10 cancer types [231], 9 for 17 cancer types [232], 21 for 14 cancer types [233], 12 for 13 cancer types [234], 15 for 7 cancer types [235], 13 for 13 cancer types [236], and 13 for 11 cancer types [237]. Twenty KCNs found in 35 (half) of the E1–E70 studies overlap with those in the reviews: KCNA1, A3, A5, D1, H1, H2, H7, J5, J8, J10, K2, K3, K5, K15, MA1, N2, N3, N4, Q1, and T2. The ten reviews combined with E1–E70 highlight KCN genes that encode H+-sensitive proteins. At least one KCN gene in each of 29 studies (82.9% of 35 overlapping E1–E70 studies) encodes an H+-sensitive protein; these studies are E4, E5, E6, E8, E9, E12, E13, E16, E22, E24, E28, E31, E32, E34, E36, E38–E40, E41, E47, E49, E52, E53, E58, E60, E64, and E68–E70. The remaining studies include E15, E29, E33, E45, E65, and E66. Citations for E1–E70 are given in Section 3.3.6.

3.3.12. Long Introns in E1–E70’s KCN DEGs

Many KCN genes have long introns (>20,000 bp). Long introns may affect gene expression and regulation and can be genetically damaged. Thirty-one (58.5%) of 53 KCN DEGs in E1–E70 contain long introns, some very long. Seven KCN DEGs (KCNC2, D3, H1, H8, MA1, Q1, and T2) contain at least one intron 100,000–199,999 bp long. Five KCN DEGs (KCNH7, IP1, Q3, MA1, and MB2) contain at least one 200,000–299,999 bp intron and KCNB2 and KCNQ5 contain at least one intron 300,000–399,999 bp long. Fourteen KCN DEGs in E1–E70 contain multiple long introns. KCNMA1 and KCNT2 each contain seven long introns. KCNN4 is among the KCN DEGs without any long introns. Forty-one (58.6%) of the E1–E70 studies contained at least one KCN DEG with at least one long intron. Nine studies had multiple (two to four) KCN DEGs with long introns. However, KCNA3, A5, and F1 in E1–E70 do not contain any introns. Intron data from Ensembl was last accessed on 22 June 2025. The contrast between KCNMA1 (seven long introns) and KCNN4 (no long introns) is notable.

3.4. Study of Selected KCN DEGs in REMBRANDT

3.4.1. Preliminary REMBRANDT Survey

The possibility of potassium acting as a chemoattractant in the CNS led to a search for KCN DEGs in the brain tumor database REMBRANDT while gathering the E1–E70 studies. Initial queries in the Online Mendelian Inheritance in Man (OMIM) database, 2018–2019, and literature descriptions of KCN genes [238], yielded 41 KCN genes with neuronal and astrocytic expressions. All 41 genes are in loci corresponding to known regions of chromosomal losses, gains, or both in glioblastomas [1]. These genes are distributed as follows: Chromosome (Ch) 1 (KCNA2, A3, C4, N3, J9, J10, K2, K1); Ch2 (KCNJ3, J13); Ch3 (KCNMB2, MB3); Ch5 (KCNN2, MB1); Ch6 (KCNK17); Ch8 (KCNQ3); Ch9 (KCNT1); Ch10 (KCNMA1, K18); Ch11 (KCNJ11, C1, A4, K4, J5); Ch12 (KCNA1, A5, A6, J8, MB4); Ch14 (KCNK10); Ch17 (KCNJ12, J16, J2); Ch19 (KCNN1, N4, C3); Ch20 (KCNB1, Q2); Ch21 (KCNJ15, J6); and Ch22 (KCNJ4). Most numerous were the KCNJ subfamily genes (KCNJ2, 3, 4, 5, 6, 8, 9, 10, 11, 12, 13, 15, and 16) encoding inwardly rectifying potassium channel forming proteins and modulators/regulators; the Kir2.1, 3.1, 2.3, 3.4, 3.2, 6.1, 3.3, 4.1, 6.2, 2.2, 7.1, 4.2, and 5.1 proteins. Kir channels can be occluded by Mg2+ or polyamines when cell membrane potentials are positive relative to the equilibrium potential of K+, allowing only a small outward current. At membrane potentials that are negative relative to the equilibrium potential of K+, Mg2+ and polyamines leave the channels by flowing into the cell, so a large inward K+ current can occur [239,240]. Kirs, including Kir4.1, may also extrude K+. Kir 4.1 and 5.1 are present in astrocytes and oligodendrocytes [39,241].
A preliminary survey of REMBRANDT for expression of 21 KCN family members, 12 other potassium flow related genes, and 4 housekeeping genes (HKGs) checked for differences in glioblastomas from patients with and without a history of seizure medications. Seizures could be an indicator for abnormal levels of potassium [242,243,244]. With “unknowns for gene expression” included, there were 84 and 18 glioblastomas with and without a history of anti-seizure medication, respectively. Compared to the absolute difference in the expression of HKG (PPIA, RPLP0, YWHAZ, and B2M) between the two groups, single or average absolute median values of reporter probes for KCNJ10, KCNK3, ATP1A1, KCNN3, KCNJ13, KCNMB2, SLC12A7/KCC4, KCNN2, KCNMB1, GJA1, KCND2, KCNK9, SLC1A2/EAAT2, KCNJ2, and KCNJ16 exceeded the average HKG difference (0.1131) of median values. Others, including ATP1A2, KCNA2, KCNA3, KCNK1, KCNIP3, KCNMB3, SLC1A3/EAAT1, SLC12A2/NKCC1, KCNT1, KCNMA1, LRRC55, KCNA1, KCNMB4, GJB2, GJB6, SLC12A6/KCC3, and FXYD7, did not exceed the average HKG difference in expression. Preliminary data suggested that the subset of glioblastomas in patients with seizures (medication history) had expressions of KCN family member genes that could be compared to the same genes in oligodendrogliomas (less invasive glial tumor) from patients also on seizure medications. Abnormalities in gene expressions among KCN family members resulting in potassium flux could be related to malignant behavior differences between oligodendrogliomas and glioblastomas.

3.4.2. Gene Expressions Reflecting Cell of Origin

Checks for expected findings were performed. Expression of ATP1A2 (alpha subunit of Na/K-ATPase specific for glial cells) was higher in both glial tumors versus neuronal-specific ATP1A1 in patients on anti-seizure medications. The medians of two reporter probes available for each gene yielded significant differences, p = 0.0345 and 0.0278, respectively, for oligodendrogliomas and glioblastomas using two-tailed, unpaired t-tests (Figure 6A). Specificity of expression for connexin genes in astrocytic cells, GJA1/Cx43, GJB6/Cx30, and GJB2/Cx26, was used to verify expected cells of origin for the glial tumors, oligodendrogliomas, and glioblastomas [241,245,246] in REMBRANDT patients on anti-seizure drugs. The median values of single-reporter probes for expression of astrocytic-type connexins, collectively as a group, were significantly greater in glioblastomas; p = 0.0477 (one-tailed, paired t-test) (Figure 6B).

3.4.3. KCN Gene Expression in Patients on Anti-Seizure Medications

Limiting tumor comparisons to patients on anti-seizure medications improved detection of differences between types of invasive glial tumors. Without using the history of seizure medications, a difference in expression of KCNJ16/Kir5.1 in glioblastomas (less expression), p = 0.05, became significant in glioblastomas using the smaller group of patients on anti-seizure drugs; p = 0.0148. Trends also improved for differences in expressions of KCNJ10/Kir4.1, KCNK3/TASK1, and KCNK10/TREK2 (all H+-sensitive, Appendix B (Table A2) and [247,248]) with reporter probe data (Figure 6C) and p values (Figure 6E,F); the same was also shown with CA12/carbonic anhydrase 12 and TP53/p53.

3.4.4. Oligodendroglioma-Minus-Glioblastoma (OMG) Normalization

Housekeeping gene (HKG) normalization corrected for loss of genomic integrity, differences in tissue handling, assay-to-assay variation within and between institutions, etc., in comparisons for patients on anti-seizure medications. Oligodendroglioma-minus-glioblastoma (OMG) differences (increased or decreased) were calculated for each reporter probe’s median expression. These OMG differences for each gene of interest were compared to the OMG differences for all four HKGs, resulting in better p values (Figure 6D–F). The OMG difference for KCNJ16, decreased in glioblastomas, became highly significant, p = 5.532 × 10−8, and trends for other genes improved. These t-tests were calculated for unpaired samples with unequal variance. Median data for probes via Adobe Flash, shown in Figure 6A–D, are no longer available online (see Section 2.3).

3.5. KCN DEGs Shared in E1–E70, REMBRANDT, and Glioblastoma Cell Lines [22]

Whole-exome and RNA sequencing data of LN18, LN229, T98G, U87, U343, and U373 human glioblastoma cell lines in Patil et al.’s supplementary materials [22] revealed data for many KCN DEGs and genes of interest found in the E1–E70 and REMBRANDT studies. Also, three and five of their cell lines were used for the cell settling and migration studies, respectively, described earlier. Their data availability is described in the Data Availability Statement following this article’s text. The methodology of Patil et al. is described in [22], with some information included here. For each GBM cell line, a 62 Mb genome region was targeted. Alignment was to human reference genome hg19. A single-nucleotide variant (SNV) was a cancer-specific mutation (CSM) if present in the COSMIC database (reference 18 in [22]) and/or was present in GBM TCGA exome sequencing data (references 8 and 9 in [22]). Data for four cell lines, U87, T98G, LN229, and LN18, had an average concordance of 65.3% with COSMIC (reference 18 in [22]) and 74.7% with the Broad Institute’s CCLE database (reference 19 in [22]). Patil et al. reported small insertions and deletions (indels), −49 to +29 bases. Their differential gene expression data was obtained via comparisons of their RNA-seq data from the six GBM cell lines with expression data of five normal brain samples from TCGA, with a cut-off value of a 2-fold change in absolute expression. Their results for 2214 genes compared with those also present in TCGA GBM’s tumor samples had 73% concordance. Of the 2214 genes from cell lines, 1831 (82.7%) were downregulated. For the 1621 genes similarly up- or downregulated in both the TCGA and Patil et al. studies, 79.6% were downregulated.
Forty-two KCN genes in their cell lines had CSMs and indels. DNA changes in 18 of the 42 KCN genes occurred in at least three glioblastoma cell lines. The KCNJ subfamily had the most members (ten) with DNA changes, including KCNJ1, 2, 3, 5, 6, 10, 12, 13, 14, and 16. Among all KCN genes there were 69 CSMs, representing 4.3% of 1594 CSMs found in all genes sequenced, whereas 68 indels (25 insertions and 43 deletions) that they found in KCN genes represented 1.8% of all 3892 indels found among six glioblastoma cell lines. KCN DEGs in E1–E70 and genes of interest in REMBRANDT were among their KCN genes with CSMs, INSs, and DELs. Twenty-three of the E1–E70 KCN DEGs and four of the REMBRANDT study’s KCN genes of interest had changes described in genomes of six glioblastoma cell lines by Patil et al. Those detected in E1–E70 (glioblastomas (7) and 28 other cancers) are marked with green squares on the right-hand side of lists in Figure 7A, and genes of interest (significant expression difference or trends) found in REMBRANDT are bolded. All 51 expressed KCN genes detected in six glioblastoma cell lines were downregulated and represented 1.5% of all the 3428 differentially expressed genes in the cell lines. The average cell line KCN gene expression levels are listed by descending abundance, including 27 and 4 of those shared with E1–E70 and REMBRANDT, respectively, with the names of the encoded proteins in Figure 7B. KCNJ16, significantly downregulated in REMBRANDT’s glioblastomas compared to oligodendrogliomas, was among the genes with the lowest expression in cell lines studied by Patil et al. (Figure 7B). The log2 fold RNA changes from normal calculated using their supplementary table S6 [22] are graphed for 51 genes, with 27 genes (52.9%) that encode H+-sensitive proteins and 24 KCN genes with DNA changes highlighted, in Figure 7C. Of the 53 KCNs in E1–E70, there are 38 (71.7%) also described by Patil et al. with DNA changes and/or RNA expression detected. Of seven GBM studies in E1–E70, six (85.7%), E5, E21, E26, E27, E31, and E51, shared 1–3 KCN DEGs with those described by Patil et al. KCNs encoding proteins with H+ sensitivity in E1–E70 and REMBRANDT overlap with those described by Patil et al. (Figure 7C). References for H+ sensitivities are cited in Appendix B, Table A2, and in Section 3.4.3, for the encoded proteins of KCNA1, KCNA3, KCNA5, KCNAB2, KCNH1, KCNH3, KCNJ10, KCNJ16, KCNK1, KCNK3, KCNMB4, KCNN3, KCNQ1, KCNQ2, KCNQ3, KCNQ5, and KCNT1. References for the H+ sensitivity of proteins encoded by additional KCN genes in Figure 7C are as follows: KCNA2/Kv1.2 [249], KCNA4/Kv1.4 [250,251], KCNC4/Kv3.4 [252], KCNH4/ELK1 [253], KCNJ3/GIRK1 [254,255], KCNJ11/Kir6.2, IKATP [256,257], KCNJ12/Kir2.2, HIRK1 [258] (variant), KCNK9/K2P9, TASK3 [259,260], and KCNK17/K2P17.1, TALK2, TASK4 [261].

4. Discussion

4.1. Input of Functional U87 Studies

The decision to stochastically investigate potassium channel genes in cancer papers and a database was based on glioblastoma cell migration to serum and attraction to dying cells in settling assays with fluid deprivation. Tight cell adherence of robustly viable cells to dying settled cells was a surprise. Uncentrifuged filtered cancer cells from clinical fluid specimens can be viewed microscopically after dissolution of filters [262,263]. They do not have this appearance routinely. However, the clustering morphology of viable cells around single necrotic cells resembled rimming of deeply stained cells (“pseudopalisades”) around necrotic foci in glioblastoma tumor tissue, as described in the Introduction. Also, basophilic cells of tumors with central necrosis and more alkaline cytoplasmic pH detected in necrotic areas have been seen experimentally in cervical cancer [264]. Initial clues suggested potassium and pH involvement. Specifically, some viable glioblastoma cells rimming single dying cells, seemingly protected from the negative effect of fluid removal, had relatively darker staining, possibly due to an internal pH change, and as ATP is depleted in dying cells, K+ is released before cells rupture [265,266]. Potassium efflux has been seen in cancer cells [267].
A chemotactic role for potassium has been supported in numerous cell migration assays with a serum constituent acting as a chemoattractant. Also, U87 cells migrated significantly to a positive gradient of K+ generated by MEM, with improvement in the presence of FBS (Figure 2D). Flattening of dose curves and band-like rimming of glioblastoma cells around necrotic regions in tissue suggest that available ranges of potassium levels may need to match whatever functional complement of potassium channels is present in tumor cells. Elucidation of relationships in cancer cells between K+ and H+, possibly to incite reversal of pH, inspired investigation of KCN genes for differential expression in cancer studies and a brain tumor database.

4.2. Mechanistic Proposal for a New Route to pH Reversal

KCN family members, many of which are pH sensitive, cooperate to control potassium levels at cell membranes in a putatively H+-sensitive manner. The hypothesis is that K+, as a monovalent exchange cation for H+, frees bound protons from inner cell membranes for rapid cytoplasmic transfers to proton extruders, and instigates pH reversal in cancer if control of potassium flux is deregulated and rapid delivery of protons to cell exits is possible. The Grotthuss mechanism in protonated water wires is a new consideration to explain how freed protons rapidly transfer to cell exits to achieve pH reversal. Proton transfer via “water wires” has been described for proteins and the blood–brain barrier [268,269,270]. However, at this time cytoplasmic Grotthuss-type proton transfer in cancer cells has not been recognized, so direct mechanistic proof of K+/H+ exchange driving pH reversal is lacking, representing a limitation. However, as migrating tumor cells flatten and squeeze into small spaces, such as between a nerve fiber and its surrounding nerve sheath or along and then through dense basement membranes, bone, cartilage, etc., microaqueous intracellular regions are generated in tumor cells. As cell settling assays removed fluid to create a microaqueous condition for all cells, there was simultaneous release of K+ from single dying cells to surrounding viable cells. These centralized dying cells could also have become “sinks” for protons released from clustered viable tumor cells, possibly with some undergoing pH reversal to appear basophilic while dying central cells became more acidic.
Putatively, entry of excess K+ into viable cells exchanges with H+ bound to inner-membrane and cytosolic anions to release H+ (almost immediately H3O+) into surrounding cytoplasm containing Grotthuss water wires, that can rapidly supply protons to extruders, resulting in pH reversal and associated changes. In tumor tissue in vivo, the initiating potassium flux may be a response to stress or oncogenes to achieve pHi > pHe that acidifies the microenvironment for activation of proteases, etc., permitting cell migration and a higher cytosolic pH that activates alkaline-sensitive proteins, promoting cell fitness. An alkaline pHi leads to maximal tumor cell proliferation coupled with increased glycolysis and adaptation to hypoxia in silico [271] and is associated with increased flux of glycolytic enzymes, migration, and invasion experimentally [272]. Also, when pHi increases, H+ diffusion increases [273], possibly enhancing pH reversal for survival and dedifferentiation as single, migrating cells.

4.3. Potassium and KCN Genes in Cancer with Related Negative Findings

The KCN family offers a variety and range of dynamic distributions of K+ via numerous members, networking, biological milieus, modulation by oncogenes, histone/DNA modifiers, etc. Potassium is supplied by blood pulsing through the vasculature, although locally haphazard in many cancers, body fluids, and by nerves firing action potentials. Formation of synapses between tumor cells and neurons provides K+ as well as other synaptic substances [274]. Activation of proteases by extracellular acidosis that degrade connective tissues increases access to neural and vascular K+ to enhance migration and metastasis. Studies of KCN genes of interest in REMBRANDT, KCN cancer reviews, and six glioblastoma cell lines [22] included many genes encoding proteins with H+ sensitivity, to support similar findings in stochastic detection of KCN DEGs in unbiased E1–E70 cancer studies. KCNs in E1–E70 overlapped with most of the KCN genes of interest in REMBRANDT (KCNJ16, KCNJ10, and KCNK3,, all encoding H+-sensitive proteins), and with 20 KCN genes covered in the ten reviews. At least one KCN DEG in 29 of 35 studies from E1–E70 overlapping with KCNs in the ten reviews encodes an H+-sensitive protein. Also, in support of the REMBRANDT findings, KCNJ16 was predicted to be negatively affected in glioblastoma cells by both Patil et al. and Clark et al. [22,139]. Potassium ions may redistribute in cancer cells to exchange with H+ in pH reversal via differential expression of KCN family members that encode H+-sensitive proteins, possibly responding to oncogenes, genetic damage, etc.
The E1–E70 studies allowed stochastic investigation of H+-related non-KCN DEGs associated with expression of KCN DEGs in malignancy to verify that genes encoding non-KCN components, needed in several steps of pH reversal, were relatively intact by detection of only a few or no DEGs that would be supportive. Multiple proton extruders have been proposed for glycolytic/plurimetabolic-type glioblastoma cells [275]. Adaptations for viability occur in cell settling assays and in some invasive scenarios without extracellular fluid being available. Aquaporin genes, AQP1 and AQP4, whose products mediate water flux, were only found in three of the E1–E70 studies. AQP3, found in four studies, encodes aquaglyceroporin that mediates glycerol, uncharged solutes, and water flux [276]. In the E1–E70 studies, important negatives included non-KCN DEGs representing proton extruders, carbonic anhydrases, taurine and carnitine transporters, etc. Importantly, only a few studies found ATP1A2-encoding Na/K-ATPase as a DEG in E1–E70. Its role in pH reversal, as an active inward K+ transporter, is supported by a report that digoxin, an inhibitor of Na+/K+ ATPase, reduced the size of circulating tumor cell clusters in women with metastatic breast cancer by −2.2 cells per cluster [277].
In support of finding KCNK genes as DEGs in 20% of the E1–E70 studies, with KCNK1/K2P1 being most frequent, loss of KCNK1 was associated with breast cancer cell survival when necrosis was induced [118]. Also, the ten reviews of KCNs in cancer that shared genes with those found in E1–E70, included four KCNK/K2P and four KCa subfamily (KCNM and KCNN groups) genes. Also, KCNN4/KCa3.1/IK/SK4 and KCNMA1/BK were the most frequently detected KCN DEGs in E1–E70.

4.4. E1–E70 Studies Related to Seven Oncogenes and Histone/DNA Modulation

Among the E1–E70 studies, 50 were in eight categories of oncogenes and histone/DNA modulator involvement, with some in multiple categories. Migration of seven glioblastoma cell lines responded to HGF; the ligand of MET’s encoded Met [278,279]. Although there was no correlation of Met with the level of HGF-stimulated migration detected, its role was not completely ruled out with additional studies (multiple Met antibodies, etc.). Others have proposed a role for multiple oncogenes, including EGFR, RAS (KRAS and HRAS), MYC, and TGFβ, that overlap with oncogenes involved in E1–E70, to interact on paths to malignancy in a context-driven manner [280]. Lactate can activate TGFβ [281,282,283,284]. Also, “upregulation of oncogenic pathways” has been described in clinically evident precancerous lesions of 12 different types of malignancies [285]. The collective influence of multiple oncogenes on downstream potassium channels may promote K+ flux via H+-sensitive KCN proteins. KCN studies related to seven oncogenes and/or histone/DNA modulation were significantly (p = 0.0325) more likely to have at least one or more KCN DEG encoding an H+-sensitive protein than the other KCN studies in “no category”. Although oncogenes may promote K+/H+ exchange, leading to pH reversal, this has not been proven.
Retrieval of citation data from PubMed revealed a significant correlation in the numbers of studies of H+ with K+ across all eight categories of the seven oncogenes and histone in two-word searches without restriction to cancer. However, the category for MET had to be removed to achieve a comparable correlation when searches were restricted to “cancer”. Although any explanation requires definitive studies, this may reflect under-reporting of negative results, and this study also found unexpected negative results for Met in HGF-stimulated migration studies. Also, the PubMed citations for BK (KCNMA1 encoded) probably constituted a large proportion (approximately one-third) of studies involving K+ and H+ across the categories of seven oncogenes and “histone” not restricted to cancer and most of the studies when restricted to “cancer”. The stochastic findings in this study suggest that at least some other H+-sensitive KCN DEGs are also influenced by oncogenes in H+ dynamics.

4.5. Non-KCN DEGs in Multiple E1–E70 Studies

In E1–E70, the most frequent non-KCN DEGs found were ITGA2, HMGA2, LIF, and PLAUR, with each found in six to seven studies. To support the relevance of cancer gene expression studies with unbiased detection of KCN DEGs, six DEGs, i.e., FGFR3, GDF15, IGFBP3, IGFBP7, KRT19, and PLAUR, each included in at least four of the E1–E70 studies, were also identified by others as encoding circulating proteins associated with a risk of cancer [223,224]. PLAUR, found in six of the E1–E70 studies, was co-expressed with the most frequent KCN genes that are also H+-sensitive, KCNN4 and KCNMA1 in four and two studies, respectively. KCNN4, PLAUR, and PLAU were detected together as DEGs in a study associated with three categories, RAS, TP53, and His/Chr/Epi. BK (KCNMA1) citations in PubMed correlated with these for PLAUR + PLAU and ITGA2. Co-expression of non-KCN DEGs may become relevant functionally and/or as a signature of K+ dynamics if they occur in additional studies.

4.6. Specific Roles for Redistribution of K+ in Cancer

In studies on intracellular pH-regulating systems, potassium channel inhibitors have been used with interesting results, including “slower recovery from CO2 acidification” in early frog skeletal muscle experiments [286]. Potassium channels have been recognized as playing important roles in cancer, including volume changes [287,288], but without also considering K+ exchange with H+ on fixed anions as inciting pH reversal. Movement of hydrated potassium (K+-7H2O) from cell interiors to inner-membrane regions may redistribute water via outward flux of K+ requiring separation from its loosely aggregated water molecules to achieve its passage through KCN channels. Large outward (KCNMA1/BK), intermediate outward (KCNN4/IK), and bidirectional (KCNJ10-KCNJ16/Kir4.1-Kir5.1 or Kir4.1 only) fluxes, etc., may alter distributions of both K+ and H2O. Simultaneous changes in Cl channels (CLCN genes) were detected only once (CLCN7 as a DEG) in E1–E70. Aquaporin participation seemed unlikely in E1–E70 stochastically. Na+ binds water tightly. Redistribution of water regionally within cells may promote Grotthuss conditions near inner membranes to aid pH reversal but this is unknown. Additionally, redistribution of K+ away from mitochondria within cell interiors may affect mitochondrial K+-dependent enzymes to influence malignant behavior. The results of the E1–E70 studies only revealed gene expression that crossed thresholds set by the study authors for significance. Lower gene expression levels may collectively also be biologically relevant in a network of KCN-encoded proteins mediating potassium flux but would require bioinformatic searches of supplemental data.

5. Conclusions

Migration and settling assays provided data to support and focus further investigation of potassium flux in cancer. Although pH reversal in cancer cells generated via proton export by multiple extruders has already been established, if initiating events involve potassium, then therapeutic options may widen. Multiple tumor models and study approaches have produced stochastic evidence for potassium flux via the KCN family playing an H+-related role in at least a subset of malignancies. Therapeutic suggestions regarding alterations of potassium channels carry concern regarding effects on the heart [289]. However, potassium channels are already being targeted in cancer [290]. Any potential therapeutic benefit through diet is considered since foods rich in potassium salts of metabolizable anions, such as citrate, etc., characteristic of fruits and vegetables, should deplete hydrogen ions and have an alkalizing effect [291]. The stochastic view of cancers based on unbiased studies detecting KCN genes incidentally in genomic landscapes suggests that encoded H+-sensitive KCN proteins may be downstream effectors of one or more drivers in cancer. Oncogenes may promote adaptation to adverse events and genetic damage to cells, including KCN genes, to optimize cell fitness for survival, possibly as a spectrum of ongoing changes that include the extreme outcome of becoming dedifferentiated migrating cells, via pH reversal. Also, cancer-specific structural alterations of surface proteins encoded by genetically flawed KCN genes may add to the repertoire of targets for immune attack in personalized medical treatments. Author curated, unbiased, published gene landscapes in complex cancer studies are useful for stochastically identifying candidates in the large KCN gene family that may encode downstream effectors of oncogenes that may initiate pH reversal. Also, tumor models for KCN genes and encoded proteins are identified, where significant changes in KCN gene expression have been seen in many types of malignancy.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/biom15081177/s1, Table S1: FullTblB1DEGalphaAZ.

Funding

The cell settling and cell migration research was funded by the Walter L. Copeland Fund for Cranial Research, The Pittsburgh Foundation, Pittsburgh, PA, D2003-0305 and D2006-0379; The Nick Eric Wichman Foundation, Ellicott City, MD, No Number; The Beez Foundation, Somerset, NJ, No Number; and the American Heart Association, Ohio Valley Affiliate, AHA 9960398V.

Institutional Review Board Statement

A172 (RRID:CVCL_0131), C6 (TKG Cat# TKG 0242, RRID:CVCL_0194), F98 (ATCC Cat# CRL-2397, RRID:CVCL_3510), LN18 (RRID:CVCL_0392), LN229 (RRID:CVCL_0393), T98G (RRID:CVCL_0556), U87 (RRID:CVCL_0022), and U373 (RRID:CVCL_2219) glioblastoma/astrocytoma cell lines (human (A172, LN18, LN229, U87, U373) and rat (F98 and C6)) were obtained from the American Type Culture Collection, Manassas, VA, USA RRID:SCR_021346. No cell lines were generated from live animals or human patients for this study by the author.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data generated in this study will be available upon request from the author. Please note that data from REMBRANDT per Adobe Flash (not available after 31 December 2020) can be seen in Figure 6, except for a few instances described, which can be made available on request. KCN gene data obtained from the published manuscript by Patil, V. et al. [22] follows the terms of their statement “an open access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited”. Of the papers used in the E1–E70 studies, 34%, 31%, 13%, and 6% were found in Cancer Research, Cancer Cell, Nature Communications, and Journal of Molecular Diagnostics, respectively. Others were from 11 journals, one paper each. Papers for E1–E70 are cited in Section 3.3.6.

Acknowledgments

Ian F. Pollack, University of Pittsburgh, Children’s Hospital of Pittsburgh, provided laboratory space, supplies, technicians, and glioblastoma cell lines. The Departments of Pathology and Neurological Surgery, University of Pittsburgh, provided institutional support for the cell settling and migration experimental work. The 1301 authors of the 70 papers (E1–E70) used for gathering stochastic information on the KCN gene family while at Kent State, Kent, OH, are thanked for outstanding contributions to the medical science literature. I thank the Brain Health Research Institute at Kent State University, Kent, Ohio, as a current member, for technical advice and ongoing academic research activities. The author performed the cell settling and migration assays, and immunoblot at a previous institution: The University of Pittsburgh, Pittsburgh, PA. The author performed the analysis of REMBRANDT, collection and analysis of 70 studies detecting KCN DEGs, etc., and all other aspects of the study (text, figures, etc.) while associated with Kent State University, Kent, OH. Consultation for statistical advice, technical assistance with the R v4.3.1 software, and the use of EndNote was obtained and appreciated but acknowledgement for advice and assistance with R and Endnote was not desired.

Conflicts of Interest

The author declares no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript.
ACLYATP citrate lyase
Arg1Arginase 1
ASAnti-seizure
BKBig K (encoded by KCNMA1)
BSABovine serum albumin
CDCluster Differentiation
CLCNChloride Channels gene family, CLCN1,2,3, etc.
COLCollagen
CSMCancer-specific mutation
CxConnexin
DEGDifferentially expressed gene
EExpression
EAATExcitatory Amino Acid Transporter
EAG1Ether-A-Go-Go-K+ channel protein 1
ELK1,2EAG-like K+ channel proteins 1,2
EMTEpithelial mesenchymal transition
ERG1,3EAG-related gene K+ channel proteins 1,3
FBSFetal bovine serum
GABAGamma-aminobutyric acid
GBMGlioblastoma
G-DOCGeorgetown Database of Cancer
Girk3,4G protein-activated inward rectif-ier K+ channel proteins, 3,4
GPRG-protein-coupled receptor
G-proteinsGuanosine-triphosphate-binding proteins
GWATERWGrotthuss water wire
HGFHepatocyte growth factor
His/Chr/EpiHistone/chromatin/epigenetic
HKGHousekeeping gene
IKIntermediate K (encoded by KCNN4)
IndelsInsertions and deletions
ITAG2Integrin 2
ITGIntegrin
K2PTandem 2 pore domain KCN subfamily proteins
KcaCalcium-stimulated KCN subfamily proteins
Kchip1,3Kv channel interacting proteins 1,3
KCNASubgroup of KCN genes, KCNA1,2,3, etc.
KCNABSubgroup of KCN genes, KCNAB1,2,3
KCNBSubgroup of KCN genes, KCNB1,2
KCNCSubgroup of KCN genes, KCNC1,2,3, etc.
KCNDSubgroup of KCN genes, KCND1,2,3
KCNESubgroup of KCN genes, KCNE1,2,3, etc.
KCNGSubgroup of KCN genes, KCNG1,2,3, etc.
KCNHSubgroup of KCN genes, KCNH1,2,3, etc.
KCNIPSubgroup of KCN genes, KCNIP1,2,3, etc.
KCNJSubgroup of KCN genes, KCNJ1,2,3, etc.
KCNKSubgroup of KCN genes, KCNK1,2,3, etc.
KCNMSubgroup of KCN genes, KCNM,A1,B1,B2, etc.
KCNNSubgroup of KCN genes, KCNN1,2,3, etc.
KCNQSubgroup of KCN genes, KCNQ1,2,3, etc.
KCNSSubgroup of KCN genes, KCNS1,2,3, etc.
KCNTSubgroup of KCN genes, KCNT1,2
KirInwardly directing KCN subfamily proteins
KvVoltage-dependent KCN subfamily proteins
LDHLactate dehydrogenase
LIFLeukemia Inhibitory Factor
MCTMonocarboxylate transporter
MEMMinimal Essential Media
MesMesenchymal
MiRPMinK (minimal K+channel Protein-related peptide)
Na/K ATPaseSodium/potassium ATPase
NOXNADPH oxidases
OligoOligodendroglioma
OMGOligodendroglioma minus glioblastoma
OMIMOnline Mendelian Inheritance in Man
PBSPhosphate-buffered saline
pHeExtracellular pH
pHiIntracellular pH
PIP2Phosphatidyl inositol 4,5 bisphosphate
RPearson product-moment correlation
REMBRANDTRepository of Molecular Brain Neoplasia Data
RGSRegulator of G protein signaling
ROSReactive oxygen species
RRIDResearch Resource Identifier
SERPINSerine proteinase inhibitors
SNVSingle-nucleotide variant
SKSmall K (encoded by some KCNN genes)
SlackSequence like a calcium-activated K+ channel protein
SLCSolute Carrier
SlickSequence like an intermediate conductance K+ channel protein
SOXSry-related HMG BOX
TasK1,2,5Twik-related acid-sensitive K+ channel proteins 1,2,5
Thik2Tandem pore domain halothane-inhibited Inhibited K+ channel protein 2
TMEMTransmembrane
TNFRSFTumor Necrosis Factor Receptor Super Family
Trek1Twik-related K+ channel protein 1
TwiK1,2Tandem pore domain in weak inward rectifying K+ channel proteins 1,2
uPAUrokinase-type plasminogen activator

Appendix A

Figure A1. Macroscopic view of settled T98G glioblastoma cells on a Neuro Probe microscopic slide. One of 8 round 7 mm diameter areas that had been the bottom of a settling chamber well. Some cell clumping is noted. Diff Quik stain.
Figure A1. Macroscopic view of settled T98G glioblastoma cells on a Neuro Probe microscopic slide. One of 8 round 7 mm diameter areas that had been the bottom of a settling chamber well. Some cell clumping is noted. Diff Quik stain.
Biomolecules 15 01177 g0a1
Figure A2. All three migration assays for each cell line (C6, F98, T98G, and LN18) in studies of HGF and FBS chemoattraction. (A) Three migration assays for C6 and F98 cell line. (B) Three migration assays for T98G and LN18 cell line. One assay per cell line is shown in Figure 2A. Portions of assays that were digitized and analyzed are shown in Figure 2B. Text in Section 3.2. and the Figure 2 legend provides additional information. Each circle of stained migrated cells is 3 mm in diameter.
Figure A2. All three migration assays for each cell line (C6, F98, T98G, and LN18) in studies of HGF and FBS chemoattraction. (A) Three migration assays for C6 and F98 cell line. (B) Three migration assays for T98G and LN18 cell line. One assay per cell line is shown in Figure 2A. Portions of assays that were digitized and analyzed are shown in Figure 2B. Text in Section 3.2. and the Figure 2 legend provides additional information. Each circle of stained migrated cells is 3 mm in diameter.
Biomolecules 15 01177 g0a2
Figure A3. All three migration assays for each cell line (U87, LN229, A172, and U373) in studies of HGF and FBS chemoattraction. (A) Three migration assays for U87 and LN229 cell line. (B) Three migration assays for A172 and U373 cell line. One assay per cell line is shown in Figure 2A. Portions of assays that were digitized and analyzed are shown in Figure 2B. Text in Section 3.2. and the Figure 2 legend provides additional information. Each circle of stained migrated cells is 3 mm in diameter.
Figure A3. All three migration assays for each cell line (U87, LN229, A172, and U373) in studies of HGF and FBS chemoattraction. (A) Three migration assays for U87 and LN229 cell line. (B) Three migration assays for A172 and U373 cell line. One assay per cell line is shown in Figure 2A. Portions of assays that were digitized and analyzed are shown in Figure 2B. Text in Section 3.2. and the Figure 2 legend provides additional information. Each circle of stained migrated cells is 3 mm in diameter.
Biomolecules 15 01177 g0a3
Figure A4. Additional images for Figure 2. (A) Gel image for the immunoblot in Figure 2C, showing equal loading of protein in each lane for all 8 cell lines. (B) Scanned image of the U87 migration assay shown in Figure 2D. Dulbecco’s PBS and MEM were tested as media in the lower wells for migration without and with FBS (2 concentrations) present. The image of the filter removed (white horizontal line) an unrelated portion of the assay that tested a different batch of FBS for suitability in future migration assays. (C) Scanned image of the U87 migration assay shown in Figure 2E. Chemoattraction of HGF and 2 concentrations of FBS and chemokinesis of HGF, with and without 2 concentrations of FBS, were analyzed in U87 cells. Text in Section 3.2. and the Figure 2 legend provides additional information. Each circle of stained, migrated cells is 3 mm in diameter.
Figure A4. Additional images for Figure 2. (A) Gel image for the immunoblot in Figure 2C, showing equal loading of protein in each lane for all 8 cell lines. (B) Scanned image of the U87 migration assay shown in Figure 2D. Dulbecco’s PBS and MEM were tested as media in the lower wells for migration without and with FBS (2 concentrations) present. The image of the filter removed (white horizontal line) an unrelated portion of the assay that tested a different batch of FBS for suitability in future migration assays. (C) Scanned image of the U87 migration assay shown in Figure 2E. Chemoattraction of HGF and 2 concentrations of FBS and chemokinesis of HGF, with and without 2 concentrations of FBS, were analyzed in U87 cells. Text in Section 3.2. and the Figure 2 legend provides additional information. Each circle of stained, migrated cells is 3 mm in diameter.
Biomolecules 15 01177 g0a4

Appendix B

Table A1. A sample of alphabetized differentially expressed genes (DEGs) in E1–E70 studies. See Supplementary Materials for the entire DEG list; Table S1: FullTblB1DEGalphaAZ can be downloaded as a Word file at Table S1.
Table A1. A sample of alphabetized differentially expressed genes (DEGs) in E1–E70 studies. See Supplementary Materials for the entire DEG list; Table S1: FullTblB1DEGalphaAZ can be downloaded as a Word file at Table S1.
DEG NameStudy Number
A2ME21
AAK1E65
AATKE54
AB113APE1
ABCA1E50
ABCA12E54
ABCB10E48
ABCB1AE69
ABCC3E70
ABCD1E31
See full list…See full list…
Table A2. KCN genes in E1–E70 studies that encode H+-sensitive proteins.
Table A2. KCN genes in E1–E70 studies that encode H+-sensitive proteins.
KCN GeneProteinReference
KCNA1Kv1.1[292]
KCNA3Kv1.3[293,294]
KCNA5Kv1.5[295,296,297,298,299,300,301]
KCNAB2Kvbeta2[302,303]
KCND3Kv4.3[304,305,306]
KCNE2MiRP1[87,307,308,309,310,311]
KCNE3MiRP2, R83H variant[252]
KCNH1Kv10.1, EAG1[253,312]
KCNH2Kv11.1, HERG1[253,313,314,315,316,317,318,319,320,321,322,323,324,325]
KCNH3Kv12.2, ELK2[253]
KCNH8Kv12.1, ELK3[253,326]
KCNJ5Kir3.4, GIRK4[254,255,327]
KCNJ8Kir6.1, KATP-1[327,328,329,330,331,332,333]
KCNJ10Kir4.1, BIRK10[334,335,336]
KCNJ10/KCNJ16Kir4.1/Kir5.1heterodimer[72,326,337,338,339,340,341,342,343]
KCNJ13Kir7.1[72,327,344]
KCNK1K2P1, TWIK1[343,345,346,347,348,349,350]
KCNK2K2P2, TREK1[343,348,351,352,353,354,355,356,357]
KCNK3K2P3, TASK1[343,346,348,355,358,359,360,361]
KCNK5K2P5, TASK2[236,343,346,348,362,363,364,365,366,367,368,369]
KCNK6K2P6, TWIK2[346,370,371]
KCNK15K2P15, TASK5[372]
KCNMA1KCa1.1, BK, Slo1, maxiK[96,373,374,375,376,377,378,379,380,381,382,383,384,385,386,387,388]
KCNMB4BKbeta4[389,390,391,392,393]
KCNN2KCa2.2, SK2[394]
KCNN3KCa2.3, SK3[394]
KCNN4KCa3.1, SK4, Gardos[395,396,397,398]
KCNQ1Kv7.1, KvLQT1[85,236,307,308,309,311,399,400,401,402,403,404,405]
KCNQ2Kv7.2[402] (described data) [406,407]
KCNQ3Kv7.3[406]
KCNQ5Kv7.5[408]
KCNT1KCa4.1, Slo2.2, KNa1.1[409,410,411,412]
KCNU1KCa5.1, Slo3, KCNMA3[389,412,413,414,415,416,417,418]
Figure A5. Types of malignancies studied in E1–E70 and informative protein abbreviations for proteins in the KCN family. (A) Specific studies for 29 malignancies with their detected KCN DEGs. Red underlining indicates KCN DEGs that encode H+-sensitive proteins. Boxes that correspond with studies of glioblastoma, breast cancer, and lung cancer are color-coded: yellow, pink, and blue, respectively. These graphics are also used in the histograms shown in Figure 3 and Figure 4. (B) Some common protein abbreviations for KCN family members in Figure 3, indicating functional information not explained in the text.
Figure A5. Types of malignancies studied in E1–E70 and informative protein abbreviations for proteins in the KCN family. (A) Specific studies for 29 malignancies with their detected KCN DEGs. Red underlining indicates KCN DEGs that encode H+-sensitive proteins. Boxes that correspond with studies of glioblastoma, breast cancer, and lung cancer are color-coded: yellow, pink, and blue, respectively. These graphics are also used in the histograms shown in Figure 3 and Figure 4. (B) Some common protein abbreviations for KCN family members in Figure 3, indicating functional information not explained in the text.
Biomolecules 15 01177 g0a5
Text A1. Potential interactions of KCN proteins with H+-sensitive G protein-coupled receptors. Multiple KCN proteins have interactions with four G protein-coupled receptors encoded by GPR4, 65, 68, and 151, that are among six reported to be H+-sensitive [419,420,421,422,423,424,425]. H+-sensitive proteins encoded by GPR genes are found in many types of tumors [423,424,426,427]. Functional complementation between proteins encoded by GPR4 and KCNK5/TASK2 [427,428,429] and related expression profiles of proteins encoded by GPR31, GPR151, KCNK3/TASK1, and KCNK9/TASK3 have been found in tumors [425]. STRING interactions (Gene Cards, 17 October 2024) of proteins encoded by GPR4, 65, 68, and 151 as a group included proteins encoded by nine KCNs: KCNJ8, K1, K2, K4, K5, K10, K16, K17, and K18. Among these, 8 of the 70 studies (E3, E8, E22, E24, E31, E35, E61, E68) detected KCNJ8, K1, K2, and K5 as DEGs. Also, although GPR39’s encoded protein is not reported to be H+-sensitive, it interacts with the four H+-sensitive proteins encoded by GPR4, 65, 68, and 132. GPR39 is a DEG found in 2 of the 70 studies, E14 [193] and E60 [176], that also detected KCNQ5 and KCNMA1, respectively. The ligand of the protein encoded by GPR39 is encoded by GPNMB [430], and GPNMB was a DEG in three studies: E26 [212], E29 [157], and E49 [166]. GPR35’s protein (also not H+-sensitive) interacts with GPR4’s encoded H+-sensitive receptor protein. GPR35 is a DEG that was found in two studies that found KCNG2 and KCNE3, E20 [210] and E42 [217], respectively. The ligand for GPR35’s receptor protein is kynurenic acid [431]. GPR84’s protein (also not H+-sensitive) interacts with GPR65’s encoded H+-sensitive receptor protein. GPR84 is a DEG that was found in two studies, with KCNK6 found in E11 [173] and KCNK12 and KCNU1 in E23 [201]. The ligands for GPR84’s receptor protein are medium-chain fatty acids [432]. Thirteen of the 14 studies (out of 70) with potential indirect GPR protein sensitivity to H+ also detected KCN DEGs that encode H+-sensitive proteins. Potentially, KCN DEG-encoded proteins in 13 (18.6%) of the E1–E70 studies have indirect involvement with H+-sensitive G protein-coupled receptor proteins via STRING interactions (Gene Cards, 17 October 2024) or as described above according to the cited literature. They may function collaboratively with GPR-encoded H+-sensitive proteins in tumors.

References

  1. Burger, P.C.; Scheithauer, B.W. Tumors of the Neuroglia and Choroid Plexus. In Tumors of the Central Nervous System; American Registry of Pathology: Washington, DC, USA, 2007; pp. 55–77. [Google Scholar]
  2. Ge, L.; Hoa, N.T.; Wilson, Z.; Arismendi-Morillo, G.; Kong, X.T.; Tajhya, R.B.; Beeton, C.; Jadus, M.R. Big Potassium (BK) ion channels in biology, disease and possible targets for cancer immunotherapy. Int. Immunopharmacol. 2014, 22, 427–443. [Google Scholar] [CrossRef]
  3. Freshney, R.I. Defined media and supplements: Table 8.6. Constituents of serum. In Culture of Animal Cells; Capes-Davis, A., Gregory, C., Przyborski, S., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2016; p. 141. [Google Scholar]
  4. Kros, J.M. Grading of gliomas: The road from eminence to evidence. J. Neuropathol. Exp. Neurol. 2011, 70, 101–109. [Google Scholar] [CrossRef]
  5. Puchalski, R.B.; Shah, N.; Miller, J.; Dalley, R.; Nomura, S.R.; Yoon, J.G.; Smith, K.A.; Lankerovich, M.; Bertagnolli, D.; Bickley, K.; et al. An anatomic transcriptional atlas of human glioblastoma. Science 2018, 360, 660–663. [Google Scholar] [CrossRef]
  6. Brat, D.J.; Castellano-Sanchez, A.A.; Hunter, S.B.; Pecot, M.; Cohen, C.; Hammond, E.H.; Devi, S.N.; Kaur, B.; Van Meir, E.G. Pseudopalisades in glioblastoma are hypoxic, express extracellular matrix proteases, and are formed by an actively migrating cell population. Cancer Res. 2004, 64, 920–927. [Google Scholar] [CrossRef]
  7. Rong, Y.; Durden, D.L.; Van Meir, E.G.; Brat, D.J. ‘Pseudopalisading’ necrosis in glioblastoma: A familiar morphologic feature that links vascular pathology, hypoxia, and angiogenesis. J. Neuropathol. Exp. Neurol. 2006, 65, 529–539. [Google Scholar] [CrossRef]
  8. Chesler, M.; Kraig, R.P. Intracellular pH of astrocytes increases rapidly with cortical stimulation. Am. J. Physiol. 1987, 253, R666–R670. [Google Scholar] [CrossRef]
  9. Chesler, M.; Kraig, R.P. Intracellular pH transients of mammalian astrocytes. J. Neurosci. 1989, 9, 2011–2019. [Google Scholar] [CrossRef] [PubMed]
  10. Brookes, N. Intracellular pH as a regulatory signal in astrocyte metabolism. Glia 1997, 21, 64–73. [Google Scholar] [CrossRef]
  11. Theparambil, S.M.; Weber, T.; Schmalzle, J.; Ruminot, I.; Deitmer, J.W. Proton Fall or Bicarbonate Rise: Glycolytic Rate in Mouse Astrocytes Is Paved By Intracellular Alkalinization. J. Biol. Chem. 2016, 291, 19108–19117. [Google Scholar] [CrossRef] [PubMed]
  12. Vaupel, P.; Kallinowski, F.; Okunieff, P. Blood flow, oxygen and nutrient supply, and metabolic microenvironment of human tumors: A review. Cancer Res. 1989, 49, 6449–6465. [Google Scholar]
  13. Boron, W.F. Regulation of intracellular pH. Adv. Physiol. Educ. 2004, 28, 160–179. [Google Scholar] [CrossRef]
  14. Andersen, A.P.; Moreira, J.M.; Pedersen, S.F. Interactions of ion transporters and channels with cancer cell metabolism and the tumour microenvironment. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130098. [Google Scholar] [CrossRef]
  15. Spugnini, E.P.; Sonveaux, P.; Stock, C.; Perez-Sayans, M.; De Milito, A.; Avnet, S.; Garcìa, A.G.; Harguindey, S.; Fais, S. Proton channels and exchangers in cancer. Biochim. Biophys. Acta 2015, 1848, 2715–2726. [Google Scholar] [CrossRef] [PubMed]
  16. Swietach, P.; Vaughan-Jones, R.D.; Harris, A.L.; Hulikova, A. The chemistry, physiology and pathology of pH in cancer. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130099. [Google Scholar] [CrossRef] [PubMed]
  17. Swietach, P.; Boedtkjer, E.; Pedersen, S.F. How protons pave the way to aggressive cancers. Nat. Rev. Cancer 2023, 23, 825–841. [Google Scholar] [CrossRef]
  18. McIntyre, A.; Hulikova, A.; Ledaki, I.; Snell, C.; Singleton, D.; Steers, G.; Seden, P.; Jones, D.; Bridges, E.; Wigfield, S.; et al. Disrupting Hypoxia-Induced Bicarbonate Transport Acidifies Tumor Cells and Suppresses Tumor Growth. Cancer Res. 2016, 76, 3744–3755. [Google Scholar] [CrossRef] [PubMed]
  19. White, K.A.; Grillo-Hill, B.K.; Barber, D.L. Cancer cell behaviors mediated by dysregulated pH dynamics at a glance. J. Cell Sci. 2017, 130, 663–669. [Google Scholar] [CrossRef]
  20. Bhuvaneshwar, K.; Belouali, A.; Singh, V.; Johnson, R.M.; Song, L.; Alaoui, A.; Harris, M.A.; Clarke, R.; Weiner, L.M.; Gusev, Y.; et al. G-DOC Plus—An integrative bioinformatics platform for precision medicine. BMC Bioinform. 2016, 17, 193. [Google Scholar] [CrossRef]
  21. Madhavan, S.; Gusev, Y.; Harris, M.; Tanenbaum, D.M.; Gauba, R.; Bhuvaneshwar, K.; Shinohara, A.; Rosso, K.; Carabet, L.A.; Song, L.; et al. G-DOC: A systems medicine platform for personalized oncology. Neoplasia 2011, 13, 771–783. [Google Scholar] [CrossRef]
  22. Patil, V.; Pal, J.; Somasundaram, K. Elucidating the cancer-specific genetic alteration spectrum of glioblastoma derived cell lines from whole exome and RNA sequencing. Oncotarget 2015, 6, 43452–43471. [Google Scholar] [CrossRef]
  23. Deplazes, E.; Poger, D.; Cornell, B.; Cranfield, C.G. The effect of H(3)O(+) on the membrane morphology and hydrogen bonding of a phospholipid bilayer. Biophys. Rev. 2018, 10, 1371–1376. [Google Scholar] [CrossRef] [PubMed]
  24. Eigen, M.; De Maeyer, L. Self-dissociation and protonic charge transport in water and ice. Proc. R. Soc. A 1958, 247 A, 505–533. [Google Scholar]
  25. Johnson, D.E.; Casey, J.R. Cytosolic H+ microdomain developed around AE1 during AE1-mediated Cl-/HCO3- exchange. J. Physiol. 2011, 589, 1551–1569. [Google Scholar] [CrossRef]
  26. Vaughan-Jones, R.D.; Peercy, B.E.; Keener, J.P.; Spitzer, K.W. Intrinsic H(+) ion mobility in the rabbit ventricular myocyte. J. Physiol. 2002, 541, 139–158. [Google Scholar] [CrossRef]
  27. Xie, F.; Tikhonov, D.S.; Schnell, M. Electric nuclear quadrupole coupling reveals dissociation of HCl with a few water molecules. Science 2024, 384, 1435–1440. [Google Scholar] [CrossRef] [PubMed]
  28. Grotthuss, C.J.T. Memoire sur la decomposition de l’eau et des corps qu’elle tient en dissolution a l’aide de l’electricite galvanique. Ann. Chim. 1806, 58, 54–74. [Google Scholar]
  29. Adelroth, P. Preface: Special issue on proton transfer in biological systems. Biochem. Biophys. Acta 2006, 1757, 867–870. [Google Scholar]
  30. Cukierman, S. Et tu, Grotthuss! And other unfinished stories. Biochim. Biophys. Acta 2006, 1757, 876–885. [Google Scholar] [CrossRef]
  31. Gutman, M.; Nachliel, E.; Friedman, R. The mechanism of proton transfer between adjacent sites on the molecular surface. Biochim. Biophys. Acta 2006, 1757, 931–941. [Google Scholar] [CrossRef]
  32. Mulkidjanian, A.Y.; Heberle, J.; Cherepanov, D.A. Protons @ interfaces: Implications for biological energy conversion. Biochim. Biophys. Acta 2006, 1757, 913–930. [Google Scholar] [CrossRef]
  33. Ben-Amotz, D. Electric buzz in a glass of pure water. Science 2022, 376, 800–801. [Google Scholar] [CrossRef] [PubMed]
  34. Kariev, A.M.; Green, M.E. Water, Protons, and the Gating of Voltage-Gated Potassium Channels. Membranes 2024, 14, 37. [Google Scholar] [CrossRef]
  35. Horowitz, S.B.; Paine, P.L. Reference phase analysis of free and bound intracellular solutes. II. Isothermal and isotopic studies of cytoplasmic sodium, potassium, and water. Biophys. J. 1979, 25, 45–62. [Google Scholar] [CrossRef] [PubMed]
  36. Schornack, P.A.; Gillies, R.J. Contributions of cell metabolism and H+ diffusion to the acidic pH of tumors. Neoplasia 2003, 5, 135–145. [Google Scholar] [CrossRef]
  37. Schwab, A.; Wojnowski, L.; Gabriel, K.; Oberleithner, H. Oscillating activity of a Ca(2+)-sensitive K+ channel. A prerequisite for migration of transformed Madin-Darby canine kidney focus cells. J. Clin. Invest. 1994, 93, 1631–1636. [Google Scholar] [CrossRef] [PubMed]
  38. Schwab, A.; Nechyporuk-Zloy, V.; Fabian, A.; Stock, C. Cells move when ions and water flow. Pflugers Arch. 2007, 453, 421–432. [Google Scholar] [CrossRef]
  39. Horio, Y. Potassium channels of glial cells: Distribution and function. Jpn. J. Pharmacol. 2001, 87, 1–6. [Google Scholar] [CrossRef]
  40. Sontheimer, H. An unexpected role for ion channels in brain tumor metastasis. Exp. Biol. Med. 2008, 233, 779–791. [Google Scholar] [CrossRef]
  41. Schwab, A.; Stock, C. Ion channels and transporters in tumour cell migration and invasion. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130102. [Google Scholar] [CrossRef]
  42. Catacuzzeno, L.; Sforna, L.; Esposito, V.; Limatola, C.; Franciolini, F. Ion Channels in Glioma Malignancy. Rev. Physiol. Biochem. Pharmacol. 2021, 181, 223–267. [Google Scholar]
  43. Beckner, M.E.; Chen, X.; An, J.; Day, B.W.; Pollack, I.F. Proteomic characterization of harvested pseudopodia with differential gel electrophoresis and specific antibodies. Lab. Invest. 2005, 85, 316–327. [Google Scholar] [CrossRef]
  44. Tjaderhane, L.; Larjava, H.; Sorsa, T.; Uitto, V.J.; Larmas, M.; Salo, T. The activation and function of host matrix metalloproteinases in dentin matrix breakdown in caries lesions. J. Dent. Res. 1998, 77, 1622–1629. [Google Scholar] [CrossRef]
  45. Fasciglione, G.F.; Marini, S.; D’Alessio, S.; Politi, V.; Coletta, M. pH- and temperature-dependence of functional modulation in metalloproteinases: A comparison between neutrophil collagenase and gelatinases A and B. Biophys. J. 2000, 79, 2138–2149. [Google Scholar] [CrossRef]
  46. Kato, Y.; Lambert, C.A.; Colige, A.C.; Mineur, P.; Noel, A.; Frankenne, F.; Foidart, J.M.; Baba, M.; Hata, R.I.; Miyazaki, K.; et al. Acidic extracellular pH induces matrix metalloproteinase-9 expression in mouse metastatic melanoma cells through the phospholipase D-mitogen-activated protein kinase signaling. J. Biol. Chem. 2005, 280, 10938–10944. [Google Scholar] [CrossRef]
  47. Gatenby, R.A.; Gawlinski, E.T.; Gmitro, A.F.; Kaylor, B.; Gillies, R.J. Acid-mediated tumor invasion: A multidisciplinary study. Cancer Res. 2006, 66, 5216–5223. [Google Scholar] [CrossRef]
  48. Gioia, M.; Fasciglione, G.F.; Monaco, S.; Iundusi, R.; Sbardella, D.; Marini, S.; Tarantino, U.; Coletta, M. pH dependence of the enzymatic processing of collagen I by MMP-1 (fibroblast collagenase), MMP-2 (gelatinase A), and MMP-14 ectodomain. J. Biol. Inorg. Chem. 2010, 15, 1219–1232. [Google Scholar] [CrossRef] [PubMed]
  49. Pillai, S.R.; Damaghi, M.; Marunaka, Y.; Spugnini, E.P.; Fais, S.; Gillies, R.J. Causes, consequences, and therapy of tumors acidosis. Cancer Metastasis Rev. 2019, 38, 205–222. [Google Scholar] [CrossRef] [PubMed]
  50. Rohani, N.; Hao, L.; Alexis, M.S.; Joughin, B.A.; Krismer, K.; Moufarrej, M.N.; Soltis, A.R.; Lauffenburger, D.A.; Yaffe, M.B.; Burge, C.B.; et al. Acidification of Tumor at Stromal Boundaries Drives Transcriptome Alterations Associated with Aggressive Phenotypes. Cancer Res. 2019, 79, 1952–1966. [Google Scholar] [CrossRef] [PubMed]
  51. Sun, H.; Zhou, Y.; Skaro, M.F.; Wu, Y.; Qu, Z.; Mao, F.; Suwen Zhao, S.; Xu, Y. Metabolic Reprogramming in Cancer Is Induced to Increase Proton Production. Cancer Res. 2020, 80, 1143–1155. [Google Scholar] [CrossRef]
  52. Buhl, M.; Hutson, T.; Missio, A.; Walton, J.C. Sulfur and Phosphorus Oxyacid Radicals. J. Phys. Chem. A 2022, 126, 760–771. [Google Scholar] [CrossRef]
  53. Walton, J.C. Radical-Enhanced Acidity: Why Bicarbonate, Carboxyl, Hydroperoxyl, and Related Radicals Are So Acidic. J. Phys. Chem. A 2017, 121, 7761–7767. [Google Scholar] [CrossRef]
  54. Walton, J.C. Microhydration and the Enhanced Acidity of Free Radicals. Molecules 2018, 23, 423. [Google Scholar] [CrossRef]
  55. Segal, M.S.; Beem, E. Effect of pH, ionic charge, and osmolality on cytochrome c-mediated caspase-3 activity. Am. J. Physiol. Cell Physiol. 2001, 281, C1196–C1204. [Google Scholar] [CrossRef]
  56. Duffy, H.S.; Ashton, A.W.; O’Donnell, P.; Coombs, W.; Taffet, S.M.; Delmar, M.; Spray, D.C. Regulation of connexin43 protein complexes by intracellular acidification. Circ. Res. 2004, 94, 215–222. [Google Scholar] [CrossRef]
  57. Hirst-Jensen, B.J.; Sahoo, P.; Kieken, F.; Delmar, M.; Sorgen, P.L. Characterization of the pH-dependent interaction between the gap junction protein connexin43 carboxyl terminus and cytoplasmic loop domains. J. Biol. Chem. 2007, 282, 5801–5813. [Google Scholar] [CrossRef]
  58. Palacios-Prado, N.; Briggs, S.W.; Skeberdis, V.A.; Pranevicius, M.; Bennett, M.V.; Bukauskas, F.F. pH-dependent modulation of voltage gating in connexin45 homotypic and connexin45/connexin43 heterotypic gap junctions. Proc. Natl. Acad. Sci. USA 2010, 107, 9897–9902. [Google Scholar] [CrossRef] [PubMed]
  59. Monterisi, S.; Michl, J.; Hulikova, A.; Koth, J.; Bridges, E.M.; Hill, A.E.; Abdullayeva, G.; Bodmer, W.F.; Swietach, P. Solute exchange through gap junctions lessens the adverse effects of inactivating mutations in metabolite-handling genes. Elife 2022, 11, e78425. [Google Scholar] [CrossRef]
  60. Ranganathan, N.S.; Srere, P.A.; Linn, T.C. Comparison of phospho- and dephospho-ATP citrate lyase. Arch. Biochem. Biophys. 1980, 204, 52–58. [Google Scholar] [CrossRef] [PubMed]
  61. Wu, T.; Jun, S.; Choi, E.J.; Sun, J.; Yang, E.B.; Lee, H.S.; Kim, S.Y.; Fahmi, N.A.; Jiang, O.; Zhang, W.; et al. 53BP1-ACLY-SLBP-coordinated activation of replication-dependent histone biogenesis maintains genomic integrity. Nucleic Acids Res. 2022, 50, 1465–1483. [Google Scholar] [CrossRef] [PubMed]
  62. Beckner, M.E.; Fellows-Mayle, W.; Zhang, Z.; Agostino, N.R.; Kant, J.A.; Day, B.W.; Pollack, I.F. Identification of ATP citrate lyase as a positive regulator of glycolytic function in glioblastomas. Int. J. Cancer 2010, 126, 2282–2295. [Google Scholar] [CrossRef]
  63. Wellen, K.E.; Hatzivassiliou, G.; Sachdeva, U.M.; Bui, T.V.; Cross, J.R.; Thompson, C.B. ATP-citrate lyase links cellular metabolism to histone acetylation. Science 2009, 324, 1076–1080. [Google Scholar] [CrossRef] [PubMed]
  64. Zheng, W.; Tasselli, L.; Li, T.-M.; Chua, K.F. Mammalian SIRT6 represses invasive cancer cell phenotypes through ATP citrate lyase (ACLY)-dependent histone acetylation. Genes 2021, 12, 1460. [Google Scholar] [CrossRef] [PubMed]
  65. Scott, J.G.; Basanta, D.; Chinnaiyan, P.; Canoll, P.; Swanson, K.R.; Anderson, A.R. Production of 2-hydroxyglutarate by isocitrate dehydrogenase 1-mutated gliomas: An evolutionary alternative to the Warburg shift? Neuro Oncol 2011, 13, 1262–1264. [Google Scholar] [CrossRef]
  66. Girault, A.; Brochiero, E. Evidence of K+ channel function in epithelial cell migration, proliferation, and repair. Am. J. Physiol. Cell Physiol. 2014, 306, C307–C319. [Google Scholar] [CrossRef]
  67. Boyle, Y.; Johns, T.G.; Fletcher, E.V. Potassium Ion Channels in Malignant Central Nervous System Cancers. Cancers 2022, 14, 4767. [Google Scholar] [CrossRef] [PubMed]
  68. Cohen, N.A.; Sha, Q.; Makhina, E.N.; Lopatin, A.N.; Linder, M.E.; Snyder, S.H.; Nichols, C.G. Inhibition of an inward rectifier potassium channel (Kir2.3) by G-protein betagamma subunits. J. Biol. Chem. 1996, 271, 32301–32305. [Google Scholar] [CrossRef]
  69. Ruppersberg, J.P. Intracellular regulation of inward rectifier K+ channels. Pflugers Arch. 2000, 441, 1–11. [Google Scholar] [CrossRef]
  70. Niemeyer, M.I.; Stutzin, A.; Sepulveda, F.V. A voltage-independent K+ conductance activated by cell swelling in Ehrlich cells is modulated by a G-protein-mediated process. Biochim. Biophys. Acta 2002, 1562, 1–5. [Google Scholar] [CrossRef]
  71. Bedard, K.; Krause, K.H. The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology. Physiol. Rev. 2007, 87, 245–313. [Google Scholar] [CrossRef]
  72. Hibino, H.; Inanobe, A.; Furutani, K.; Murakami, S.; Findlay, I.; Kurachi, Y. Inwardly rectifying potassium channels: Their structure, function, and physiological roles. Physiol. Rev. 2010, 90, 291–366. [Google Scholar] [CrossRef]
  73. Sahoo, N.; Hoshi, T.; Heinemann, S.H. Oxidative modulation of voltage-gated potassium channels. Antioxid. Redox Signal 2014, 21, 933–952. [Google Scholar] [CrossRef]
  74. Rivera-Aponte, D.E.; Mendez-Gonzalez, M.P.; Rivera-Pagan, A.F.; Kucheryavykh, Y.V.; Kucheryavykh, L.Y.; Skatchkov, S.N.; Eaton, M.J. Hyperglycemia reduces functional expression of astrocytic Kir4.1 channels and glial glutamate uptake. Neuroscience 2015, 310, 216–223. [Google Scholar] [CrossRef]
  75. Wang, J.; Li, Y.; Jiang, C. MiR-133b contributes to arsenic-induced apoptosis in U251 glioma cells by targeting the hERG channel. J. Mol. Neurosci. 2015, 55, 985–994. [Google Scholar] [CrossRef]
  76. Pegg, A.E. Functions of Polyamines in Mammals. J. Biol. Chem. 2016, 291, 14904–14912. [Google Scholar] [CrossRef]
  77. Cabanos, C.; Wang, M.; Han, X.; Hansen, S.B. A Soluble Fluorescent Binding Assay Reveals PIP(2) Antagonism of TREK-1 Channels. Cell Rep. 2017, 20, 1287–1294. [Google Scholar] [CrossRef]
  78. Li, D.; Xia, L.; Chen, M.; Lin, C.; Wu, H.; Zhang, Y.; Pan, S.; Li, X. miR-133b, a particular member of myomiRs, coming into playing its unique pathological role in human cancer. Oncotarget 2017, 8, 50193–50208. [Google Scholar] [CrossRef]
  79. Verkhratsky, A.; Nedergaard, M. Physiology of Astroglia. Physiol. Rev. 2018, 98, 239–389. [Google Scholar] [CrossRef]
  80. Bukiya, A.N.; Dopico, A.M. Regulation of BK Channel Activity by Cholesterol and Its Derivatives. Adv. Exp. Med. Biol. 2019, 1115, 53–75. [Google Scholar]
  81. Burton, M.J.; Cresser-Brown, J.; Thomas, M.; Portolano, N.; Basran, J.; Freeman, S.L.; Kwon, H.; Bottrill, A.R.; Llansola-Portoles, M.J.; Pascal, A.A.; et al. Discovery of a heme-binding domain in a neuronal voltage-gated potassium channel. J. Biol. Chem. 2020, 295, 13277–13286. [Google Scholar] [CrossRef]
  82. Dudem, S.; Boon, P.X.; Mullins, N.; McClafferty, H.; Shipston, M.J.; Wilkinson, R.D.A.; Lobb, I.; Sergeant, G.P.; Thornbury, K.D.; Tikhonova, I.G.; et al. Oxidation modulates LINGO2-induced inactivation of large conductance, Ca(2+)-activated potassium channels. J. Biol. Chem. 2023, 299, 102975. [Google Scholar] [CrossRef]
  83. Dudem, S.; Large, R.J.; Kulkarni, S.; McClafferty, H.; Tikhonova, I.G.; Sergeant, G.P.; Keith D Thornbury, K.D.; Shipston, M.J.; Perrino, B.A.; Hollywood, M.A. LINGO1 is a regulatory subunit of large conductance, Ca(2+)-activated potassium channels. Proc. Natl. Acad. Sci. USA 2020, 117, 2194–2200. [Google Scholar] [CrossRef]
  84. Rotko, D.; Kunz, W.S.; Szewczyk, A.; Kulawiak, B. Signaling pathways targeting mitochondrial potassium channels. Int. J. Biochem. Cell Biol. 2020, 125, 105792. [Google Scholar] [CrossRef]
  85. Abbott, G.W. Control of Biophysical and Pharmacological Properties of Potassium Channels by Ancillary Subunits. Handb. Exp. Pharmacol. 2021, 267, 445–480. [Google Scholar]
  86. Abbott, G.W. Kv Channel Ancillary Subunits: Where Do We Go from Here? Physiology 2022, 37, 225–241. [Google Scholar] [CrossRef]
  87. North, K.C.; Zhang, M.; Singh, A.K.; Zaytseva, D.; Slayden, A.V.; Bukiya, A.N.; Dopico, A.M. Cholesterol Inhibition of Slo1 Channels Is Calcium-Dependent and Can Be Mediated by Either High-Affinity Calcium-Sensing Site in the Slo1 Cytosolic Tail. Mol. Pharmacol. 2022, 101, 132–143. [Google Scholar] [CrossRef]
  88. Kim, S.S.; Bae, Y.; Kwon, O.; Kwon, S.H.; Seo, J.B.; Hwang, E.M.; Park, J.Y. beta-COP Regulates TWIK1/TREK1 Heterodimeric Channel-Mediated Passive Conductance in Astrocytes. Cells 2022, 11, 3322. [Google Scholar] [CrossRef]
  89. Pardo, L.A. Voltage-Gated Potassium Channels Beyond the Action Potential. Bioelectricity 2022, 4, 117–125. [Google Scholar] [CrossRef]
  90. Vaithianathan, T.; Schneider, E.H.; Bukiya, A.N.; Dopico, A.M. Cholesterol and PIP(2) Modulation of BK(Ca) Channels. Adv. Exp. Med. Biol. 2023, 1422, 217–243. [Google Scholar]
  91. Huang, X.; He, Y.; Dubuc, A.M.; Hashizume, R.; Zhang, W.; Reimand, J.; Yang, H.; Wang, T.A.; Stehbens, S.J.; Younger, S.; et al. EAG2 potassium channel with evolutionarily conserved function as a brain tumor target. Nat. Neurosci. 2015, 18, 1236–1246. [Google Scholar] [CrossRef]
  92. Santos, R.; Ursu, O.; Gaulton, A.; Bento, A.P.; Donadi, R.S.; Bologa, C.G.; Karlsson, A.; Al-Lazikani, B.; Hersey, A.; Oprea, T.I.; et al. A comprehensive map of molecular drug targets. Nat. Rev. Drug Discov. 2017, 16, 19–34. [Google Scholar] [CrossRef]
  93. Hutchings, C.J.; Colussi, P.; Clark, T.G. Ion channels as therapeutic antibody targets. MAbs 2019, 11, 265–296. [Google Scholar] [CrossRef]
  94. Weise-Cross, L.; Resta, T.C.; Jernigan, N.L. Redox Regulation of Ion Channels and Receptors in Pulmonary Hypertension. Antioxid. Redox Signal 2019, 31, 898–915. [Google Scholar] [CrossRef]
  95. Wang, Y.; Qi, H.; Liu, Y.; Duan, C.; Liu, X.; Xia, T.; Chen, D.; Piao, H.L.; Liu, H.X. The double-edged roles of ROS in cancer prevention and therapy. Theranostics 2021, 11, 4839–4857. [Google Scholar] [CrossRef]
  96. Park, J.K.; Kim, Y.C.; Sim, J.H.; Choi, M.Y.; Choi, W.; Hwang, K.K.; Cho, M.C.; Kim, K.W.; Lim, S.W.; Lee, S.J. Regulation of membrane excitability by intracellular pH (pHi) changers through Ca2+-activated K+ current (BK channel) in single smooth muscle cells from rabbit basilar artery. Pflug. Arch. 2007, 454, 307–319. [Google Scholar] [CrossRef]
  97. Lewandowska, J.; Kalenik, B.; Wrzosek, A.; Szewczyk, A. Redox Regulation of Mitochondrial Potassium Channels Activity. Antioxidants 2024, 13, 434. [Google Scholar] [CrossRef]
  98. Lee, Y.M.; Kim, B.J.; Chun, Y.S.; So, I.; Choi, H.; Kim, M.S.; Park, J.W. NOX4 as an oxygen sensor to regulate TASK-1 activity. Cell Signal 2006, 18, 499–507. [Google Scholar] [CrossRef]
  99. Michl, J.; White, B.; Monterisi, S.; Bodmer, W.F.; Swietach, P. Phenotypic screen of sixty-eight colorectal cancer cell lines identifies CEACAM6 and CEACAM5 as markers of acid resistance. Proc. Natl. Acad. Sci. USA 2024, 121, e2319055121. [Google Scholar] [CrossRef]
  100. Selivanov, V.A.; Zeak, J.A.; Roca, J.; Cascante, M.; Trucco, M.; Votyakova, T.V. The role of external and matrix pH in mitochondrial reactive oxygen species generation. J. Biol. Chem. 2008, 283, 29292–29300. [Google Scholar] [CrossRef]
  101. Barar, J.; Omidi, Y. Dysregulated pH in Tumor Microenvironment Checkmates Cancer Therapy. Bioimpacts 2013, 3, 149–162. [Google Scholar]
  102. Damaghi, M.; Wojtkowiak, J.W.; Gillies, R.J. pH sensing and regulation in cancer. Front. Physiol. 2013, 4, 370. [Google Scholar] [CrossRef]
  103. Vrbka, L.; Vondrasek, J.; Jagoda-Cwiklik, B.; Vacha, R.; Jungwirth, P. Quantification and rationalization of the higher affinity of sodium over potassium to protein surfaces. Proc. Natl. Acad. Sci. USA 2006, 103, 15440–15444. [Google Scholar] [CrossRef]
  104. Tansel, B. Significance of thermodynamic and physical characteristics on permeation of ions during membrane separation: Hydrated radius, hydration free energy and viscous effects. Sep. Purif. Technol. 2012, 86, 119–126. [Google Scholar] [CrossRef]
  105. Danchin, A.; Nikel, P.I. Why Nature Chose Potassium. J. Mol. Evol. 2019, 87, 271–288. [Google Scholar] [CrossRef]
  106. Kopec, W.; Kopfer, D.A.; Vickery, O.N.; Bondarenko, A.S.; Jansen, T.L.C.; de Groot, B.L.; Zachariae, U. Direct knock-on of desolvated ions governs strict ion selectivity in K(+) channels. Nat. Chem. 2018, 10, 813–820. [Google Scholar] [CrossRef]
  107. Jing, Z.; Rackers, J.A.; Pratt, L.R.; Liu, C.; Rempe, S.B.; Ren, P. Thermodynamics of ion binding and occupancy in potassium channels. Chem. Sci. 2021, 12, 8920–8930. [Google Scholar] [CrossRef]
  108. Lam, C.K.; de Groot, B.L. Ion Conduction Mechanisms in Potassium Channels Revealed by Permeation Cycles. J. Chem. Theory Comput. 2023, 19, 2574–2589. [Google Scholar] [CrossRef]
  109. Collins, K.D. Charge density-dependent strength of hydration and biological structure. Biophys. J. 1997, 72, 65–76. [Google Scholar] [CrossRef]
  110. Nagy, I.Z.; Lustyik, G.; Nagy, V.Z.; Zarandi, B.; Bertoni-Freddari, C. Intracellular Na+:K+ ratios in human cancer cells as revealed by energy dispersive x-ray microanalysis. J. Cell Biol. 1981, 90, 769–777. [Google Scholar] [CrossRef]
  111. Leslie, T.K.; James, A.D.; Zaccagna, F.; Grist, J.T.; Deen, S.; Kennerley, A.; Riemer, F.; Kaggie, J.D.; Gallagher, F.A.; Gilbert, F.J.; et al. Sodium homeostasis in the tumour microenvironment. Biochim. Biophys. Acta Rev. Cancer 2019, 1872, 188304. [Google Scholar] [CrossRef]
  112. Allu, A.S.; Tiriveedhi, V. Cancer Salt Nostalgia. Cells 2021, 10, 1285. [Google Scholar] [CrossRef]
  113. Bortner, C.D.; Hughes, F.M., Jr.; Cidlowski, J.A. A primary role for K+ and Na+ efflux in the activation of apoptosis. J. Biol. Chem. 1997, 272, 32436–32442. [Google Scholar] [CrossRef]
  114. Pietrzak, M.; Meyerhoff, M.E. Determination of potassium in red blood cells using unmeasured volumes of whole blood and combined sodium/potassium-selective membrane electrode measurements. Anal. Chem. 2009, 81, 5961–5965. [Google Scholar] [CrossRef]
  115. Martinez-Valverde, T.; Vidal-Jorge, M.; Montoya, N.; Sanchez-Guerrero, A.; Manrique, S.; Munar, F.; Pellegri, M.D.; Poca, M.A.; Sahuquillo, J. Brain microdialysis as a tool to explore the ionic profile of the brain extracellular space in neurocritical patients: A methodological approach and feasibility study. J. Neurotrauma 2015, 32, 7–16. [Google Scholar] [CrossRef]
  116. Eil, R.; Vodnala, S.K.; Clever, D.; Klebanoff, C.A.; Sukumar, M.; Pan, J.H.; Palmer, D.C.; Gros, A.; Yamamoto, T.N.; Patel, S.J.; et al. Ionic immune suppression within the tumour microenvironment limits T cell effector function. Nature 2016, 537, 539–543. [Google Scholar] [CrossRef]
  117. Lehmann, H.P.; Henry, J.B. Appendix 4, Table A4-4. Whole Blood, Serum, and Plasma Chemistry. In Clinical Diagnosis and Management by Laboratory Methods; Henry, J.B., Ed.; W.B. Saunders: Philadelphia, PA, USA, 1996; pp. 1450–1454. [Google Scholar]
  118. Ghosh, S.; Yang, R.; Duraki, D.; Zhu, J.; Kim, J.E.; Jabeen, M.; Mao, C.; Dai, X.; Livezey, M.R.; Boudreau, M.W.; et al. Plasma Membrane Channel TRPM4 Mediates Immunogenic Therapy-Induced Necrosis. Cancer Res. 2023, 83, 3115–3130. [Google Scholar] [CrossRef]
  119. Oliver, D.; Baukrowitz, T.; Fakler, B. Polyamines as gating molecules of inward-rectifier K+ channels. Eur. J. Biochem. 2000, 267, 5824–5829. [Google Scholar] [CrossRef]
  120. Hopp, L.; Lasker, N.; Bamforth, R.; Aviv, A. Characterization of Na(+)-K+ homeostasis of cultured human skin fibroblasts in the presence and absence of fetal bovine serum. J. Cell Physiol. 1992, 151, 427–432. [Google Scholar] [CrossRef]
  121. Abdullah, N.A.; Hirata, M.; Matsumoto, K.; Aizawa, H.; Inoue, R.; Hamano, S.; Ikeda, S.; Xie, Z.; Hara, N.; Ito, Y. Contraction and depolarization induced by fetal bovine serum in airway smooth muscle. Am. J. Physiol. 1994, 266, L528–L535. [Google Scholar] [CrossRef]
  122. Marton, L.J.; Heby, O.; Levin, V.A.; Lubich, W.P.; Crafts, D.C.; Wilson, C.B. The relationship of polyamines in cerebrospinal fluid to the presence of central nervous system tumors. Cancer Res. 1976, 36, 973–977. [Google Scholar]
  123. Miska, J.; Rashidi, A.; Lee-Chang, C.; Gao, P.; Lopez-Rosas, A.; Zhang, P.; Burga, R.; Castro, B.; Xiao, T.; Han, Y.; et al. Polyamines drive myeloid cell survival by buffering intracellular pH to promote immunosuppression in glioblastoma. Sci. Adv. 2021, 7, eabc8929. [Google Scholar] [CrossRef]
  124. Monsorno, K.; Buckinx, A.; Paolicelli, R.C. Microglial metabolic flexibility: Emerging roles for lactate. Trends Endocrinol. Metab. 2022, 33, 186–195. [Google Scholar] [CrossRef]
  125. Augustynek, B.; Kudin, A.P.; Bednarczyk, P.; Szewczyk, A.; Kunz, W.S. Hemin inhibits the large conductance potassium channel in brain mitochondria: A putative novel mechanism of neurodegeneration. Exp. Neurol. 2014, 257, 70–75. [Google Scholar] [CrossRef]
  126. Bailey, C.S.; Moldenhauer, H.J.; Park, S.M.; Keros, S.; Meredith, A.L. KCNMA1-linked channelopathy. J. Gen. Physiol. 2019, 151, 1173–1189. [Google Scholar] [CrossRef]
  127. Zhao, Y.; Wang, D.; Xu, T.; Liu, P.; Cao, Y.; Wang, Y.; Yang, X.; Xu, X.; Wang, X.; Niu, H. Bladder cancer cells re-educate TAMs through lactate shuttling in the microfluidic cancer microenvironment. Oncotarget 2015, 6, 39196–39210. [Google Scholar] [CrossRef]
  128. Zhang, L.; Li, S. Lactic acid promotes macrophage polarization through MCT-HIF1alpha signaling in gastric cancer. Exp. Cell Res. 2020, 388, 111846. [Google Scholar] [CrossRef]
  129. Zhou, H.C.; Xin-Yan, Y.; Yu, W.W.; Liang, X.Q.; Du, X.Y.; Liu, Z.C.; Long, J.P.; Zhao, G.H.; Liu, H.B. Lactic acid in macrophage polarization: The significant role in inflammation and cancer. Int. Rev. Immunol. 2022, 41, 4–18. [Google Scholar] [CrossRef]
  130. Liebig, C.; Ayala, G.; Wilks, J.A.; Berger, D.H.; Albo, D. Perineural invasion in cancer: A review of the literature. Cancer 2009, 115, 3379–3391. [Google Scholar] [CrossRef]
  131. Bakst, R.L.; Wong, R.J. Mechanisms of Perineural Invasion. J. Neurol. Surg. B Skull Base 2016, 77, 96–106. [Google Scholar] [CrossRef]
  132. Narayan, P.; Flynn, J.; Zhang, Z.; Gillespie, E.F.; Mueller, B.; Xu, A.J.; Cuaron, J.; McCormick, B.; Khan, A.J.; Cahlon, O.; et al. Perineural invasion as a risk factor for locoregional recurrence of invasive breast cancer. Sci. Rep. 2021, 11, 12781. [Google Scholar] [CrossRef]
  133. Hosoya, K.; Wakahara, M.; Ikeda, K.; Umekita, Y. Perineural Invasion Predicts Unfavorable Prognosis in Patients With Invasive Breast Cancer. Cancer Diagn. Progn. 2023, 3, 208–214. [Google Scholar] [CrossRef]
  134. Beckner, M.E.; Fellows-Mayle, W.; Agostino, N.R.; Gobbel, G.T.; Pollack, I.F. Tumor markers in pseudopodia for invasive migration. In Progress in Tumor Marker Research; Swenson, L.I., Ed.; Nova Science Publishers, Inc.: New York, NY, USA, 2007; pp. 43–66. [Google Scholar]
  135. Beckner, M.E.; Gobbel, G.T.; Abounader, R.; Burovic, F.; Agostino, N.R.; Laterra, J.; Pollack, I.F. Glycolytic glioma cells with active glycogen synthase are sensitive to PTEN and inhibitors of PI3K and gluconeogenesis. Lab. Investig. 2005, 85, 1457–1470. [Google Scholar] [CrossRef]
  136. Beckner, M.E.; Zhang, Z.; Agostino, N.R.; Day, B.W.; Pollack, I.F. Albumin marks pseudopodia of astrocytoma cells responding to hepatocyte growth factor or serum. Lab. Investig. 2006, 86, 1103–1114. [Google Scholar] [CrossRef]
  137. Silverman, J.F.; Frable, W.J. The use of the diff-quik stain in the immediate interpretation of fine-needle aspiration biopsies. Diagn. Cytopathol. 1990, 6, 366–369. [Google Scholar] [CrossRef]
  138. Eagle, H. Amino acid metabolism in mammalian cell cultures. Science 1959, 130, 432–437. [Google Scholar] [CrossRef]
  139. Clark, M.J.; Homer, N.; O’Connor, B.D.; Chen, Z.; Eskin, A.; Lee, H.; Merriman, B.; Nelson, S.F. U87MG decoded: The genomic sequence of a cytogenetically aberrant human cancer cell line. PLoS Genet. 2010, 6, e1000832, Erratum in PLoS Genet. 2018, 14, e1007392. [Google Scholar] [CrossRef]
  140. Becker, H.M. Carbonic anhydrase IX and acid transport in cancer. Br. J. Cancer 2020, 122, 157–167. [Google Scholar] [CrossRef]
  141. Noor, S.I.; Jamali, S.; Ames, S.; Langer, S.; Deitmer, J.W.; Becker, H.M. A surface proton antenna in carbonic anhydrase II supports lactate transport in cancer cells. Elife 2018, 7, e35176. [Google Scholar] [CrossRef] [PubMed]
  142. Cheng, J.; Wen, J.; Wang, N.; Wang, C.; Xu, Q.; Yang, Y. Ion Channels and Vascular Diseases. Arterioscler. Thromb. Vasc. Biol. 2019, 39, e146–e156. [Google Scholar] [CrossRef] [PubMed]
  143. Crepalde, M.A.; Faria-Campos, A.C.; Campos, S.V. Modeling and analysis of cell membrane systems with probabilistic model checking. BMC Genom. 2011, 12 (Suppl. 4), S14. [Google Scholar] [CrossRef] [PubMed]
  144. Sajic, J.L.; Langthaler, S.; Baumgartner, C. Creating a novel mathematical model of the Kv10.1 ion channel and controlling channel activity with nanoelectromechanical systems. Appl. Sci. 2022, 12, 3836. [Google Scholar] [CrossRef]
  145. Sun, X.; Lee, H.C.; Lu, T. Sorbs2 Deficiency and Vascular BK Channelopathy in Diabetes. Circ. Res. 2024, 134, 858–871. [Google Scholar] [CrossRef]
  146. Wulff, H.; Castle, N.A. Therapeutic potential of KCa3.1 blockers: An overview of recent advances, and promising trends. Expert. Rev. Clin. Pharmacol. 2010, 3, 385–396. [Google Scholar] [CrossRef]
  147. Zaika, O.L.; Mamenko, M.; Palygin, O.; Boukelmoune, N.; Staruschenko, A.; Pochynyuk, O. Direct inhibition of basolateral Kir4.1/5.1 and Kir4.1 channels in the cortical collecting duct by dopamine. Am. J. Physiol. Renal Physiol. 2013, 305, F1277–F1287. [Google Scholar] [CrossRef]
  148. Fournier, P.A.; Fairchild, T.J.; Ferreira, L.D.; Brau, L. Post-exercise muscle glycogen repletion in the extreme: Effect of food absence and active recovery. J. Sports Sci. Med. 2004, 3, 139–146. [Google Scholar] [PubMed]
  149. Bangsbo, J.; Gollnick, P.D.; Graham, T.E.; Saltin, B. Substrates for muscle glycogen synthesis in recovery from intense exercise in man. J. Physiol. 1991, 434, 423–440. [Google Scholar] [CrossRef]
  150. Talmadge, R.J.; Scheide, J.I.; Silverman, H. Glycogen synthesis from lactate in a chronically active muscle. J. Appl. Physiol. 1989, 66, 2231–2238. [Google Scholar] [CrossRef]
  151. Hydrean, C.; Ghosh, A.; Nallin, M.; Ghosh, B.K. Interrelationship of carbohydrate metabolism and alkaline phosphatase synthesis in Bacillus licheniformis 749/c. J. Biol. Chem. 1977, 252, 6806–6812. [Google Scholar] [CrossRef]
  152. Tucci, V.; Isles, A.R.; Kelsey, G.; Ferguson-Smith, A.C.; Erice Imprinting, G. Genomic Imprinting and Physiological Processes in Mammals. Cell 2019, 176, 952–965. [Google Scholar] [CrossRef]
  153. Berlin, C.; Cottard, F.; Willmann, D.; Urban, S.; Tirier, S.M.; Marx, L.; Rippe, K.; Schmitt, M.; Petrocelli, V.; Greten, F.R.; et al. KMT9 Controls Stemness and Growth of Colorectal Cancer. Cancer Res. 2022, 82, 210–220. [Google Scholar] [CrossRef] [PubMed]
  154. Contreras-Espinosa, L.; Alcaraz, N.; De La Rosa-Velazquez, I.A.; Diaz-Chavez, J.; Cabrera-Galeana, P.; Rebollar-Vega, R.; Reynoso-Noverón, N.; Maldonado-Martínez, H.A.; González-Barrios, R.; Rogelio Montiel-Manriquez, R.; et al. Transcriptome Analysis Identifies GATA3-AS1 as a Long Noncoding RNA Associated with Resistance to Neoadjuvant Chemotherapy in Locally Advanced Breast Cancer Patients. J. Mol. Diagn. 2021, 23, 1306–1323. [Google Scholar] [CrossRef] [PubMed]
  155. Nargund, A.M.; Xu, C.; Mandoli, A.; Okabe, A.; Chen, G.B.; Huang, K.K.; Sheng, T.; Yao, X.; Teo, J.M.T.; Sundar, R.; et al. Chromatin Rewiring by Mismatch Repair Protein MSH2 Alters Cell Adhesion Pathways and Sensitivity to BET Inhibition in Gastric Cancer. Cancer Res. 2022, 82, 2538–2551. [Google Scholar] [CrossRef] [PubMed]
  156. Pine, A.R.; Cirigliano, S.M.; Singhania, R.; Nicholson, J.; da Silva, B.; Leslie, C.S.; Fine, H.A. Microenvironment-Driven Dynamic Chromatin Changes in Glioblastoma Recapitulate Early Neural Development at Single-Cell Resolution. Cancer Res. 2023, 83, 1581–1595. [Google Scholar] [CrossRef]
  157. Yu, Z.; Lv, Y.; Su, C.; Lu, W.; Zhang, R.; Li, J.; Guo, B.; Yan, H.; Liu, D.; Yang, Z. Integrative Single-Cell Analysis Reveals Transcriptional and Epigenetic Regulatory Features of Clear Cell Renal Cell Carcinoma. Cancer Res. 2023, 83, 700–719. [Google Scholar] [CrossRef]
  158. Zhu, Q.; Zhang, C.; Qu, T.; Lu, X.; He, X.; Li, W.; Yin, D.; Han, L.; Guo, R.; Zhang, E. MNX1-AS1 Promotes Phase Separation of IGF2BP1 to Drive c-Myc-Mediated Cell-Cycle Progression and Proliferation in Lung Cancer. Cancer Res. 2022, 82, 4340–4358. [Google Scholar] [CrossRef]
  159. Guetta-Terrier, C.; Karambizi, D.; Akosman, B.; Zepecki, J.P.; Chen, J.S.; Kamle, S.; Fajardo, E.; Fiser, A.; Singh, R.; Toms, S.A.; et al. Chi3l1 Is a Modulator of Glioma Stem Cell States and a Therapeutic Target in Glioblastoma. Cancer Res. 2023, 83, 1984–1999. [Google Scholar] [CrossRef]
  160. Mohan, D.R.; Borges, K.S.; Finco, I.; LaPensee, C.R.; Rege, J.; Solon, A.L.; Little, D.W.; Else, T.; Almeida, M.Q.; Dang, D.; et al. beta-Catenin-Driven Differentiation Is a Tissue-Specific Epigenetic Vulnerability in Adrenal Cancer. Cancer Res. 2023, 83, 2123–2141. [Google Scholar] [CrossRef] [PubMed]
  161. Abruzzo, L.V.; Barron, L.L.; Anderson, K.; Newman, R.J.; Wierda, W.G.; O’Brien, S.; Ferrajoli, A.; Luthra, M.; Talwalkar, S.; Luthra, R.; et al. Identification and validation of biomarkers of IgV(H) mutation status in chronic lymphocytic leukemia using microfluidics quantitative real-time polymerase chain reaction technology. J. Mol. Diagn. 2007, 9, 546–555. [Google Scholar] [CrossRef] [PubMed]
  162. Emmons, M.F.; Bennett, R.L.; Riva, A.; Gupta, K.; Carvalho, L.; Zhang, C.; Macaulay, R.; Dupéré-Richér, D.; Fang, B.; Seto, E.; et al. HDAC8-mediated inhibition of EP300 drives a transcriptional state that increases melanoma brain metastasis. Nat. Commun. 2023, 14, 7759. [Google Scholar] [CrossRef]
  163. Newman, A.C.; Kemp, A.J.; Drabsch, Y.; Behrends, C.; Wilkinson, S. Autophagy acts through TRAF3 and RELB to regulate gene expression via antagonism of SMAD proteins. Nat. Commun. 2017, 8, 1537. [Google Scholar] [CrossRef]
  164. Min, J.; Vega, P.N.; Engevik, A.C.; Williams, J.A.; Yang, Q.; Patterson, L.M.; Simmons, A.J.; Bliton, R.J.; Betts, J.W.; Lau, K.S.; et al. Heterogeneity and dynamics of active Kras-induced dysplastic lineages from mouse corpus stomach. Nat. Commun. 2019, 10, 5549. [Google Scholar] [CrossRef]
  165. Banito, A.; Li, X.; Laporte, A.N.; Roe, J.S.; Sanchez-Vega, F.; Huang, C.H.; Dancsok, A.R.; Hatzi, K.; Chen, C.C.; Tschaharganeh, D.F.; et al. The SS18-SSX Oncoprotein Hijacks KDM2B-PRC1.1 to Drive Synovial Sarcoma. Cancer Cell 2018, 33, 527–541.e8. [Google Scholar] [CrossRef]
  166. Pan, Y.; Han, H.; Hu, H.; Wang, H.; Song, Y.; Hao, Y.; Tong, X.; Patel, A.S.; Misirlioglu, S.; Tang, S.; et al. KMT2D deficiency drives lung squamous cell carcinoma and hypersensitivity to RTK-RAS inhibition. Cancer Cell 2023, 41, 88–105.e8. [Google Scholar] [CrossRef] [PubMed]
  167. Chapuy, B.; McKeown, M.R.; Lin, C.Y.; Monti, S.; Roemer, M.G.; Qi, J.; Rahl, P.B.; Sun, H.H.; Yeda, K.T.; Doench, J.G.; et al. Discovery and characterization of super-enhancer-associated dependencies in diffuse large B cell lymphoma. Cancer Cell 2013, 24, 777–790. [Google Scholar] [CrossRef]
  168. Beddows, I.; Fan, H.; Heinze, K.; Johnson, B.K.; Leonova, A.; Senz, J.; Djirackor, S.; Cho, K.R.; Pearce, C.L.; Huntsman, D.G.; et al. Cell State of Origin Impacts Development of Distinct Endometriosis-Related Ovarian Carcinoma Histotypes. Cancer Res. 2024, 84, 26–38. [Google Scholar] [CrossRef] [PubMed]
  169. Shih, A.H.; Jiang, Y.; Meydan, C.; Shank, K.; Pandey, S.; Barreyro, L.; Antony-Debre, I.; Viale, A.; Socci, N.; Sun, Y.; et al. Mutational cooperativity linked to combinatorial epigenetic gain of function in acute myeloid leukemia. Cancer Cell 2015, 27, 502–515. [Google Scholar] [CrossRef]
  170. Veschi, V.; Liu, Z.; Voss, T.C.; Ozbun, L.; Gryder, B.; Yan, C.; Hu, Y.; Ma, A.; Jin, J.; Mazur, S.J.; et al. Epigenetic siRNA and Chemical Screens Identify SETD8 Inhibition as a Therapeutic Strategy for p53 Activation in High-Risk Neuroblastoma. Cancer Cell 2017, 31, 50–63. [Google Scholar] [CrossRef]
  171. da Silva, L.S.; Mancano, B.M.; de Paula, F.E.; Dos Reis, M.B.; de Almeida, G.C.; Matsushita, M.; Junior, C.A.; Evangelista, A.F.; Saggioro, F.; Serafini, L.N.; et al. Expression of GNAS, TP53, and PTEN Improves the Patient Prognostication in Sonic Hedgehog (SHH) Medulloblastoma Subgroup. J. Mol. Diagn. 2020, 22, 957–966. [Google Scholar] [CrossRef]
  172. East, P.; Kelly, G.P.; Biswas, D.; Marani, M.; Hancock, D.C.; Creasy, T.; Sachsenmeier, K.; Swanton, C.; TRACERx Consortium; Downward, J.; et al. RAS oncogenic activity predicts response to chemotherapy and outcome in lung adenocarcinoma. Nat. Commun. 2022, 13, 5632. [Google Scholar] [CrossRef]
  173. Ishizawa, J.; Nakamaru, K.; Seki, T.; Tazaki, K.; Kojima, K.; Chachad, D.; Zhao, R.; Heese, L.; Ma, W.; Ma, M.C.J.; et al. Predictive Gene Signatures Determine Tumor Sensitivity to MDM2 Inhibition. Cancer Res. 2018, 78, 2721–2731. [Google Scholar] [CrossRef]
  174. Wang, X.Q.; Wang, X.Q.; Hsu, A.; Goytain, A.; Ng, T.L.T.; Nielsen, T.O. A Rapid and Cost-Effective Gene Expression Assay for the Diagnosis of Well-Differentiated and Dedifferentiated Liposarcomas. J. Mol. Diagn. 2021, 23, 274–284. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, L.B.; Karpova, A.; Gritsenko, M.A.; Kyle, J.E.; Cao, S.; Li, Y.; Rykunov, D.; Colaprico, A.; Rothstein, J.H.; Hong, R.; et al. Proteogenomic and metabolomic characterization of human glioblastoma. Cancer Cell 2021, 39, 509–528.e20. [Google Scholar] [CrossRef]
  176. Soragni, A.; Janzen, D.M.; Johnson, L.M.; Lindgren, A.G.; Thai-Quynh Nguyen, A.; Tiourin, E.; Antony-Debre, I.; Viale, A.; Socci, N.; Sun, Y.; et al. A Designed Inhibitor of p53 Aggregation Rescues p53 Tumor Suppression in Ovarian Carcinomas. Cancer Cell 2016, 29, 90–103. [Google Scholar] [CrossRef] [PubMed]
  177. Mello, S.S.; Valente, L.J.; Raj, N.; Seoane, J.A.; Flowers, B.M.; McClendon, J.; Bieging-Rolett, K.T.; Lee, J.; Ivanochko, D.; Kozak, M.M.; et al. A p53 Super-tumor Suppressor Reveals a Tumor Suppressive p53-Ptpn14-Yap Axis in Pancreatic Cancer. Cancer Cell 2017, 32, 460–473.e6. [Google Scholar] [CrossRef]
  178. Dong, L.; Lu, D.; Chen, R.; Lin, Y.; Zhu, H.; Zhang, Z.; Cai, S.; Cui, P.; Song, G.; Rao, D.; et al. Proteogenomic characterization identifies clinically relevant subgroups of intrahepatic cholangiocarcinoma. Cancer Cell 2022, 40, 70–87.e15. [Google Scholar] [CrossRef]
  179. Zhang, H.; Christensen, C.L.; Dries, R.; Oser, M.G.; Deng, J.; Diskin, B.; Li, F.; Pan, Y.; Zhang, X.; Yin, Y.; et al. CDK7 Inhibition Potentiates Genome Instability Triggering Anti-tumor Immunity in Small Cell Lung Cancer. Cancer Cell 2020, 37, 37–54.e9. [Google Scholar] [CrossRef]
  180. Schwab, A.; Siddiqui, A.; Vazakidou, M.E.; Napoli, F.; Bottcher, M.; Menchicchi, B.; Raza, U.; Saatci, O.; Krebs, A.M.; Ferrazzi, F.; et al. Polyol Pathway Links Glucose Metabolism to the Aggressiveness of Cancer Cells. Cancer Res. 2018, 78, 1604–1618. [Google Scholar] [CrossRef]
  181. Zhang, Y.; Cruickshanks, N.; Yuan, F.; Wang, B.; Pahuski, M.; Wulfkuhle, J.; Gallagher, I.; Koeppel, A.F.; Hatef, S.; Papanicolas, C.; et al. Targetable T-type Calcium Channels Drive Glioblastoma. Cancer Res. 2017, 77, 3479–3490. [Google Scholar] [CrossRef] [PubMed]
  182. Halvorsen, S.C.; Benita, Y.; Hopton, M.; Hoppe, B.; Gunnlaugsson, H.O.; Korgaonkar, P.; Vanderburg, C.R.; Nielsen, G.P.; Trepanowski, N.; Cheah, J.H.; et al. Transcriptional Profiling Supports the Notochordal Origin of Chordoma and Its Dependence on a TGFB1-TBXT Network. Am. J. Pathol. 2023, 193, 532–547. [Google Scholar] [CrossRef] [PubMed]
  183. Sahu, S.K.; Tiwari, N.; Pataskar, A.; Zhuang, Y.; Borisova, M.; Diken, M.; Strand, S.; Beli, P.; Tiwari, V.K. FBXO32 promotes microenvironment underlying epithelial-mesenchymal transition via CtBP1 during tumour metastasis and brain development. Nat. Commun. 2017, 8, 1523. [Google Scholar] [CrossRef] [PubMed]
  184. Verfaillie, A.; Imrichova, H.; Atak, Z.K.; Dewaele, M.; Rambow, F.; Hulselmans, G.; Christiaens, V.; Svetlichnyy, D.; Luciani, F.; Van den Mooter, L.; et al. Decoding the regulatory landscape of melanoma reveals TEADS as regulators of the invasive cell state. Nat. Commun. 2015, 6, 6683. [Google Scholar] [CrossRef]
  185. Wu, L.M.N.; Deng, Y.; Wang, J.; Zhao, C.; Wang, J.; Rao, R.; Xu, L.; Zhou, W.; Choi, K.; Rizvi, T.A.; et al. Programming of Schwann Cells by Lats1/2-TAZ/YAP Signaling Drives Malignant Peripheral Nerve Sheath Tumorigenesis. Cancer Cell 2018, 33, 292–308.e7. [Google Scholar] [CrossRef]
  186. Haider, S.; McIntyre, A.; van Stiphout, R.G.; Winchester, L.M.; Wigfield, S.; Harris, A.L.; Buffa, F.M. Genomic alterations underlie a pan-cancer metabolic shift associated with tumour hypoxia. Genome Biol. 2016, 17, 140. [Google Scholar] [CrossRef]
  187. Zhang, D.; Park, D.; Zhong, Y.; Lu, Y.; Rycaj, K.; Gong, S.; Chen, X.; Liu, X.; Chao, H.P.; Whitney, P.; et al. Stem cell and neurogenic gene-expression profiles link prostate basal cells to aggressive prostate cancer. Nat. Commun. 2016, 7, 10798. [Google Scholar] [CrossRef]
  188. Ni, Y.; Schmidt, K.R.; Werner, B.A.; Koenig, J.K.; Guldner, I.H.; Schnepp, P.M.; Tan, X.; Jiang, L.; Host, M.; Sun, L.; et al. Death effector domain-containing protein induces vulnerability to cell cycle inhibition in triple-negative breast cancer. Nat. Commun. 2019, 10, 2860. [Google Scholar] [CrossRef]
  189. Wang, Z.; Sun, D.; Chen, Y.J.; Xie, X.; Shi, Y.; Tabar, V.; Brennan, C.W.; Bale, T.A.; Jayewickreme, C.D.; Laks, D.R.; et al. Cell Lineage-Based Stratification for Glioblastoma. Cancer Cell 2020, 38, 366–379.e8. [Google Scholar] [CrossRef]
  190. Zhang, Z.; Karthaus, W.R.; Lee, Y.S.; Gao, V.R.; Wu, C.; Russo, J.W.; Liu, M.; Mota, J.M.; Abida, W.; Linton, E.; et al. Tumor Microenvironment-Derived NRG1 Promotes Antiandrogen Resistance in Prostate Cancer. Cancer Cell 2020, 38, 279–296.e9. [Google Scholar] [CrossRef] [PubMed]
  191. Cheung, W.K.; Zhao, M.; Liu, Z.; Stevens, L.E.; Cao, P.D.; Fang, J.E.; Westbrook, T.F.; Nguyen, D.X. Control of alveolar differentiation by the lineage transcription factors GATA6 and HOPX inhibits lung adenocarcinoma metastasis. Cancer Cell 2013, 23, 725–738. [Google Scholar] [CrossRef] [PubMed]
  192. Kelenis, D.P.; Rodarte, K.E.; Kollipara, R.K.; Pozo, K.; Choudhuri, S.P.; Spainhower, K.B.; Wait, S.J.; Stastny, V.; Oliver, T.G.; Johnson, J.E. Inhibition of Karyopherin beta1-Mediated Nuclear Import Disrupts Oncogenic Lineage-Defining Transcription Factor Activity in Small Cell Lung Cancer. Cancer Res. 2022, 82, 3058–3073. [Google Scholar] [CrossRef]
  193. Pozniak, J.; Nsengimana, J.; Laye, J.P.; O’Shea, S.J.; Diaz, J.M.S.; Droop, A.P.; Filia, A.; Harland, M.; Davies, J.R.; Mell, T.; et al. Genetic and Environmental Determinants of Immune Response to Cutaneous Melanoma. Cancer Res. 2019, 79, 2684–2696. [Google Scholar] [CrossRef]
  194. Van de Velde, L.A.; Allen, E.K.; Crawford, J.C.; Wilson, T.L.; Guy, C.S.; Russier, M.; Zeitler, L.; Bahrami, A.; Finkelstein, D.; Pelletier, S.; et al. Neuroblastoma Formation Requires Unconventional CD4 T Cells and Arginase-1-Dependent Myeloid Cells. Cancer Res. 2021, 81, 5047–5059. [Google Scholar] [CrossRef] [PubMed]
  195. Cai, L.; DeBerardinis, R.J.; Xie, Y.; Minna, J.D.; Xiao, G. A Comparative Study of Neuroendocrine Heterogeneity in Small Cell Lung Cancer and Neuroblastoma. Mol. Cancer Res. 2023, 21, 795–807. [Google Scholar] [CrossRef]
  196. Kawauchi, D.; Robinson, G.; Uziel, T.; Gibson, P.; Rehg, J.; Gao, C.; Finkelstein, D.; Qu, C.; Pounds, S.; Ellison, D.W.; et al. A mouse model of the most aggressive subgroup of human medulloblastoma. Cancer Cell 2012, 21, 168–180. [Google Scholar] [CrossRef]
  197. Topham, C.; Tighe, A.; Ly, P.; Bennett, A.; Sloss, O.; Nelson, L.; Ridgway, R.A.; Huels, D.; Littler, S.; Schandl, C.; et al. MYC Is a Major Determinant of Mitotic Cell Fate. Cancer Cell 2015, 28, 129–140. [Google Scholar] [CrossRef] [PubMed]
  198. Alam, A.; Levanduski, E.; Denz, P.; Villavicencio, H.S.; Bhatta, M.; Alhorebi, L.; Zhang, Y.; Gomez, E.C.; Morreale, B.; Senchanthisai, S.; et al. Fungal mycobiome drives IL-33 secretion and type 2 immunity in pancreatic cancer. Cancer Cell 2022, 40, 153–167.e11. [Google Scholar] [CrossRef] [PubMed]
  199. Altwegg, K.A.; Viswanadhapalli, S.; Mann, M.; Chakravarty, D.; Krishnan, S.; Liu, Z.; Liu, J.; Pratap, U.P.; Ebrahimi, B.; Sanchez, J.R.; et al. A First-in-Class Inhibitor of ER Coregulator PELP1 Targets ER+ Breast Cancer. Cancer Res. 2022, 82, 3830–3844. [Google Scholar] [CrossRef] [PubMed]
  200. Guieze, R.; Liu, V.M.; Rosebrock, D.; Jourdain, A.A.; Hernandez-Sanchez, M.; Martinez Zurita, A.; Sun, J.; Ten Hacken, E.; Baranowski, K.; Thompson, P.A.; et al. Mitochondrial Reprogramming Underlies Resistance to BCL-2 Inhibition in Lymphoid Malignancies. Cancer Cell 2019, 36, 369–384.e13. [Google Scholar] [CrossRef]
  201. Roh, W.; Geffen, Y.; Cha, H.; Miller, M.; Anand, S.; Kim, J.; Heiman, D.I.; Gainor, J.F.; Laird, P.W.; Cherniack, A.D.; et al. High-Resolution Profiling of Lung Adenocarcinoma Identifies Expression Subtypes with Specific Biomarkers and Clinically Relevant Vulnerabilities. Cancer Res. 2022, 82, 3917–3931. [Google Scholar] [CrossRef]
  202. Vidal, S.J.; Rodriguez-Bravo, V.; Quinn, S.A.; Rodriguez-Barrueco, R.; Lujambio, A.; Williams, E.; Sun, X.; de la Iglesia-Vicente, J.; Lee, A.; Readhead, B.; et al. A targetable GATA2-IGF2 axis confers aggressiveness in lethal prostate cancer. Cancer Cell 2015, 27, 223–239. [Google Scholar] [CrossRef]
  203. Alothman, S.J.; Kang, K.; Liu, X.; Krawczyk, E.; Azhar, R.I.; Hu, R.; Goerlitz, D.; Kallakury, B.V.; Furth, P.A. Characterization of transcriptome diversity and in vitro behavior of primary human high-risk breast cells. Sci. Rep. 2022, 12, 6159. [Google Scholar] [CrossRef]
  204. Chen, R.; Khatri, P.; Mazur, P.K.; Polin, M.; Zheng, Y.; Vaka, D.; Hoang, C.D.; Shrager, J.; Xu, Y.; Vicent, S.; et al. A meta-analysis of lung cancer gene expression identifies PTK7 as a survival gene in lung adenocarcinoma. Cancer Res. 2014, 74, 2892–2902. [Google Scholar] [CrossRef]
  205. Darmanis, S.; Sloan, S.A.; Croote, D.; Mignardi, M.; Chernikova, S.; Samghababi, P.; Zhang, Y.; Neff, N.; Kowarsky, M.; Caneda, C.; et al. Single-Cell RNA-Seq Analysis of Infiltrating Neoplastic Cells at the Migrating Front of Human Glioblastoma. Cell Rep. 2017, 21, 1399–1410. [Google Scholar] [CrossRef] [PubMed]
  206. de Jong, J.J.; Narayan, V.M.; Cronican, A.A.; Gupta, S.; van Leenders, G.J.L.H.; Boormans, J.L.; Gibb, E.A.; Konety, B.R. Gene expression profiling of muscle-invasive bladder cancer with secondary variant histology. Am. J. Clin. Pathol. 2021, 156, 895–905. [Google Scholar] [CrossRef] [PubMed]
  207. Harlow, M.L.; Maloney, N.; Roland, J.; Guillen Navarro, M.J.; Easton, M.K.; Kitchen-Goosen, S.M.; Boguslawski, E.A.; Madaj, Z.B.; Johnson, B.K.; Bowman, M.J.; et al. Lurbinectedin Inactivates the Ewing Sarcoma Oncoprotein EWS-FLI1 by Redistributing It within the Nucleus. Cancer Res. 2016, 76, 6657–6668. [Google Scholar] [CrossRef] [PubMed]
  208. Scheidmann, M.C.; Castro-Giner, F.; Strittmatter, K.; Krol, I.; Paasinen-Sohns, A.; Scherrer, R.; Donato, C.; Gkountela, S.; Szczerba, B.M.; Diamantopoulou, Z.; et al. An In Vivo CRISPR Screen Identifies Stepwise Genetic Dependencies of Metastatic Progression. Cancer Res. 2022, 82, 681–694. [Google Scholar] [CrossRef]
  209. Xu, Q.; Chen, J.; Ni, S.; Tan, C.; Xu, M.; Dong, L.; Yuan, L.; Wang, Q.; Du, X. Pan-cancer transcriptome analysis reveals a gene expression signature for the identification of tumor tissue origin. Mod. Pathol. 2016, 29, 546–556. [Google Scholar] [CrossRef]
  210. Zhang, S.; Zhou, Y.F.; Cao, J.; Burley, S.K.; Wang, H.Y.; Zheng, X.F.S. mTORC1 Promotes ARID1A Degradation and Oncogenic Chromatin Remodeling in Hepatocellular Carcinoma. Cancer Res. 2021, 81, 5652–5665. [Google Scholar] [CrossRef]
  211. Khlebus, E.; Vuttaradhi, V.K.; Welte, T.; Khurana, N.; Celestino, J.; Beird, H.C.; Gumbs, C.; Little, L. Legarreta, A.F.; Fellman, B.M.; et al. Comparative Tumor Microenvironment Analysis of Primary and Recurrent Ovarian Granulosa Cell Tumors. Mol. Cancer Res. 2023, 21, 483–494. [Google Scholar] [CrossRef]
  212. Mirzaei, R.; D’Mello, C.; Liu, M.; Nikolic, A.; Kumar, M.; Visser, F.; Bose, P.; Gallo, M.; Yong, V.W. Single-Cell Spatial Analysis Identifies Regulators of Brain Tumor-Initiating Cells. Cancer Res. 2023, 83, 1725–1741. [Google Scholar] [CrossRef]
  213. Yao, L.; Wang, J.T.; Jayasinghe, R.G.; O’Neal, J.; Tsai, C.F.; Rettig, M.P.; Guo, B.; Yan, H.; Liu, D.; Yang, Z. Single-Cell Discovery and Multiomic Characterization of Therapeutic Targets in Multiple Myeloma. Cancer Res. 2023, 83, 1214–1233. [Google Scholar] [CrossRef]
  214. Roda, N.; Cossa, A.; Hillje, R.; Tirelli, A.; Ruscitto, F.; Cheloni, S.; Priami, C.; Dalmasso, A.; Gambino, V.; Blandano, G.; et al. A Rare Subset of Primary Tumor Cells with Concomitant Hyperactivation of Extracellular Matrix Remodeling and dsRNA-IFN1 Signaling Metastasizes in Breast Cancer. Cancer Res. 2023, 83, 2155–2170. [Google Scholar] [CrossRef]
  215. Restaino, A.C.; Walz, A.; Vermeer, S.J.; Barr, J.; Kovacs, A.; Fettig, R.R.; Vermeer, D.W.; Reavis, H.; Williamson, C.S.; Lucido, C.T.; et al. Functional neuronal circuits promote disease progression in cancer. Sci. Adv. 2023, 9, eade4443. [Google Scholar] [CrossRef]
  216. Song, Y.; Deng, Z.; Sun, H.; Zhao, Y.; Zhao, R.; Cheng, J.; Huang, Q. Predicting tumor repopulation through the gene panel derived from radiation resistant colorectal cancer cells. J. Transl. Med. 2023, 21, 390. [Google Scholar] [CrossRef] [PubMed]
  217. Zhou, X.; An, J.; Kurilov, R.; Brors, B.; Hu, K.; Peccerella, T.; Roessler, S.; Pfütze, K.; Schulz, A.; Wolf, S.; et al. Persister cell phenotypes contribute to poor patient outcomes after neoadjuvant chemotherapy in PDAC. Nat. Cancer 2023, 4, 1362–1381. [Google Scholar] [CrossRef] [PubMed]
  218. Inman, G.J.; Wang, J.; Nagano, A.; Alexandrov, L.B.; Purdie, K.J.; Taylor, R.G.; Sherwood, V.; Thomson, J.; Hogan, S.; Spender, L.C.; et al. The genomic landscape of cutaneous SCC reveals drivers and a novel azathioprine associated mutational signature. Nat. Commun. 2018, 9, 3667. [Google Scholar] [CrossRef] [PubMed]
  219. Pandey, G.K.; Mitra, S.; Subhash, S.; Hertwig, F.; Kanduri, M.; Mishra, K.; Fransson, S.; Ganeshram, A.; Mondal, T.; Bandaru, S.; et al. The risk-associated long noncoding RNA NBAT-1 controls neuroblastoma progression by regulating cell proliferation and neuronal differentiation. Cancer Cell 2014, 26, 722–737. [Google Scholar] [CrossRef]
  220. Witt, H.; Mack, S.C.; Ryzhova, M.; Bender, S.; Sill, M.; Isserlin, R.; Benner, A.; Hielscher, T.; Milde, T.; Remke, M.; et al. Delineation of two clinically and molecularly distinct subgroups of posterior fossa ependymoma. Cancer Cell 2011, 20, 143–157. [Google Scholar] [CrossRef]
  221. Neou, M.; Villa, C.; Armignacco, R.; Jouinot, A.; Raffin-Sanson, M.L.; Septier, A.; Letourneur, F.; Diry, S.; Diedisheim, M.; Izac, B.; et al. Pangenomic Classification of Pituitary Neuroendocrine Tumors. Cancer Cell 2020, 37, 123–134.e5. [Google Scholar] [CrossRef]
  222. Fukumoto, T.; Umakoshi, H.; Iwahashi, N.; Ogasawara, T.; Yokomoto-Umakoshi, M.; Kaneko, H.; Fujita, M.; Uchida, N.; Nakao, H.; Kawamura, N.; et al. Steroids-producing nodules: A two-layered adrenocortical nodular structure as a precursor lesion of cortisol-producing adenoma. EBioMedicine 2024, 103, 105087. [Google Scholar] [CrossRef]
  223. Papier, K.; Atkins, J.R.; Tong, T.Y.N.; Gaitskell, K.; Desai, T.; Ogamba, C.F.; Parsaeian, M.; Reeves, G.K.; Mills, I.G.; Key, T.J.; et al. Identifying proteomic risk factors for cancer using prospective and exome analyses of 1463 circulating proteins and risk of 19 cancers in the UK Biobank. Nat. Commun. 2024, 15, 4010. [Google Scholar] [CrossRef]
  224. Smith-Byrne, K.; Hedman, A.; Dimitriou, M.; Desai, T.; Sokolov, A.V.; Schioth, H.B.; Koprulu, M.; Pietzner, M.; Langenberg, C.; Atkins, J.; et al. Identifying therapeutic targets for cancer among 2074 circulating proteins and risk of nine cancers. Nat. Commun. 2024, 15, 3621. [Google Scholar] [CrossRef]
  225. Suchak, S.K.; Baloyianni, N.V.; Perkinton, M.S.; Williams, R.J.; Meldrum, B.S.; Rattray, M. The ‘glial’ glutamate transporter, EAAT2 (Glt-1) accounts for high affinity glutamate uptake into adult rodent nerve endings. J. Neurochem. 2003, 84, 522–532. [Google Scholar] [CrossRef]
  226. Al Awabdh, S.; Gupta-Agarwal, S.; Sheehan, D.F.; Muir, J.; Norkett, R.; Twelvetrees, A.E.; Griffin, L.D.; Kittler, J.T. Neuronal activity mediated regulation of glutamate transporter GLT-1 surface diffusion in rat astrocytes in dissociated and slice cultures. Glia 2016, 64, 1252–1264. [Google Scholar] [CrossRef]
  227. Mai, D.; Chen, R.; Wang, J.; Zheng, J.; Zhang, X.; Qu, S. Critical amino acids in the TM2 of EAAT2 are essential for membrane-bound localization, substrate binding, transporter function and anion currents. J. Cell Mol. Med. 2021, 25, 2530–2548. [Google Scholar] [CrossRef]
  228. Molenaar, R.J. Ion channels in glioblastoma. ISRN Neurol. 2011, 2011, 590249. [Google Scholar] [CrossRef]
  229. Pedersen, S.F.; Stock, C. Ion channels and transporters in cancer: Pathophysiology, regulation, and clinical potential. Cancer Res. 2013, 73, 1658–1661. [Google Scholar] [CrossRef] [PubMed]
  230. Huang, X.; Jan, L.Y. Targeting potassium channels in cancer. J. Cell Biol. 2014, 206, 151–162. [Google Scholar] [CrossRef] [PubMed]
  231. Pardo, L.A.; Stuhmer, W. The roles of K(+) channels in cancer. Nat. Rev. Cancer 2014, 14, 39–48. [Google Scholar] [CrossRef] [PubMed]
  232. Comes, N.; Serrano-Albarras, A.; Capera, J.; Serrano-Novillo, C.; Condom, E.; Ramon, Y.C.S.; Ferreres, J.C.; Felipe, A. Involvement of potassium channels in the progression of cancer to a more malignant phenotype. Biochim. Biophys. Acta 2015, 1848, 2477–2492. [Google Scholar] [CrossRef]
  233. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion Channels in Cancer: Are Cancer Hallmarks Oncochannelopathies? Physiol. Rev. 2018, 98, 559–621. [Google Scholar] [CrossRef]
  234. Iorio, J.; Petroni, G.; Duranti, C.; Lastraioli, E. Potassium and Sodium Channels and the Warburg Effect: Biophysical Regulation of Cancer Metabolism. Bioelectricity 2019, 1, 188–200. [Google Scholar] [CrossRef]
  235. Girault, A.; Ahidouch, A.; Ouadid-Ahidouch, H. Roles for Ca(2+) and K(+) channels in cancer cells exposed to the hypoxic tumour microenvironment. Biochim. Biophys. Acta Mol. Cell Res. 2020, 1867, 118644. [Google Scholar] [CrossRef]
  236. Petho, Z.; Najder, K.; Carvalho, T.; McMorrow, R.; Todesca, L.M.; Rugi, M.; Bulk, E.; Chan, A.; Löwik, C.W.G.M.; Reshkin, S.J.; et al. pH-Channeling in Cancer: How pH-Dependence of Cation Channels Shapes Cancer Pathophysiology. Cancers 2020, 12, 2484. [Google Scholar] [CrossRef] [PubMed]
  237. Ganser, K.; Klumpp, L.; Bischof, H.; Lukowski, R.; Eckert, F.; Huber, S.M. Potassium Channels in Cancer. Handb. Exp. Pharmacol. 2021, 267, 253–275. [Google Scholar] [PubMed]
  238. Verkhratsky, A.; Parpura, V. Physiology of Astroglia: Channels, Receptors, Transporters, Ion Signaling, and Gliotransmission; Morgan & Claypool Publishers: San Rafael, CA, USA, 2015. [Google Scholar]
  239. Lopatin, A.N.; Makhina, E.N.; Nichols, C.G. The mechanism of inward rectification of potassium channels: “long-pore plugging” by cytoplasmic polyamines. J. Gen. Physiol. 1995, 106, 923–955. [Google Scholar] [CrossRef]
  240. Taura, J.; Kircher, D.M.; Gameiro-Ros, I.; Slesinger, P.A. Comparison of K(+) Channel Families. Handb. Exp. Pharmacol. 2021, 267, 1–49. [Google Scholar]
  241. Beckner, M.E. A roadmap for potassium buffering/dispersion via the glial network of the CNS. Neurochem. Int. 2020, 136, 104727. [Google Scholar] [CrossRef] [PubMed]
  242. Schroder, W.; Hinterkeuser, S.; Seifert, G.; Schramm, J.; Jabs, R.; Wilkin, G.P.; Steinhäuser, C. Functional and molecular properties of human astrocytes in acute hippocampal slices obtained from patients with temporal lobe epilepsy. Epilepsia 2000, 41 (Suppl. 6), S181–S184. [Google Scholar] [CrossRef]
  243. Larson, V.A.; Mironova, Y.; Vanderpool, K.G.; Waisman, A.; Rash, J.E.; Agarwal, A.; Bergles, D.E. Oligodendrocytes control potassium accumulation in white matter and seizure susceptibility. Elife 2018, 7, e34829. [Google Scholar] [CrossRef]
  244. Song, S.; Luo, L.; Sun, B.; Sun, D. Roles of glial ion transporters in brain diseases. Glia 2020, 68, 472–494. [Google Scholar] [CrossRef]
  245. Aronica, E.; Gorter, J.A.; Jansen, G.H.; Leenstra, S.; Yankaya, B.; Troost, D. Expression of connexin 43 and connexin 32 gap-junction proteins in epilepsy-associated brain tumors and in the perilesional epileptic cortex. Acta Neuropathol. 2001, 101, 449–459. [Google Scholar] [CrossRef]
  246. Caltabiano, R.; Torrisi, A.; Condorelli, D.; Albanese, V.; Lanzafame, S. High levels of connexin 43 mRNA in high grade astrocytomas. Study of 32 cases with in situ hybridization. Acta Histochem. 2010, 112, 529–535. [Google Scholar] [CrossRef]
  247. McClenaghan, C.; Schewe, M.; Aryal, P.; Carpenter, E.P.; Baukrowitz, T.; Tucker, S.J. Polymodal activation of the TREK-2 K2P channel produces structurally distinct open states. J. Gen. Physiol. 2016, 147, 497–505. [Google Scholar] [CrossRef]
  248. Woo, J.; Jun, Y.K.; Zhang, Y.H.; Nam, J.H.; Shin, D.H.; Kim, S.J. Identification of critical amino acids in the proximal C-terminal of TREK-2 K(+) channel for activation by acidic pH(i) and ATP-dependent inhibition. Pflugers Arch. 2018, 470, 327–337. [Google Scholar] [CrossRef]
  249. Lee, J.; Kang, M.; Kim, S.; Chang, I. Structural and molecular insight into the pH-induced low-permeability of the voltage-gated potassium channel Kv1.2 through dewetting of the water cavity. PLoS Comput. Biol. 2020, 16, e1007405. [Google Scholar] [CrossRef]
  250. Claydon, T.W.; Boyett, M.R.; Sivaprasadarao, A.; Ishii, K.; Owen, J.M.; O’Beirne, H.A.; Leach, R.; Komukai, K.; Orchard, C.H. Inhibition of the K+ channel kv1.4 by acidosis: Protonation of an extracellular histidine slows the recovery from N-type inactivation. J. Physiol. 2000, 526 Pt 2, 253–264. [Google Scholar] [CrossRef] [PubMed]
  251. Li, X.; Bett, G.C.; Jiang, X.; Bondarenko, V.E.; Morales, M.J.; Rasmusson, R.L. Regulation of N- and C-type inactivation of Kv1.4 by pHo and K+: Evidence for transmembrane communication. Am. J. Physiol. Heart Circ. Physiol. 2003, 284, H71–H80. [Google Scholar] [CrossRef] [PubMed]
  252. Abbott, G.W.; Butler, M.H.; Goldstein, S.A. Phosphorylation and protonation of neighboring MiRP2 sites: Function and pathophysiology of MiRP2-Kv3.4 potassium channels in periodic paralysis. FASEB J. 2006, 20, 293–301. [Google Scholar] [CrossRef]
  253. Kazmierczak, M.; Zhang, X.; Chen, B.; Mulkey, D.K.; Shi, Y.; Wagner, P.G.; Pivaroff-Ward, K.; Sassic, J.K.; Bayliss, D.A.; Jegla, T. External pH modulates EAG superfamily K+ channels through EAG-specific acidic residues in the voltage sensor. J. Gen. Physiol. 2013, 141, 721–735. [Google Scholar] [CrossRef] [PubMed]
  254. Mao, J.; Li, L.; McManus, M.; Wu, J.; Cui, N.; Jiang, C. Molecular determinants for activation of G-protein-coupled inward rectifier K+ (GIRK) channels by extracellular acidosis. J. Biol. Chem. 2002, 277, 46166–46171. [Google Scholar] [CrossRef]
  255. Mao, J.; Wu, J.; Chen, F.; Wang, X.; Jiang, C. Inhibition of G-protein-coupled inward rectifying K+ channels by intracellular acidosis. J. Biol. Chem. 2003, 278, 7091–7098. [Google Scholar] [CrossRef]
  256. Wu, J.; Xu, H.; Yang, Z.; Wang, Y.; Mao, J.; Jiang, C. Protons activate homomeric Kir6.2 channels by selective suppression of the long and intermediate closures. J. Membr. Biol. 2002, 190, 105–116. [Google Scholar] [CrossRef] [PubMed]
  257. Cui, N.; Wu, J.; Xu, H.; Wang, R.; Rojas, A.; Piao, H.; Mao, J.; Abdulkadir, L.; Li, L.; Jiang, C. A threonine residue (Thr71) at the intracellular end of the M1 helix plays a critical role in the gating of Kir6.2 channels by intracellular ATP and protons. J. Membr. Biol. 2003, 192, 111–122. [Google Scholar] [CrossRef]
  258. Maksaev, G.; Brundl-Jirout, M.; Stary-Weinzinger, A.; Zangerl-Plessl, E.M.; Lee, S.J.; Nichols, C.G. Subunit gating resulting from individual protonation events in Kir2 channels. Nat. Commun. 2023, 14, 4538. [Google Scholar] [CrossRef] [PubMed]
  259. Meuth, S.G.; Kleinschnitz, C.; Broicher, T.; Austinat, M.; Braeuninger, S.; Bittner, S.; Fischer, S.; Bayliss, D.A.; Budde, T.; Stoll, G.; et al. The neuroprotective impact of the leak potassium channel TASK1 on stroke development in mice. Neurobiol. Dis. 2009, 33, 1–11. [Google Scholar] [CrossRef]
  260. Buckler, K.J. TASK channels in arterial chemoreceptors and their role in oxygen and acid sensing. Pflugers Arch. 2015, 467, 1013–1025. [Google Scholar] [CrossRef] [PubMed]
  261. Decher, N.; Maier, M.; Dittrich, W.; Gassenhuber, J.; Bruggemann, A.; Busch, A.E.; Steinmeyer, K. Characterization of TASK-4, a novel member of the pH-sensitive, two-pore domain potassium channel family. FEBS Lett. 2001, 492, 84–89. [Google Scholar] [CrossRef]
  262. Bales, C.E.; Durfee, G.R. Part I: Cytologic Techniques. In Diagnostic Cytology and Its Histopathologic Bases; Koss, L.G., Ed.; J.B. Lippincott Company: Philadelphia, PA, USA, 1992; pp. 1484–1486. [Google Scholar]
  263. Folstad, L.; Look, M.; Pallavicini, M. A polycarbonate filter technique for collection of sorted cells. Cytometry 1982, 3, 64–65. [Google Scholar] [CrossRef]
  264. Druzhkova, I.; Lukina, M.; Dudenkova, V.; Ignatova, N.; Snopova, L.; Gavrina, A.; Shimolina, L.; Belousov, V.; Zagaynova, E.; Shirmanova, M.; et al. Tracing of intracellular pH in cancer cells in response to Taxol treatment. Cell Cycle 2021, 20, 1540–1551. [Google Scholar] [CrossRef]
  265. Lemasters, J.J. Molecular Mechanisms of Cell Death. In Molecular Pathology, The Molecular Basis of Human Disease; Coleman, W.B., Tsongalis, G.J., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 1–24. [Google Scholar]
  266. Green, D.R. Nonapoptotic cell death pathways. In Cell Death, Apoptosis and Other Means to an End; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2018; p. 135. [Google Scholar]
  267. Gow, I.F.; Thomson, J.; Davidson, J.; Shennan, D.B. The effect of a hyposmotic shock and purinergic agonists on K+(Rb+) efflux from cultured human breast cancer cells. Biochim. Biophys. Acta 2005, 1712, 52–61. [Google Scholar] [CrossRef]
  268. Nakada, T. The Molecular Mechanisms of Neural Flow Coupling: A New Concept. J. Neuroimaging 2015, 25, 861–865. [Google Scholar] [CrossRef]
  269. Li, C.; Voth, G.A. A quantitative paradigm for water-assisted proton transport through proteins and other confined spaces. Proc. Natl. Acad. Sci. USA 2021, 118, e2113141118. [Google Scholar] [CrossRef]
  270. Drew, D.; Boudker, O. Ion and lipid orchestration of secondary active transport. Nature 2024, 626, 963–974. [Google Scholar] [CrossRef] [PubMed]
  271. Persi, E.; Duran-Frigola, M.; Damaghi, M.; Roush, W.R.; Aloy, P.; Cleveland, J.L.; Gillies, R.J.; Ruppin, E. Systems analysis of intracellular pH vulnerabilities for cancer therapy. Nat. Commun. 2018, 9, 2997. [Google Scholar] [CrossRef] [PubMed]
  272. Russell, S.; Xu, L.; Kam, Y.; Abrahams, D.; Ordway, B.; Lopez, A.S.; Bui, M.M.; Johnson, J.; Epstein, T.; Ruiz, E.; et al. Proton export upregulates aerobic glycolysis. BMC Biol. 2022, 20, 163. [Google Scholar] [CrossRef] [PubMed]
  273. Swietach, P.; Spitzer, K.W.; Vaughan-Jones, R.D. pH-Dependence of extrinsic and intrinsic H(+)-ion mobility in the rat ventricular myocyte, investigated using flash photolysis of a caged-H(+) compound. Biophys. J. 2007, 92, 641–653. [Google Scholar] [CrossRef]
  274. Barron, T.; Yalcin, B.; Su, M.; Byun, Y.G.; Gavish, A.; Shamardani, K.; Xu, H.; Ni, L.; Soni, N.; Mehta, V.; et al. GABAergic neuron-to-glioma synapses in diffuse midline gliomas. Nature 2025, 639, 1060–1068. [Google Scholar] [CrossRef]
  275. Garofano, L.; Migliozzi, S.; Oh, Y.T.; D’Angelo, F.; Najac, R.D.; Ko, A.; Frangaj, B.; Caruso, F.P.; Yu, K.; Yuan, J.; et al. Pathway-based classification of glioblastoma uncovers a mitochondrial subtype with therapeutic vulnerabilities. Nat. Cancer 2021, 2, 141–156. [Google Scholar] [CrossRef]
  276. Xia, J.; Wang, H.; Li, S.; Wu, Q.; Sun, L.; Huang, H.; Zeng, M. Ion channels or aquaporins as novel molecular targets in gastric cancer. Mol. Cancer 2017, 16, 54. [Google Scholar] [CrossRef]
  277. Kurzeder, C.; Nguyen-Strauli, B.D.; Krol, I.; Ring, A.; Castro-Giner, F.; Nuesch, M.; Asawa, S.; Zhang, Y.W.; Budinjas, S.; Gvozdenovic, A.; et al. Digoxin for reduction of circulating tumor cell cluster size in metastatic breast cancer: A proof-of-concept trial. Nat. Med. 2025, 31, 1120–1124. [Google Scholar] [CrossRef]
  278. Moriyama, T.; Kataoka, H.; Seguchi, K.; Tsubouchi, H.; Koono, M. Effects of hepatocyte growth factor (HGF) on human glioma cells in vitro: HGF acts as a motility factor in glioma cells. Int. J. Cancer 1996, 66, 678–685. [Google Scholar] [CrossRef]
  279. Yamamoto, S.; Wakimoto, H.; Aoyagi, M.; Hirakawa, K.; Hamada, H. Modulation of motility and proliferation of glioma cells by hepatocyte growth factor. Jpn. J. Cancer Res. 1997, 88, 564–577. [Google Scholar] [CrossRef]
  280. Yuan, S.; Almagro, J.; Fuchs, E. Beyond genetics: Driving cancer with the tumour microenvironment behind the wheel. Nat. Rev. Cancer 2024, 24, 274–286. [Google Scholar] [CrossRef] [PubMed]
  281. Baumann, F.; Leukel, P.; Doerfelt, A.; Beier, C.P.; Dettmer, K.; Oefner, P.J.; Kastenberger, M.; Kreutz, M.; Nickl-Jockschat, T.; Bogdahn, U.; et al. Lactate promotes glioma migration by TGF-beta2-dependent regulation of matrix metalloproteinase-2. Neuro Oncol. 2009, 11, 368–380. [Google Scholar] [CrossRef] [PubMed]
  282. Hirschhaeuser, F.; Sattler, U.G.; Mueller-Klieser, W. Lactate: A metabolic key player in cancer. Cancer Res. 2011, 71, 6921–6925. [Google Scholar] [CrossRef] [PubMed]
  283. Seliger, C.; Leukel, P.; Moeckel, S.; Jachnik, B.; Lottaz, C.; Kreutz, M.; Brawanski, A.; Proescholdt, M.; Bogdahn, U.; Bosserhoff, A.K.; et al. Lactate-modulated induction of THBS-1 activates transforming growth factor (TGF)-beta2 and migration of glioma cells in vitro. PLoS ONE 2013, 8, e78935. [Google Scholar] [CrossRef]
  284. Ward, C.; Meehan, J.; Gray, M.E.; Murray, A.F.; Argyle, D.J.; Kunkler, I.H.; Langdon, S.P. The impact of tumour pH on cancer progression: Strategies for clinical intervention. Explor. Target. Antitumor Ther. 2020, 1, 71–100. [Google Scholar] [CrossRef]
  285. Zhou, R.; Tang, X.; Wang, Y. Emerging strategies to investigate the biology of early cancer. Nat. Rev. Cancer 2024, 24, 850–866. [Google Scholar] [CrossRef]
  286. Putnam, R.W.; Roos, A.; Wilding, T.J. Properties of the intracellular pH-regulating systems of frog skeletal muscle. J. Physiol. 1986, 381, 205–219. [Google Scholar] [CrossRef]
  287. Fraser, S.P.; Pardo, L.A. Ion channels: Functional expression and therapeutic potential in cancer. Colloquium on Ion Channels and Cancer. EMBO Rep. 2008, 9, 512–515. [Google Scholar] [CrossRef]
  288. Lang, F.; Stournaras, C. Ion channels in cancer: Future perspectives and clinical potential. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130108. [Google Scholar] [CrossRef]
  289. Martina, M. Ion channel screening: Advances in technologies and analysis. Front. Pharmacol. 2012, 3, 86. [Google Scholar] [CrossRef] [PubMed]
  290. Li, M.; Tian, P.; Zhao, Q.; Ma, X.; Zhang, Y. Potassium channels: Novel targets for tumor diagnosis and chemoresistance. Front. Oncol. 2023, 12, 1074469. [Google Scholar] [CrossRef]
  291. Kahleova, H.; Maracine, C.; Himmelfarb, J.; Jayaraman, A.; Znayenko-Miller, T.; Holubkov, R.; Barnard, N.D. Dietary acid load on the Mediterranean and a vegan diet: A secondary analysis of a randomized, cross-over trial. Front. Nutr. 2025, 12, 1634215. [Google Scholar] [CrossRef]
  292. Bretschneider, F.; Wrisch, A.; Lehmann-Horn, F.; Grissmer, S. External tetraethylammonium as a molecular caliper for sensing the shape of the outer vestibule of potassium channels. Biophys. J. 1999, 76, 2351–2360. [Google Scholar] [CrossRef]
  293. Somodi, S.; Hajdu, P.; Gaspar, R.; Panyi, G.; Varga, Z. Effects of changes in extracellular pH and potassium concentration on Kv1.3 inactivation. Eur. Biophys. J. 2008, 37, 1145–1156. [Google Scholar] [CrossRef]
  294. Somodi, S.; Varga, Z.; Hajdu, P.; Starkus, J.G.; Levy, D.I.; Gaspar, R.; Panyi, G. pH-dependent modulation of Kv1.3 inactivation: Role of His399. Am. J. Physiol. Cell Physiol. 2004, 287, C1067–C1076. [Google Scholar]
  295. Steidl, J.V.; Yool, A.J. Differential sensitivity of voltage-gated potassium channels Kv1.5 and Kv1.2 to acidic pH and molecular identification of pH sensor. Mol. Pharmacol. 1999, 55, 812–820. [Google Scholar] [CrossRef]
  296. Jager, H.; Grissmer, S. Regulation of a mammalian Shaker-related potassium channel, hKv1.5, by extracellular potassium and pH. FEBS Lett. 2001, 488, 45–50. [Google Scholar] [CrossRef]
  297. Kehl, S.J.; Eduljee, C.; Kwan, D.C.; Zhang, S.; Fedida, D. Molecular determinants of the inhibition of human Kv1.5 potassium currents by external protons and Zn(2+). J. Physiol. 2002, 541, 9–24. [Google Scholar] [CrossRef] [PubMed]
  298. Trapani, J.G.; Korn, S.J. Effect of external pH on activation of the Kv1.5 potassium channel. Biophys. J. 2003, 84, 195–204. [Google Scholar] [CrossRef] [PubMed]
  299. Zhang, S.; Kurata, H.T.; Kehl, S.J.; Fedida, D. Rapid induction of P/C-type inactivation is the mechanism for acid-induced K+ current inhibition. J. Gen. Physiol. 2003, 121, 215–225. [Google Scholar] [CrossRef] [PubMed]
  300. Fedida, D.; Zhang, S.; Kwan, D.C.; Eduljee, C.; Kehl, S.J. Synergistic inhibition of the maximum conductance of Kv1.5 channels by extracellular K+ reduction and acidification. Cell Biochem. Biophys. 2005, 43, 231–242. [Google Scholar] [CrossRef]
  301. Wang, S.; Ding, W.G.; Bai, J.Y.; Toyoda, F.; Wei, M.J.; Matsuura, H. Regulation of human cardiac Kv1.5 channels by extracellular acidification. Pflug. Arch. 2016, 468, 1885–1894. [Google Scholar] [CrossRef]
  302. Tipparaju, S.M.; Barski, O.A.; Srivastava, S.; Bhatnagar, A. Catalytic mechanism and substrate specificity of the beta-subunit of the voltage-gated potassium channel. Biochemistry 2008, 47, 8840–8854. [Google Scholar] [CrossRef] [PubMed]
  303. Raph, S.M.; Bhatnagar, A.; Nystoriak, M.A. Biochemical and physiological properties of K(+) channel-associated AKR6A (Kvbeta) proteins. Chem. Biol. Interact. 2019, 305, 21–27. [Google Scholar] [CrossRef]
  304. Singarayar, S.; Bursill, J.; Wyse, K.; Bauskin, A.; Wu, W.; Vandenberg, J.; Breit, S.; Campbell, T. Extracellular acidosis modulates drug block of Kv4.3 currents by flecainide and quinidine. J. Cardiovasc. Electrophysiol. 2003, 14, 641–650. [Google Scholar] [CrossRef]
  305. Singarayar, S.; Singleton, C.; Tie, H.; Wyse, K.; Bursill, J.; Bauskin, A.; Wu, W.; Valenzuela, S.; Breit, S.; Campbell, T. Effects of components of ischemia on the Kv4.3 current stably expressed in Chinese hamster ovary cells. J. Mol. Cell Cardiol. 2002, 34, 197–207. [Google Scholar] [CrossRef]
  306. Saegusa, N.; Garg, V.; Spitzer, K.W. Modulation of ventricular transient outward K(+) current by acidosis and its effects on excitation-contraction coupling. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H1680–H1696. [Google Scholar] [CrossRef]
  307. Heitzmann, D.; Grahammer, F.; von Hahn, T.; Schmitt-Graff, A.; Romeo, E.; Nitschke, R.; Gerlach, U.; Lang, H.J.; Verrey, F.; Barhanin, J.; et al. Heteromeric KCNE2/KCNQ1 potassium channels in the luminal membrane of gastric parietal cells. J. Physiol. 2004, 561, 547–557. [Google Scholar] [CrossRef] [PubMed]
  308. Heitzmann, D.; Koren, V.; Wagner, M.; Sterner, C.; Reichold, M.; Tegtmeier, I.; Volk, T.; Warth, R. KCNE beta subunits determine pH sensitivity of KCNQ1 potassium channels. Cell Physiol. Biochem. 2007, 19, 21–32. [Google Scholar] [CrossRef] [PubMed]
  309. Heitzmann, D.; Warth, R. No potassium, no acid: K+ channels and gastric acid secretion. Physiology 2007, 22, 335–341. [Google Scholar] [CrossRef]
  310. Abbott, G.W.; Roepke, T.K. KCNE2 and gastric cancer: Bench to bedside. Oncotarget 2016, 7, 17286–17287. [Google Scholar] [CrossRef]
  311. Engevik, A.C.; Kaji, I.; Goldenring, J.R. The Physiology of the Gastric Parietal Cell. Physiol. Rev. 2020, 100, 573–602. [Google Scholar] [CrossRef]
  312. Terlau, H.; Ludwig, J.; Steffan, R.; Pongs, O.; Stuhmer, W.; Heinemann, S.H. Extracellular Mg2+ regulates activation of rat eag potassium channel. Pflugers Arch. 1996, 432, 301–312. [Google Scholar] [CrossRef] [PubMed]
  313. Anumonwo, J.M.; Horta, J.; Delmar, M.; Taffet, S.M.; Jalife, J. Proton and zinc effects on HERG currents. Biophys. J. 1999, 77, 282–298. [Google Scholar] [CrossRef] [PubMed]
  314. Berube, J.; Chahine, M.; Daleau, P. Modulation of HERG potassium channel properties by external pH. Pflug. Arch. 1999, 438, 419–422. [Google Scholar]
  315. Jiang, M.; Dun, W.; Tseng, G.N. Mechanism for the effects of extracellular acidification on HERG-channel function. Am. J. Physiol. 1999, 277, H1283–H1292. [Google Scholar] [CrossRef]
  316. Jo, S.H.; Youm, J.B.; Kim, I.; Lee, C.O.; Earm, Y.E.; Ho, W.K. Blockade of HERG channels expressed in Xenopus oocytes by external H+. Pflug. Arch. 1999, 438, 23–29. [Google Scholar] [CrossRef]
  317. Terai, T.; Furukawa, T.; Katayama, Y.; Hiraoka, M. Effects of external acidosis on HERG current expressed in Xenopus oocytes. J. Mol. Cell Cardiol. 2000, 32, 11–21. [Google Scholar] [CrossRef]
  318. Vereecke, J.; Carmeliet, E. The effect of external pH on the delayed rectifying K+ current in cardiac ventricular myocytes. Pflug. Arch. 2000, 439, 739–751. [Google Scholar]
  319. Zhou, Q.; Bett, G.C. Regulation of the voltage-insensitive step of HERG activation by extracellular pH. Am. J. Physiol. Heart Circ. Physiol. 2010, 298, H1710–H1718. [Google Scholar] [CrossRef] [PubMed]
  320. Van Slyke, A.C.; Cheng, Y.M.; Mafi, P.; Allard, C.R.; Hull, C.M.; Shi, Y.P.; Claydon, T.W. Proton block of the pore underlies the inhibition of hERG cardiac K+ channels during acidosis. Am. J. Physiol. Cell Physiol. 2012, 302, C1797–C1806. [Google Scholar] [CrossRef]
  321. Shi, Y.P.; Cheng, Y.M.; Van Slyke, A.C.; Claydon, T.W. External protons destabilize the activated voltage sensor in hERG channels. Eur. Biophys. J. 2014, 43, 59–69. [Google Scholar] [CrossRef] [PubMed]
  322. Shi, Y.P.; Thouta, S.; Cheng, Y.M.; Claydon, T.W. Extracellular protons accelerate hERG channel deactivation by destabilizing voltage sensor relaxation. J. Gen. Physiol. 2019, 151, 231–246. [Google Scholar] [CrossRef] [PubMed]
  323. Shi, Y.P.; Thouta, S.; Claydon, T.W. Modulation of hERG K(+) Channel Deactivation by Voltage Sensor Relaxation. Front. Pharmacol. 2020, 11, 139. [Google Scholar] [CrossRef]
  324. Wilson, S.L.; Dempsey, C.E.; Hancox, J.C.; Marrion, N.V. Identification of a proton sensor that regulates conductance and open time of single hERG channels. Sci. Rep. 2019, 9, 19825. [Google Scholar] [CrossRef]
  325. Ukachukwu, C.U.; Jimenez-Vazquez, E.N.; Jain, A.; Jones, D.K. hERG1 channel subunit composition mediates proton inhibition of rapid delayed rectifier potassium current (I(Kr)) in cardiomyocytes derived from hiPSCs. J. Biol. Chem. 2023, 299, 102778. [Google Scholar] [CrossRef]
  326. Shi, W.; Wang, H.S.; Pan, Z.; Wymore, R.S.; Cohen, I.S.; McKinnon, D.; Dixon, J.E. Cloning of a mammalian elk potassium channel gene and EAG mRNA distribution in rat sympathetic ganglia. J. Physiol. 1998, 511 Pt 3, 675–682. [Google Scholar] [CrossRef] [PubMed]
  327. Sala-Rabanal, M.; Nichols, C.G. Inward rectifying potassium channels. In Handbook of Ion Channels; Zheng, J., Trudeau, M.C., Eds.; CRC Press: Boca Raton, FL, USA, 2015; pp. 241–260. [Google Scholar]
  328. Davies, N.W. Modulation of ATP-sensitive K+ channels in skeletal muscle by intracellular protons. Nature 1990, 343, 375–377. [Google Scholar] [CrossRef]
  329. Davies, N.W.; Standen, N.B.; Stanfield, P.R. The effect of intracellular pH on ATP-dependent potassium channels of frog skeletal muscle. J. Physiol. 1992, 445, 549–568. [Google Scholar] [CrossRef]
  330. Vivaudou, M.; Forestier, C. Modification by protons of frog skeletal muscle KATP channels: Effects on ion conduction and nucleotide inhibition. J. Physiol. 1995, 486 Pt 3, 629–645. [Google Scholar] [CrossRef]
  331. Forestier, C.; Pierrard, J.; Vivaudou, M. Mechanism of action of K channel openers on skeletal muscle KATP channels: Interactions with nucleotides and protons. J. Gen. Physiol. 1996, 107, 489–502. [Google Scholar] [CrossRef]
  332. Santa, N.; Kitazono, T.; Ago, T.; Ooboshi, H.; Kamouchi, M.; Wakisaka, M.; Ibayashi, S.; Iida, M. ATP-sensitive potassium channels mediate dilatation of basilar artery in response to intracellular acidification in vivo. Stroke 2003, 34, 1276–1280. [Google Scholar] [CrossRef]
  333. Butt, A.M.; Kalsi, A. Inwardly rectifying potassium channels (Kir) in central nervous system glia: A special role for Kir4.1 in glial functions. J. Cell Mol. Med. 2006, 10, 33–44. [Google Scholar] [CrossRef]
  334. Shuck, M.E.; Piser, T.M.; Bock, J.H.; Slightom, J.L.; Lee, K.S.; Bienkowski, M.J. Cloning and characterization of two K+ inward rectifier (Kir) 1.1 potassium channel homologs from human kidney (Kir1.2 and Kir1.3). J. Biol. Chem. 1997, 272, 586–593. [Google Scholar] [CrossRef]
  335. Yang, Z.; Jiang, C. Opposite effects of pH on open-state probability and single channel conductance of kir4.1 channels. J. Physiol. 1999, 520 Pt 3, 921–927. [Google Scholar] [CrossRef]
  336. Xu, H.; Yang, Z.; Cui, N.; Giwa, L.R.; Abdulkadir, L.; Patel, M.; Zeng, M. Molecular determinants for the distinct pH sensitivity of Kir1.1 and Kir4.1 channels. Am. J. Physiol. Cell Physiol. 2000, 279, C1464–C1471. [Google Scholar] [CrossRef]
  337. Xu, H.; Cui, N.; Yang, Z.; Qu, Z.; Jiang, C. Modulation of kir4.1 and kir5.1 by hypercapnia and intracellular acidosis. J. Physiol. 2000, 524 Pt 3, 725–735. [Google Scholar] [CrossRef] [PubMed]
  338. Tucker, S.J.; Imbrici, P.; Salvatore, L.; D’Adamo, M.C.; Pessia, M. pH dependence of the inwardly rectifying potassium channel, Kir5.1, and localization in renal tubular epithelia. J. Biol. Chem. 2000, 275, 16404–16407. [Google Scholar] [CrossRef] [PubMed]
  339. Tanemoto, M.; Kittaka, N.; Inanobe, A.; Kurachi, Y. In vivo formation of a proton-sensitive K+ channel by heteromeric subunit assembly of Kir5.1 with Kir4.1. J. Physiol. 2000, 525 Pt 3, 587–592. [Google Scholar] [CrossRef] [PubMed]
  340. Yang, Z.; Xu, H.; Cui, N.; Qu, Z.; Chanchevalap, S.; Shen, W.; Jiang, C. Biophysical and molecular mechanisms underlying the modulation of heteromeric Kir4.1-Kir5.1 channels by CO2 and pH. J. Gen. Physiol. 2000, 116, 33–45. [Google Scholar] [CrossRef] [PubMed]
  341. Pessia, M.; Imbrici, P.; D’Adamo, M.C.; Salvatore, L.; Tucker, S.J. Differential pH sensitivity of Kir4.1 and Kir4.2 potassium channels and their modulation by heteropolymerisation with Kir5.1. J. Physiol. 2001, 532, 359–367. [Google Scholar] [CrossRef]
  342. Hibino, H.; Fujita, A.; Iwai, K.; Yamada, M.; Kurachi, Y. Differential assembly of inwardly rectifying K+ channel subunits, Kir4.1 and Kir5.1, in brain astrocytes. J. Biol. Chem. 2004, 279, 44065–44073. [Google Scholar] [CrossRef]
  343. Sepulveda, F.V.; Pablo Cid, L.; Teulon, J.; Niemeyer, M.I. Molecular aspects of structure, gating, and physiology of pH-sensitive background K2P and Kir K+-transport channels. Physiol. Rev. 2015, 95, 179–217. [Google Scholar] [CrossRef]
  344. Hughes, B.A.; Swaminathan, A. Modulation of the Kir7.1 potassium channel by extracellular and intracellular pH. Am. J. Physiol. Cell Physiol. 2008, 294, C423–C431. [Google Scholar] [CrossRef]
  345. Lesage, F.; Guillemare, E.; Fink, M.; Duprat, F.; Lazdunski, M.; Romey, G.; Barhanin, J. TWIK-1, a ubiquitous human weakly inward rectifying K+ channel with a novel structure. EMBO J. 1996, 15, 1004–1011. [Google Scholar] [CrossRef] [PubMed]
  346. Lesage, F.; Lazdunski, M. Molecular and functional properties of two-pore-domain potassium channels. Am. J. Physiol. Renal Physiol. 2000, 279, F793–F801. [Google Scholar] [CrossRef]
  347. Rajan, S.; Plant, L.D.; Rabin, M.L.; Butler, M.H.; Goldstein, S.A. Sumoylation silences the plasma membrane leak K+ channel K2P1. Cell 2005, 121, 37–47. [Google Scholar] [CrossRef]
  348. Feliciangeli, S.; Chatelain, F.C.; Bichet, D.; Lesage, F. The family of K2P channels: Salient structural and functional properties. J. Physiol. 2015, 593, 2587–2603. [Google Scholar] [CrossRef]
  349. Turney, T.S.; Li, V.; Brohawn, S.G. Structural Basis for pH-gating of the K(+) channel TWIK1 at the selectivity filter. Nat. Commun. 2022, 13, 3232. [Google Scholar] [CrossRef] [PubMed]
  350. Chatelain, F.C.; Gilbert, N.; Bichet, D.; Jauch, A.; Feliciangeli, S.; Lesage, F.; Bignucolo, O. Mechanistic basis of the dynamic response of TWIK1 ionic selectivity to pH. Nat. Commun. 2024, 15, 3849. [Google Scholar] [CrossRef] [PubMed]
  351. Maingret, F.; Patel, A.J.; Lesage, F.; Lazdunski, M.; Honore, E. Mechano- or acid stimulation, two interactive modes of activation of the TREK-1 potassium channel. J. Biol. Chem. 1999, 274, 26691–26696. [Google Scholar] [CrossRef]
  352. Honore, E.; Maingret, F.; Lazdunski, M.; Patel, A.J. An intracellular proton sensor commands lipid- and mechano-gating of the K(+) channel TREK-1. EMBO J. 2002, 21, 2968–2976. [Google Scholar] [CrossRef]
  353. Cohen, A.; Ben-Abu, Y.; Hen, S.; Zilberberg, N. A novel mechanism for human K2P2.1 channel gating. Facilitation of C-type gating by protonation of extracellular histidine residues. J. Biol. Chem. 2008, 283, 19448–19455. [Google Scholar] [CrossRef]
  354. Sandoz, G.; Douguet, D.; Chatelain, F.; Lazdunski, M.; Lesage, F. Extracellular acidification exerts opposite actions on TREK1 and TREK2 potassium channels via a single conserved histidine residue. Proc. Natl. Acad. Sci. USA 2009, 106, 14628–14633. [Google Scholar] [CrossRef]
  355. Ryoo, K.; Park, J.Y. Two-pore Domain Potassium Channels in Astrocytes. Exp. Neurobiol. 2016, 25, 222–232. [Google Scholar] [CrossRef]
  356. Nasr, N.; Faucherre, A.; Borsotto, M.; Heurteaux, C.; Mazella, J.; Jopling, C.; Maati, H.M.O. Identification and characterization of two zebrafish Twik related potassium channels, Kcnk2a and Kcnk2b. Sci. Rep. 2018, 8, 15311. [Google Scholar] [CrossRef] [PubMed]
  357. Luo, Y.; Huang, L.; Liao, P.; Jiang, R. Contribution of Neuronal and Glial Two-Pore-Domain Potassium Channels in Health and Neurological Disorders. Neural Plast. 2021, 2021, 8643129. [Google Scholar] [CrossRef] [PubMed]
  358. Duprat, F.; Lesage, F.; Fink, M.; Reyes, R.; Heurteaux, C.; Lazdunski, M. TASK, a human background K+ channel to sense external pH variations near physiological pH. EMBO J. 1997, 16, 5464–5471. [Google Scholar] [CrossRef] [PubMed]
  359. Millar, J.A.; Barratt, L.; Southan, A.P.; Page, K.M.; Fyffe, R.E.; Robertson, B.; Mathie, A. A functional role for the two-pore domain potassium channel TASK-1 in cerebellar granule neurons. Proc. Natl. Acad. Sci. USA 2000, 97, 3614–3618. [Google Scholar] [CrossRef]
  360. Torborg, C.L.; Berg, A.P.; Jeffries, B.W.; Bayliss, D.A.; McBain, C.J. TASK-like conductances are present within hippocampal CA1 stratum oriens interneuron subpopulations. J. Neurosci. 2006, 26, 7362–7367. [Google Scholar] [CrossRef]
  361. Shvetsova, A.A.; Lazarenko, V.S.; Gaynullina, D.K.; Tarasova, O.S.; Schubert, R. TWIK-Related Acid-Sensitive Potassium Channels (TASK-1) Emerge as Contributors to Tone Regulation in Renal Arteries at Alkaline pH. Front. Physiol. 2022, 13, 895863. [Google Scholar] [CrossRef] [PubMed]
  362. Reyes, R.; Duprat, F.; Lesage, F.; Fink, M.; Salinas, M.; Farman, N.; Lazdunski, M. Cloning and expression of a novel pH-sensitive two pore domain K+ channel from human kidney. J. Biol. Chem. 1998, 273, 30863–30869. [Google Scholar] [CrossRef] [PubMed]
  363. Morton, M.J.; Abohamed, A.; Sivaprasadarao, A.; Hunter, M. pH sensing in the two-pore domain K+ channel, TASK2. Proc. Natl. Acad. Sci. USA 2005, 102, 16102–16106. [Google Scholar] [CrossRef] [PubMed]
  364. Gonczi, M.; Szentandrassy, N.; Johnson, I.T.; Heagerty, A.M.; Weston, A.H. Investigation of the role of TASK-2 channels in rat pulmonary arteries; pharmacological and functional studies following RNA interference procedures. Br. J. Pharmacol. 2006, 147, 496–505. [Google Scholar] [CrossRef]
  365. Niemeyer, M.I.; Cid, L.P.; Pena-Munzenmayer, G.; Sepulveda, F.V. Separate gating mechanisms mediate the regulation of K2P potassium channel TASK-2 by intra- and extracellular pH. J. Biol. Chem. 2010, 285, 16467–16475. [Google Scholar] [CrossRef]
  366. Zuniga, L.; Marquez, V.; Gonzalez-Nilo, F.D.; Chipot, C.; Cid, L.P.; Sepulveda, F.V.; Niemeyer, M.I. Gating of a pH-sensitive K(2P) potassium channel by an electrostatic effect of basic sensor residues on the selectivity filter. PLoS ONE 2011, 6, e16141. [Google Scholar] [CrossRef]
  367. Cid, L.P.; Roa-Rojas, H.A.; Niemeyer, M.I.; Gonzalez, W.; Araki, M.; Araki, K.; Sepulveda, F.V. TASK-2: A K2P K(+) channel with complex regulation and diverse physiological functions. Front. Physiol. 2013, 4, 198. [Google Scholar] [CrossRef]
  368. Bustos, D.; Bedoya, M.; Ramirez, D.; Concha, G.; Zuniga, L.; Decher, N.; Hernández-Rodríguez, E.W.; Sepúlveda, F.V.; Martínez, L.; Gonzále, W. Elucidating the Structural Basis of the Intracellular pH Sensing Mechanism of TASK-2 K(2)P Channels. Int. J. Mol. Sci. 2020, 21, 532. [Google Scholar] [CrossRef]
  369. Khoubza, L.; Gilbert, N.; Kim, E.J.; Chatelain, F.C.; Feliciangeli, S.; Abelanet, S.; Kang, D.; Lesage, F.; Bichet, D. Alkaline-sensitive two-pore domain potassium channels form functional heteromers in pancreatic beta-cells. J. Biol. Chem. 2022, 298, 102447. [Google Scholar] [CrossRef]
  370. Chavez, R.A.; Gray, A.T.; Zhao, B.B.; Kindler, C.H.; Mazurek, M.J.; Mehta, Y.; Forsayeth, J.R.; Yost, C.S. TWIK-2, a new weak inward rectifying member of the tandem pore domain potassium channel family. J. Biol. Chem. 1999, 274, 7887–7892. [Google Scholar] [CrossRef]
  371. Patel, A.J.; Maingret, F.; Magnone, V.; Fosset, M.; Lazdunski, M.; Honore, E. TWIK-2, an inactivating 2P domain K+ channel. J. Biol. Chem. 2000, 275, 28722–28730. [Google Scholar] [CrossRef]
  372. Rinne, S.; Schick, F.; Vowinkel, K.; Schutte, S.; Krasel, C.; Kauferstein, S.; Schäfer, M.K.J.; Kiper, A.K.; Müller, T.; Decher, N. Potassium channel TASK-5 forms functional heterodimers with TASK-1 and TASK-3 to break its silence. Nat. Commun. 2024, 15, 7548. [Google Scholar] [CrossRef]
  373. Cook, D.L.; Ikeuchi, M.; Fujimoto, W.Y. Lowering of pHi inhibits Ca2+-activated K+ channels in pancreatic B-cells. Nature 1984, 311, 269–271. [Google Scholar] [CrossRef]
  374. Christensen, O.; Zeuthen, T. Maxi K+ channels in leaky epithelia are regulated by intracellular Ca2+, pH and membrane potential. Pflug. Arch. 1987, 408, 249–259. [Google Scholar] [CrossRef]
  375. Kume, H.; Takagi, K.; Satake, T.; Tokuno, H.; Tomita, T. Effects of intracellular pH on calcium-activated potassium channels in rabbit tracheal smooth muscle. J. Physiol. 1990, 424, 445–457. [Google Scholar] [CrossRef] [PubMed]
  376. Peers, C.; Green, F.K. Inhibition of Ca(2+)-activated K+ currents by intracellular acidosis in isolated type I cells of the neonatal rat carotid body. J. Physiol. 1991, 437, 589–602. [Google Scholar] [CrossRef] [PubMed]
  377. Schubert, R.; Krien, U.; Gagov, H. Protons inhibit the BK(Ca) channel of rat small artery smooth muscle cells. J. Vasc. Res. 2001, 38, 30–38. [Google Scholar] [CrossRef]
  378. Avdonin, V.; Tang, X.D.; Hoshi, T. Stimulatory action of internal protons on Slo1 BK channels. Biophys. J. 2003, 84, 2969–2980. [Google Scholar] [CrossRef] [PubMed]
  379. Brelidze, T.I.; Magleby, K.L. Protons block BK channels by competitive inhibition with K+ and contribute to the limits of unitary currents at high voltages. J. Gen. Physiol. 2004, 123, 305–319. [Google Scholar] [CrossRef]
  380. Raingo, J.; Rebolledo, A.; Grassi de Gende, A.O.; Sanz, N.; Tommasi, J.; Milesi, V. pH effects on high conductance Ca2+-activated K+ channels (BK(Ca)) in human internal mammary artery smooth muscle cells. Life Sci. 2005, 77, 1993–2003. [Google Scholar] [CrossRef]
  381. Petroff, E.; Snitsarev, V.; Gong, H.; Abboud, F.M. Acid sensing ion channels regulate neuronal excitability by inhibiting BK potassium channels. Biochem. Biophys. Res. Commun. 2012, 426, 511–515. [Google Scholar] [CrossRef]
  382. Petroff, E.Y.; Price, M.P.; Snitsarev, V.; Gong, H.; Korovkina, V.; Abboud, F.M.; Welsh, M.J. Acid-sensing ion channels interact with and inhibit BK K+ channels. Proc. Natl. Acad. Sci. USA 2008, 105, 3140–3144. [Google Scholar] [CrossRef]
  383. Hou, S.; Horrigan, F.T.; Xu, R.; Heinemann, S.H.; Hoshi, T. Comparative effects of H+ and Ca2+ on large-conductance Ca2+- and voltage-gated Slo1 K+ channels. Channels 2009, 3, 249–258. [Google Scholar] [CrossRef]
  384. Yang, H.; Cui, J. BK channels. In Handbook of Ion Channels; Zheng, J., Trudeau, M.C., Eds.; CRC Press: Boca Raton, FL, USA, 2015; pp. 227–239. [Google Scholar]
  385. Guarina, L.; Vandael, D.H.; Carabelli, V.; Carbone, E. Low pH(o) boosts burst firing and catecholamine release by blocking TASK-1 and BK channels while preserving Cav1 channels in mouse chromaffin cells. J. Physiol. 2017, 595, 2587–2609. [Google Scholar] [CrossRef]
  386. Zhou, Y.; Xia, X.M.; Lingle, C.J. BK channel inhibition by strong extracellular acidification. Elife 2018, 7, e38060. [Google Scholar] [CrossRef] [PubMed]
  387. Dopico, A.M.; Bukiya, A.N.; Jaggar, J.H. Calcium- and voltage-gated BK channels in vascular smooth muscle. Pflug. Arch. 2018, 470, 1271–1289. [Google Scholar] [CrossRef] [PubMed]
  388. Sancho, M.; Kyle, B.D. The Large-Conductance, Calcium-Activated Potassium Channel: A Big Key Regulator of Cell Physiology. Front. Physiol. 2021, 12, 750615. [Google Scholar] [CrossRef] [PubMed]
  389. Yang, C.T.; Zeng, X.H.; Xia, X.M.; Lingle, C.J. Interactions between beta subunits of the KCNMB family and Slo3: Beta4 selectively modulates Slo3 expression and function. PLoS ONE 2009, 4, e6135. [Google Scholar] [CrossRef]
  390. Wen, D.; Cornelius, R.J.; Yuan, Y.; Sansom, S.C. Regulation of BK-alpha expression in the distal nephron by aldosterone and urine pH. Am. J. Physiol. Renal Physiol. 2013, 305, F463–F476. [Google Scholar] [CrossRef]
  391. Wen, D.; Cornelius, R.J.; Rivero-Hernandez, D.; Yuan, Y.; Li, H.; Weinstein, A.M.; Sansom, S.C. Relation between BK-alpha/beta4-mediated potassium secretion and ENaC-mediated sodium reabsorption. Kidney Int. 2014, 86, 139–145. [Google Scholar] [CrossRef]
  392. Wang, B.; Wang-France, J.; Li, H.; Sansom, S.C. Furosemide reduces BK-alphabeta4-mediated K(+) secretion in mice on an alkaline high-K(+) diet. Am. J. Physiol. Renal Physiol. 2019, 316, F341–F350. [Google Scholar] [CrossRef]
  393. Al-Qusairi, L.; Ferdaus, M.Z.; Pham, T.D.; Li, D.; Grimm, P.R.; Zapf, A.M.; Abood, D.C.; Tahaei, E.; Delpire, E.; Wal, S.M.; et al. Dietary anions control potassium excretion: It is more than a poorly absorbable anion effect. Am. J. Physiol. Renal Physiol. 2023, 325, F377–F393. [Google Scholar] [CrossRef] [PubMed]
  394. Goodchild, S.J.; Lamy, C.; Seutin, V.; Marrion, N.V. Inhibition of K(Ca)2.2 and K(Ca)2.3 channel currents by protonation of outer pore histidine residues. J. Gen. Physiol. 2009, 134, 295–308. [Google Scholar] [CrossRef] [PubMed]
  395. Srivastava, S.; Li, Z.; Ko, K.; Choudhury, P.; Albaqumi, M.; Johnson, A.K.; Yan, Y.; Backer, J.M.; Unutmaz, D.; Coetzee, W.A.; et al. Histidine phosphorylation of the potassium channel KCa3.1 by nucleoside diphosphate kinase B is required for activation of KCa3.1 and CD4 T cells. Mol. Cell 2006, 24, 665–675. [Google Scholar] [CrossRef]
  396. Kee, J.M.; Muir, T.W. Chasing phosphohistidine, an elusive sibling in the phosphoamino acid family. ACS Chem. Biol. 2012, 7, 44–51. [Google Scholar] [CrossRef]
  397. Kalagiri, R.; Hunter, T. The many ways that nature has exploited the unusual structural and chemical properties of phosphohistidine for use in proteins. Biochem. J. 2021, 478, 3575–3596. [Google Scholar] [CrossRef]
  398. Ning, J.; Sala, M.; Reina, J.; Kalagiri, R.; Hunter, T.; McCullough, B.S. Histidine Phosphorylation: Protein Kinases and Phosphatases. Int. J. Mol. Sci. 2024, 25, 7975. [Google Scholar] [CrossRef] [PubMed]
  399. Yamane, T.; Furukawa, T.; Horikawa, S.; Hiraoka, M. External pH regulates the slowly activating potassium current IsK expressed in Xenopus oocytes. FEBS Lett. 1993, 319, 229–232. [Google Scholar] [CrossRef]
  400. Freeman, L.C.; Lippold, J.J.; Mitchell, K.E. Glycosylation influences gating and pH sensitivity of I(sK). J Membr Biol 2000, 177, 65–79. [Google Scholar] [CrossRef]
  401. Unsold, B.; Kerst, G.; Brousos, H.; Hubner, M.; Schreiber, R.; Nitschke, R.; Greger, R.; Bleich, M. KCNE1 reverses the response of the human K+ channel KCNQ1 to cytosolic pH changes and alters its pharmacology and sensitivity to temperature. Pflug. Arch. 2000, 441, 368–378. [Google Scholar] [CrossRef]
  402. Peretz, A.; Schottelndreier, H.; Aharon-Shamgar, L.B.; Attali, B. Modulation of homomeric and heteromeric KCNQ1 channels by external acidification. J. Physiol. 2002, 545, 751–766. [Google Scholar] [CrossRef]
  403. Roepke, T.K.; King, E.C.; Purtell, K.; Kanda, V.A.; Lerner, D.J.; Abbott, G.W. Genetic dissection reveals unexpected influence of beta subunits on KCNQ1 K+ channel polarized trafficking in vivo. FASEB J. 2011, 25, 727–736. [Google Scholar] [CrossRef]
  404. Dixit, G.; Dabney-Smith, C.; Lorigan, G.A. The membrane protein KCNQ1 potassium ion channel: Functional diversity and current structural insights. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183148. [Google Scholar] [CrossRef] [PubMed]
  405. Abbott, G.W. KCNQs: Ligand- and Voltage-Gated Potassium Channels. Front. Physiol. 2020, 11, 583. [Google Scholar] [CrossRef]
  406. Prole, D.L.; Lima, P.A.; Marrion, N.V. Mechanisms underlying modulation of neuronal KCNQ2/KCNQ3 potassium channels by extracellular protons. J. Gen. Physiol. 2003, 122, 775–793. [Google Scholar] [CrossRef] [PubMed]
  407. Mehrdel, B.; Villalba-Galea, C.A. Effect of a sensing charge mutation on the deactivation of KV7.2 channels. J. Gen. Physiol. 2024, 156, e202213284. [Google Scholar] [CrossRef] [PubMed]
  408. Jensen, H.S.; Callo, K.; Jespersen, T.; Jensen, B.S.; Olesen, S.P. The KCNQ5 potassium channel from mouse: A broadly expressed M-current like potassium channel modulated by zinc, pH, and volume changes. Brain Res. Mol. Brain Res. 2005, 139, 52–62. [Google Scholar] [CrossRef] [PubMed]
  409. Ruffin, V.A.; Gu, X.Q.; Zhou, D.; Douglas, R.M.; Sun, X.; Trouth, C.O.; Haddad, G.G. The sodium-activated potassium channel Slack is modulated by hypercapnia and acidosis. Neuroscience 2008, 151, 410–418. [Google Scholar] [CrossRef]
  410. Ruffin, V.A.; Salameh, A.I.; Boron, W.F.; Parker, M.D. Intracellular pH regulation by acid-base transporters in mammalian neurons. Front. Physiol. 2014, 5, 43. [Google Scholar] [CrossRef]
  411. Kaczmarek, L.K. Slack, Slick and Sodium-Activated Potassium Channels. ISRN Neurosci. 2013, 2013, 354262. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  412. Kaczmarek, L.K.; Aldrich, R.W.; Chandy, K.G.; Grissmer, S.; Wei, A.D.; Wulff, H. International Union of Basic and Clinical Pharmacology. C. Nomenclature and Properties of Calcium-Activated and Sodium-Activated Potassium Channels. Pharmacol. Rev. 2017, 69, 1–11. [Google Scholar] [CrossRef]
  413. Schreiber, M.; Wei, A.; Yuan, A.; Gaut, J.; Saito, M.; Salkoff, L. Slo3, a novel pH-sensitive K+ channel from mammalian spermatocytes. J. Biol. Chem. 1998, 273, 3509–3516. [Google Scholar] [CrossRef] [PubMed]
  414. Zhang, X.; Zeng, X.; Lingle, C.J. Slo3 K+ channels: Voltage and pH dependence of macroscopic currents. J. Gen. Physiol. 2006, 128, 317–336. [Google Scholar] [CrossRef] [PubMed]
  415. Zhang, X.; Zeng, X.; Xia, X.M.; Lingle, C.J. pH-regulated Slo3 K+ channels: Properties of unitary currents. J. Gen. Physiol. 2006, 128, 301–315. [Google Scholar] [CrossRef] [PubMed]
  416. Navarro, B.; Kirichok, Y.; Clapham, D.E. KSper, a pH-sensitive K+ current that controls sperm membrane potential. Proc. Natl. Acad. Sci. USA 2007, 104, 7688–7692. [Google Scholar] [CrossRef] [PubMed]
  417. Leonetti, M.D.; Yuan, P.; Hsiung, Y.; Mackinnon, R. Functional and structural analysis of the human SLO3 pH- and voltage-gated K+ channel. Proc. Natl. Acad. Sci. USA 2012, 109, 19274–19279. [Google Scholar] [CrossRef]
  418. Brenker, C.; Zhou, Y.; Muller, A.; Echeverry, F.A.; Trotschel, C.; Poetsch, A.; Xia, X.M.; Bönigk, W.; Lingle, C.J.; Kaupp, U.B.; et al. The Ca2+-activated K+ current of human sperm is mediated by Slo3. Elife 2014, 3, e01438. [Google Scholar] [CrossRef]
  419. Ludwig, M.G.; Vanek, M.; Guerini, D.; Gasser, J.A.; Jones, C.E.; Junker, U.; Hofstetter, H.; Wolf, R.M.; Seuwen, K. Proton-sensing G-protein-coupled receptors. Nature 2003, 425, 93–98. [Google Scholar] [CrossRef]
  420. Weiss, K.T.; Fante, M.; Kohl, G.; Schreml, J.; Haubner, F.; Kreutz, M.; Haverkampf, S.; Berneburg, M.; Schreml, S. Proton-sensing G protein-coupled receptors as regulators of cell proliferation and migration during tumor growth and wound healing. Exp. Dermatol. 2017, 26, 127–132. [Google Scholar] [CrossRef]
  421. Hosford, P.S.; Mosienko, V.; Kishi, K.; Jurisic, G.; Seuwen, K.; Kinzel, B.; Ludwig, M.G.; Wells, J.A.; Christie, I.N.; Koolen, L.; et al. CNS distribution, signalling properties and central effects of G-protein coupled receptor 4. Neuropharmacology 2018, 138, 381–392. [Google Scholar] [CrossRef]
  422. Mashiko, M.; Kurosawa, A.; Tani, Y.; Tsuji, T.; Takeda, S. GPR31 and GPR151 are activated under acidic conditions. J. Biochem. 2019, 166, 317–322. [Google Scholar] [CrossRef] [PubMed]
  423. Klatt, W.; Wallner, S.; Brochhausen, C.; Stolwijk, J.A.; Schreml, S. Expression profiles of proton-sensing G-protein coupled receptors in common skin tumors. Sci. Rep. 2020, 10, 15327. [Google Scholar] [CrossRef]
  424. Sisignano, M.; Fischer, M.J.M.; Geisslinger, G. Proton-Sensing GPCRs in Health and Disease. Cells 2021, 10, 2050. [Google Scholar] [CrossRef] [PubMed]
  425. Forch, A.; Wallner, S.; Zeman, F.; Ettl, T.; Brochhausen, C.; Schreml, S. Expression of Proton-Sensitive GPR31, GPR151, TASK1 and TASK3 in Common Skin Tumors. Cells 2021, 11, 27. [Google Scholar] [CrossRef]
  426. Imenez Silva, P.H.; Camara, N.O.; Wagner, C.A. Role of proton-activated G protein-coupled receptors in pathophysiology. Am. J. Physiol. Cell Physiol. 2022, 323, C400–C414. [Google Scholar] [CrossRef]
  427. Imenez Silva, P.H.; Wagner, C.A. Physiological relevance of proton-activated GPCRs. Pflug. Arch. 2022, 474, 487–504. [Google Scholar] [CrossRef]
  428. Kumar, N.N.; Velic, A.; Soliz, J.; Shi, Y.; Li, K.; Wang, S.; Weaver, J.L.; Sen, J.; Abbott, S.B.G.; Lazarenko, R.M.; et al. PHYSIOLOGY. Regulation of breathing by CO(2) requires the proton-activated receptor GPR4 in retrotrapezoid nucleus neurons. Science 2015, 348, 1255–1260. [Google Scholar] [CrossRef]
  429. Guyenet, P.G.; Bayliss, D.A. Neural Control of Breathing and CO2 Homeostasis. Neuron 2015, 87, 946–961. [Google Scholar] [CrossRef] [PubMed]
  430. Ramadoss, S.; Qin, J.; Tao, B.; Thomas, N.E.; Cao, E.; Wu, R.; Sandoval, D.R.; Piermatteo, A.; Grunddal, K.V.; Ma, F.; et al. Bone-marrow macrophage-derived GPNMB protein binds to orphan receptor GPR39 and plays a critical role in cardiac repair. Nat. Cardiovasc. Res. 2024, 3, 1356–1373. [Google Scholar] [CrossRef]
  431. Wang, J.; Simonavicius, N.; Wu, X.; Swaminath, G.; Reagan, J.; Tian, H.; Ling, L. Kynurenic acid as a ligand for orphan G protein-coupled receptor GPR35. J. Biol. Chem. 2006, 281, 22021–22028. [Google Scholar] [CrossRef]
  432. Wang, J.; Wu, X.; Simonavicius, N.; Tian, H.; Ling, L. Medium-chain fatty acids as ligands for orphan G protein-coupled receptor GPR84. J. Biol. Chem. 2006, 281, 34457–34464. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Settled glioblastoma cells deprived of extracellular fluid. (A,B) The morphology of each cell line (A172, LN229, T98G, and U87) after cell settling with fluid wicked away, included a few clumps of robust appearing, viable tumor cells clustered around single degenerating cells of the same lineage. Although Diff Quik staining varied, some viable clustered cells’ cytoplasm stained basophilic. Black and white photographs suggest some thin strands between cells crossing narrow linear spaces separating them. Single and grouped cells without a visible degenerating central cell were also seen. (C) Cytoplasmic vacuolization in U87 cells is shown in representative single and grouped cells, outlined digitally with white strokes around each merged layer. The basophilic cytoplasm seen in some cells would be consistent with an alkaline pH, but no measurements were made. Numbers of vacuoles were highest in single cells and decreased as numbers of cells in cellular groups increased. (D) For counts performed on cell monolayers from two separate wells, 51% of 75 and 72% of 25 single cells were vacuolated,1st and 2nd wells, respectively. However, vacuolated cells in groups were only 11% of 122 and 18% of 39 cell groups with greater than twenty cells, 1st and 2nd wells, respectively. Pearson regressions of the numbers of cells in groups with vacuoles had R2 values of 0.266 (p = 0.0410) and 0.544 (p = 0.0062), 1st and 2nd wells, respectively. Diff Quik stain. Scale bars, 25 μm.
Figure 1. Settled glioblastoma cells deprived of extracellular fluid. (A,B) The morphology of each cell line (A172, LN229, T98G, and U87) after cell settling with fluid wicked away, included a few clumps of robust appearing, viable tumor cells clustered around single degenerating cells of the same lineage. Although Diff Quik staining varied, some viable clustered cells’ cytoplasm stained basophilic. Black and white photographs suggest some thin strands between cells crossing narrow linear spaces separating them. Single and grouped cells without a visible degenerating central cell were also seen. (C) Cytoplasmic vacuolization in U87 cells is shown in representative single and grouped cells, outlined digitally with white strokes around each merged layer. The basophilic cytoplasm seen in some cells would be consistent with an alkaline pH, but no measurements were made. Numbers of vacuoles were highest in single cells and decreased as numbers of cells in cellular groups increased. (D) For counts performed on cell monolayers from two separate wells, 51% of 75 and 72% of 25 single cells were vacuolated,1st and 2nd wells, respectively. However, vacuolated cells in groups were only 11% of 122 and 18% of 39 cell groups with greater than twenty cells, 1st and 2nd wells, respectively. Pearson regressions of the numbers of cells in groups with vacuoles had R2 values of 0.266 (p = 0.0410) and 0.544 (p = 0.0062), 1st and 2nd wells, respectively. Diff Quik stain. Scale bars, 25 μm.
Biomolecules 15 01177 g001
Figure 2. Migration of glioblastoma cell lines to hepatocyte growth factor (HGF) and fetal bovine serum (FBS). (A) Representative Boyden assays of 2 rat astrocytoma (C6 and F98) and 6 glioblastoma cell lines (T98G, LN18, U87, LN229, A172, and U373) showed background (unstimulated) cell migration (row 1), cell migration to HGF (row 2), HGF and FBS (rows 3–9), and to FBS alone (rows 10–12) using concentrations in the bottom wells as indicated on the left. With the same amount of HGF present, adding increasing amounts of FBS increased chemotactic cell migration in all cell lines except A172. Images were obtained with a transparency scanner. (B) Portions of assays (regular scanner) with FBS alone as the chemoattractant were digitized for densitometry, with results (average total pixels) shown. All cell lines showed significant migration to FBS except A172, which only had a trend, p = 0.0638, for 0.5% FBS. All assays (3 per cell line) are in Appendix A, for C6, F98, T98G, and LN18 (Figure A2) and U87, LN229, A172, and U373 (Figure A3). (C) Immunoblot with densitometry (total pixels) underneath quantifies HGF’s receptor, Met (β subunit, 145 kDa), among cell lines. For comparison, HGF-stimulated cell migration is below the bar graph as times (X) background cell migration and lacks notable correlation. The immunoblot’s gel is in Appendix A (Figure A4A). (D) MEM as media in the bottom wells, with and without FBS, outperformed Dulbecco’s PBS for U87 cell migration with “Migration” media in all top wells. Migration was 1.7X, 1.1X, and 1.4X greater (significantly, p < 0.05) using MEM compared to PBS for 0, 0.1, and 1% FBS in the bottom wells, respectively. MEM’s greater potassium content (5.3 mM) in the bottom wells than in “Migration” media (4.5 mM K+; see text) in the top wells putatively provided a chemoattractant K+ gradient. p =0.0144 in MEM versus Dulbecco’s PBS with 0% FBS, and with 0.1 and 1% FBS it was also present, p = 0.0136 and 0.000422, respectively. The scanned image of the assay is in Appendix A (Figure A4B). (E) Chemotaxis to HGF and FBS outperformed chemokinesis and FBS enhanced HGF’s effects on U87 cells in MEM (top and bottom wells). Chemotaxis was consistently greater than chemokinesis in paired comparisons designated by brackets under the x-axis with significance, p = 0.012, for row 8 versus row 9. The assay image is in Appendix A (Figure A4C). Averages of rows’ densitometry units indicated by bar heights with 95% confidence intervals. Variable staining of background migration and assay-to-assay variation for each cell line in Appendix A are routine. Each circle of migrated cells is 3 mm in diameter.
Figure 2. Migration of glioblastoma cell lines to hepatocyte growth factor (HGF) and fetal bovine serum (FBS). (A) Representative Boyden assays of 2 rat astrocytoma (C6 and F98) and 6 glioblastoma cell lines (T98G, LN18, U87, LN229, A172, and U373) showed background (unstimulated) cell migration (row 1), cell migration to HGF (row 2), HGF and FBS (rows 3–9), and to FBS alone (rows 10–12) using concentrations in the bottom wells as indicated on the left. With the same amount of HGF present, adding increasing amounts of FBS increased chemotactic cell migration in all cell lines except A172. Images were obtained with a transparency scanner. (B) Portions of assays (regular scanner) with FBS alone as the chemoattractant were digitized for densitometry, with results (average total pixels) shown. All cell lines showed significant migration to FBS except A172, which only had a trend, p = 0.0638, for 0.5% FBS. All assays (3 per cell line) are in Appendix A, for C6, F98, T98G, and LN18 (Figure A2) and U87, LN229, A172, and U373 (Figure A3). (C) Immunoblot with densitometry (total pixels) underneath quantifies HGF’s receptor, Met (β subunit, 145 kDa), among cell lines. For comparison, HGF-stimulated cell migration is below the bar graph as times (X) background cell migration and lacks notable correlation. The immunoblot’s gel is in Appendix A (Figure A4A). (D) MEM as media in the bottom wells, with and without FBS, outperformed Dulbecco’s PBS for U87 cell migration with “Migration” media in all top wells. Migration was 1.7X, 1.1X, and 1.4X greater (significantly, p < 0.05) using MEM compared to PBS for 0, 0.1, and 1% FBS in the bottom wells, respectively. MEM’s greater potassium content (5.3 mM) in the bottom wells than in “Migration” media (4.5 mM K+; see text) in the top wells putatively provided a chemoattractant K+ gradient. p =0.0144 in MEM versus Dulbecco’s PBS with 0% FBS, and with 0.1 and 1% FBS it was also present, p = 0.0136 and 0.000422, respectively. The scanned image of the assay is in Appendix A (Figure A4B). (E) Chemotaxis to HGF and FBS outperformed chemokinesis and FBS enhanced HGF’s effects on U87 cells in MEM (top and bottom wells). Chemotaxis was consistently greater than chemokinesis in paired comparisons designated by brackets under the x-axis with significance, p = 0.012, for row 8 versus row 9. The assay image is in Appendix A (Figure A4C). Averages of rows’ densitometry units indicated by bar heights with 95% confidence intervals. Variable staining of background migration and assay-to-assay variation for each cell line in Appendix A are routine. Each circle of migrated cells is 3 mm in diameter.
Biomolecules 15 01177 g002
Figure 3. KCN DEGs and related DEGs detected in E1–E70 studies. (A) The 53 KCN DEGs are listed horizontally, with specific studies listed vertically showing their frequency in 70 studies. The 3 most frequent tumors (breast, glioblastoma, and lung) have color-coded superscripts. The 33 (62.3%) KCN DEGs that encode H+-sensitive proteins are labeled with red H+ above each column and have their names underlined in red. A total of 52 (74.3%) of 70 studies contain at least one KCN DEG encoding an H+-sensitive protein. Yellow shading indicates the study was in 1 or more of 8 categories (see Figure 4A). Non-coding KCN DEGs are in separate histograms. (B) A stochastic view of K+-related dynamics is based on collective incidences of KCN and related DEGs in E1–E70. The most frequent members of KCN subfamilies with pertinent non-KCN proteins shown reflect literature representations [140,141,142,143,144,145,146,147]. Relatively few studies in E1–E70 detected DEGs encoding Na/K-ATPases (9% for 5 genes), the monocarboxylate transporter complex (5% for 3 genes), lactate dehydrogenases (3% for 1 gene), taurine transporters (2% for 2 genes), and 0% for those not listed. Therefore, it is proposed that abnormal K+ flux via KCN DEGs (assorted combinations) initiates pH changes by releasing protons for transport (putatively Grotthuss-type transport) to exit sites using relatively intact supporting systems. Specifically, abnormal accumulation of K+ at inner-membrane surfaces that forces exchange with bound H+ is proposed to initiate pH reversal. Also, movement of K+-7H2O (2 H2O per K+ in “cross-section”) may bring water to inner-membrane regions to enhance Grotthuss conditions in cancer cells. Putative K+-related changes in cancer cells maintaining relatively normal expression of pertinent non-KCN genes, as seen in most of the 70 KCN DEG studies, would be consistent with pH reversal, as shown (B). E1–E70 citations are listed in Section 3.3.6.
Figure 3. KCN DEGs and related DEGs detected in E1–E70 studies. (A) The 53 KCN DEGs are listed horizontally, with specific studies listed vertically showing their frequency in 70 studies. The 3 most frequent tumors (breast, glioblastoma, and lung) have color-coded superscripts. The 33 (62.3%) KCN DEGs that encode H+-sensitive proteins are labeled with red H+ above each column and have their names underlined in red. A total of 52 (74.3%) of 70 studies contain at least one KCN DEG encoding an H+-sensitive protein. Yellow shading indicates the study was in 1 or more of 8 categories (see Figure 4A). Non-coding KCN DEGs are in separate histograms. (B) A stochastic view of K+-related dynamics is based on collective incidences of KCN and related DEGs in E1–E70. The most frequent members of KCN subfamilies with pertinent non-KCN proteins shown reflect literature representations [140,141,142,143,144,145,146,147]. Relatively few studies in E1–E70 detected DEGs encoding Na/K-ATPases (9% for 5 genes), the monocarboxylate transporter complex (5% for 3 genes), lactate dehydrogenases (3% for 1 gene), taurine transporters (2% for 2 genes), and 0% for those not listed. Therefore, it is proposed that abnormal K+ flux via KCN DEGs (assorted combinations) initiates pH changes by releasing protons for transport (putatively Grotthuss-type transport) to exit sites using relatively intact supporting systems. Specifically, abnormal accumulation of K+ at inner-membrane surfaces that forces exchange with bound H+ is proposed to initiate pH reversal. Also, movement of K+-7H2O (2 H2O per K+ in “cross-section”) may bring water to inner-membrane regions to enhance Grotthuss conditions in cancer cells. Putative K+-related changes in cancer cells maintaining relatively normal expression of pertinent non-KCN genes, as seen in most of the 70 KCN DEG studies, would be consistent with pH reversal, as shown (B). E1–E70 citations are listed in Section 3.3.6.
Biomolecules 15 01177 g003
Figure 4. Oncogenes and DNA modifiers may play a role in cancer studies via KCN DEGs encoding H+-sensitive proteins. (A) KCN DEGs that encode H+-sensitive proteins are highlighted with red underlining of study names. Histograms of 8 categories for cancer studies among E1–E70 are shown on the left (yellow boxes). The right histogram also includes 20 studies in “no category” (white boxes). Comparison of studies in “no category” versus those in at least one category shown in the left histogram found a significant difference in the number of studies detecting KCN DEGs encoding H+-sensitive proteins using Fisher’s exact test, p = 0.0325, indicating a higher likelihood for studies included in at least one category. (BD) Word searches of PubMed for numbers of oncogene and histone publications related to H+, K+, BK(KCNMA1), and Cl (contrasting ion) were performed to further evaluate categories in E1–E70. (B) On 12 May 2025, searches for each category of “TP53”, “MYC”, “TGFbeta”, “BCL2”, “EGFR”, “histone”, “RAS”, and “MET” were performed separately for “proton”, “potassium”, and “chloride”. For all 8 categories, the numbers of citations for “proton”, “potassium”, and “chloride” were 1565, 1418, and 1261, respectively, and Pearson’s product-moment correlations for “proton” with “potassium” and “proton” with “chloride” were R = 0.891 (p = 0.00148) and 0.599 (p = 0.0584), respectively. Pearson’s product-moment correlation for “potassium” and “chloride” was 0.689 (p = 0.0295). R2 for “proton” with “potassium” was 0.79. “BK” and “potassium” searches on 7 June 2025revealed “BK” as a potential large component of “potassium” citations, approximately one-third for 7 categories. (C) When searches also included “cancer” as a third word, the numbers of citations for “proton”, “potassium”, “BK”, and “chloride” on 14 May 2025, were 1061, 577, 513 (88.9% of “potassium” search results), and 952, respectively. When MET was left out for better “proton” correlations, the numbers of citations for “proton”, “potassium”, “BK”, and “chloride” on 14 May 2025, were 728, 496, 439 (88.5% of “potassium” search results), and 879, respectively, and Pearson’s product-moment correlations for “proton” with “potassium” and “chloride” were R = 0.860 (p = 0.00649) and 0.462 (p = 0.148), respectively, across the 7 remaining categories. Pearson’s product-moment correlation for “proton” with “BK” was R = 0.901 (p = 0.00283) across 7 categories. R2 for “proton” with “BK” was 0.81. R2 for “proton” with “potassium” was 0.74. Results among the ions for correlation, with numbers of “Met” citations included, differed markedly in “cancer” searches compared to search results not restricted to “cancer” for unknown reasons. (D) Numbers of citations shared among categories of 7 oncogenes and histone are shown. E1–E70 citations are listed in Section 3.3.6.
Figure 4. Oncogenes and DNA modifiers may play a role in cancer studies via KCN DEGs encoding H+-sensitive proteins. (A) KCN DEGs that encode H+-sensitive proteins are highlighted with red underlining of study names. Histograms of 8 categories for cancer studies among E1–E70 are shown on the left (yellow boxes). The right histogram also includes 20 studies in “no category” (white boxes). Comparison of studies in “no category” versus those in at least one category shown in the left histogram found a significant difference in the number of studies detecting KCN DEGs encoding H+-sensitive proteins using Fisher’s exact test, p = 0.0325, indicating a higher likelihood for studies included in at least one category. (BD) Word searches of PubMed for numbers of oncogene and histone publications related to H+, K+, BK(KCNMA1), and Cl (contrasting ion) were performed to further evaluate categories in E1–E70. (B) On 12 May 2025, searches for each category of “TP53”, “MYC”, “TGFbeta”, “BCL2”, “EGFR”, “histone”, “RAS”, and “MET” were performed separately for “proton”, “potassium”, and “chloride”. For all 8 categories, the numbers of citations for “proton”, “potassium”, and “chloride” were 1565, 1418, and 1261, respectively, and Pearson’s product-moment correlations for “proton” with “potassium” and “proton” with “chloride” were R = 0.891 (p = 0.00148) and 0.599 (p = 0.0584), respectively. Pearson’s product-moment correlation for “potassium” and “chloride” was 0.689 (p = 0.0295). R2 for “proton” with “potassium” was 0.79. “BK” and “potassium” searches on 7 June 2025revealed “BK” as a potential large component of “potassium” citations, approximately one-third for 7 categories. (C) When searches also included “cancer” as a third word, the numbers of citations for “proton”, “potassium”, “BK”, and “chloride” on 14 May 2025, were 1061, 577, 513 (88.9% of “potassium” search results), and 952, respectively. When MET was left out for better “proton” correlations, the numbers of citations for “proton”, “potassium”, “BK”, and “chloride” on 14 May 2025, were 728, 496, 439 (88.5% of “potassium” search results), and 879, respectively, and Pearson’s product-moment correlations for “proton” with “potassium” and “chloride” were R = 0.860 (p = 0.00649) and 0.462 (p = 0.148), respectively, across the 7 remaining categories. Pearson’s product-moment correlation for “proton” with “BK” was R = 0.901 (p = 0.00283) across 7 categories. R2 for “proton” with “BK” was 0.81. R2 for “proton” with “potassium” was 0.74. Results among the ions for correlation, with numbers of “Met” citations included, differed markedly in “cancer” searches compared to search results not restricted to “cancer” for unknown reasons. (D) Numbers of citations shared among categories of 7 oncogenes and histone are shown. E1–E70 citations are listed in Section 3.3.6.
Biomolecules 15 01177 g004
Figure 5. Repeat non-KCN DEGs were often found with KCNN4 and some with KCNMA1. (A) Among non-KCN DEGs, ITGA2/Integrin Alpha-2 was most frequent (7 studies). Six also included H+-sensitive coding KCN DEGs (bolded): E8 (KCNK1, -N4), E15, E39 (KCNMA1), E41 (KCNMA1, -N4), E49 (KCNN4), E64 (KCNN4), and E69 (KCNN4). KCNN4 was the most frequent, 9 times in total and was also in 5 of 7 studies with ITGA2. Studies of co-expressed KCNN4, ITGA2, and PLAUR, E8 [172], E41 [162], E49 [166], and E69 [198], included 2 lung cancers, 1 melanoma, and 1 pancreatic cancer, in categories of His/Chr/Epi, TP53, and RAS (top left Venn diagram). E49 detected KCNN4, LIF, and PLAU together (top right Venn diagram). E41 [162] detected KCNMA1, KCNN4, ITGA2, and PLAUR (not shown). (B) PubMed word searches for shared citations among “BK” (KCNMA1) and “PLAUR” + “PLAU” and “ITGA2” found correlations, R2 = 0.55, p = 0.0175 and R2 = 0.52, p = 0.0224, respectively, across 8 categories. # = number.
Figure 5. Repeat non-KCN DEGs were often found with KCNN4 and some with KCNMA1. (A) Among non-KCN DEGs, ITGA2/Integrin Alpha-2 was most frequent (7 studies). Six also included H+-sensitive coding KCN DEGs (bolded): E8 (KCNK1, -N4), E15, E39 (KCNMA1), E41 (KCNMA1, -N4), E49 (KCNN4), E64 (KCNN4), and E69 (KCNN4). KCNN4 was the most frequent, 9 times in total and was also in 5 of 7 studies with ITGA2. Studies of co-expressed KCNN4, ITGA2, and PLAUR, E8 [172], E41 [162], E49 [166], and E69 [198], included 2 lung cancers, 1 melanoma, and 1 pancreatic cancer, in categories of His/Chr/Epi, TP53, and RAS (top left Venn diagram). E49 detected KCNN4, LIF, and PLAU together (top right Venn diagram). E41 [162] detected KCNMA1, KCNN4, ITGA2, and PLAUR (not shown). (B) PubMed word searches for shared citations among “BK” (KCNMA1) and “PLAUR” + “PLAU” and “ITGA2” found correlations, R2 = 0.55, p = 0.0175 and R2 = 0.52, p = 0.0224, respectively, across 8 categories. # = number.
Biomolecules 15 01177 g005
Figure 6. Results for oligodendrogliomas (Oligos) and glioblastomas (GBMs) from REMBRANDT. (AD) Median reporter probe values of microarray gene expression results from Oligo and GBM tumor tissue are shown. Pink backgrounds (AF) indicate results are from patients on anti-seizure medications (Anti-seizure Rx). Results with a blue background (C, E, F) are from large-group patients (medication history not used). (A) Comparisons of ATP1A1/Na/K-ATPase subunit alpha 1 (neural-specific) versus ATP1A2/Na/K-ATPase subunit alpha 2 (glial-specific) in Oligos and GBMs with significant differences confirmed glial origins. (B) Median reporter probe values for astrocytic-specific (versus oligodendroglial) values for gap junction genes with higher levels of expression in GBMs versus Oligos. (C) Values for median reporter probes for 4 KCN genes, CA12/carbonic anhydrase 12, and TP53/p53. (D) Another method to evaluate gene expression calculated Oligo-minus-GBM (OMG) differences of median probe results for each gene of interest and the housekeeping genes (HKGs). This method provided correction for variation in tissue handling, assay-to-assay variation within and between institutions, etc. (EF) The p values are graphed for decreased and increased gene expressions in GBMs. Significance improved for the KCNJ16 difference and trends strengthened for other genes. Oligo = oligodendroglioma, GBM = glioblastoma, Rx = medication, HKG = housekeeping gene. REMBRANDT (part of the Georgetown Database of Cancer (G-DOC)). * p < 0.05.
Figure 6. Results for oligodendrogliomas (Oligos) and glioblastomas (GBMs) from REMBRANDT. (AD) Median reporter probe values of microarray gene expression results from Oligo and GBM tumor tissue are shown. Pink backgrounds (AF) indicate results are from patients on anti-seizure medications (Anti-seizure Rx). Results with a blue background (C, E, F) are from large-group patients (medication history not used). (A) Comparisons of ATP1A1/Na/K-ATPase subunit alpha 1 (neural-specific) versus ATP1A2/Na/K-ATPase subunit alpha 2 (glial-specific) in Oligos and GBMs with significant differences confirmed glial origins. (B) Median reporter probe values for astrocytic-specific (versus oligodendroglial) values for gap junction genes with higher levels of expression in GBMs versus Oligos. (C) Values for median reporter probes for 4 KCN genes, CA12/carbonic anhydrase 12, and TP53/p53. (D) Another method to evaluate gene expression calculated Oligo-minus-GBM (OMG) differences of median probe results for each gene of interest and the housekeeping genes (HKGs). This method provided correction for variation in tissue handling, assay-to-assay variation within and between institutions, etc. (EF) The p values are graphed for decreased and increased gene expressions in GBMs. Significance improved for the KCNJ16 difference and trends strengthened for other genes. Oligo = oligodendroglioma, GBM = glioblastoma, Rx = medication, HKG = housekeeping gene. REMBRANDT (part of the Georgetown Database of Cancer (G-DOC)). * p < 0.05.
Biomolecules 15 01177 g006
Figure 7. Many KCN DEGs detected in E1–E70 studies (Figure 3, Figure 4 and Figure 5) and in REMBRANDT (Figure 6) were included in whole-genome and expression studies of 6 glioblastoma cell lines, LN18, LN229, T98G, U87, U343, and U373, by Patil et al. [22]. For comparisons, shared KCN DEGs in E1–E70 (29 cancers including 7 glioblastomas) are marked with green squares and KCN genes of interest found in REMBRANDT are bolded (A,B). (A) Sequencing revealed DNA alterations including cancer-specific mutations (CSMs), insertions (INSs), and deletions (DELs). A total of 23 KCN DEGs (of 53 in E1–E70; see Figure 3 and Figure 4) and those of interest in REMBRANDT (3 with trends and 1 significantly differentially expressed; see Figure 6C–F) had changes in genomes of 6 cell lines reported by Patil et al. Several shared KCN DEGs were genetically altered in all 6 cell lines. (B) Average RNA expressions determined by Patil et al. are listed in descending order of abundance, including 27 and 4 of E1–E70 KCN DEGs and REMBRANDT’s KCN DEGs and genes of interest, respectively, with names of proteins. (C) Log2 fold RNA changes in combined tumor cell lines versus normal (see text) are shown with KCNs encoding H+-sensitive proteins indicated with red (H+) and KCNs with DNA changes in (A) with underlying blue brackets (˽). Of 51 KCNs in the glioblastoma cell lines, 27 (52.9%) encoded H+-sensitive proteins. Of 53 KCNs in E1–E70, 38 (71.7%) were found in their glioblastoma cell lines, with altered DNA and/or RNA expression. Of seven GBM studies in E1–E70, six (85.7%) shared 1–3 KCN DEGs with those described by Patil et al. [22]. See the text and Appendix B (Figure A5A) for specific glioblastoma studies in E1–E70. References for studies are in Section 3.3.6.
Figure 7. Many KCN DEGs detected in E1–E70 studies (Figure 3, Figure 4 and Figure 5) and in REMBRANDT (Figure 6) were included in whole-genome and expression studies of 6 glioblastoma cell lines, LN18, LN229, T98G, U87, U343, and U373, by Patil et al. [22]. For comparisons, shared KCN DEGs in E1–E70 (29 cancers including 7 glioblastomas) are marked with green squares and KCN genes of interest found in REMBRANDT are bolded (A,B). (A) Sequencing revealed DNA alterations including cancer-specific mutations (CSMs), insertions (INSs), and deletions (DELs). A total of 23 KCN DEGs (of 53 in E1–E70; see Figure 3 and Figure 4) and those of interest in REMBRANDT (3 with trends and 1 significantly differentially expressed; see Figure 6C–F) had changes in genomes of 6 cell lines reported by Patil et al. Several shared KCN DEGs were genetically altered in all 6 cell lines. (B) Average RNA expressions determined by Patil et al. are listed in descending order of abundance, including 27 and 4 of E1–E70 KCN DEGs and REMBRANDT’s KCN DEGs and genes of interest, respectively, with names of proteins. (C) Log2 fold RNA changes in combined tumor cell lines versus normal (see text) are shown with KCNs encoding H+-sensitive proteins indicated with red (H+) and KCNs with DNA changes in (A) with underlying blue brackets (˽). Of 51 KCNs in the glioblastoma cell lines, 27 (52.9%) encoded H+-sensitive proteins. Of 53 KCNs in E1–E70, 38 (71.7%) were found in their glioblastoma cell lines, with altered DNA and/or RNA expression. Of seven GBM studies in E1–E70, six (85.7%) shared 1–3 KCN DEGs with those described by Patil et al. [22]. See the text and Appendix B (Figure A5A) for specific glioblastoma studies in E1–E70. References for studies are in Section 3.3.6.
Biomolecules 15 01177 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Beckner, M.E. Cell Settling, Migration, and Stochastic Cancer Gene Expression Suggest Potassium Membrane Flux May Initiate pH Reversal. Biomolecules 2025, 15, 1177. https://doi.org/10.3390/biom15081177

AMA Style

Beckner ME. Cell Settling, Migration, and Stochastic Cancer Gene Expression Suggest Potassium Membrane Flux May Initiate pH Reversal. Biomolecules. 2025; 15(8):1177. https://doi.org/10.3390/biom15081177

Chicago/Turabian Style

Beckner, Marie E. 2025. "Cell Settling, Migration, and Stochastic Cancer Gene Expression Suggest Potassium Membrane Flux May Initiate pH Reversal" Biomolecules 15, no. 8: 1177. https://doi.org/10.3390/biom15081177

APA Style

Beckner, M. E. (2025). Cell Settling, Migration, and Stochastic Cancer Gene Expression Suggest Potassium Membrane Flux May Initiate pH Reversal. Biomolecules, 15(8), 1177. https://doi.org/10.3390/biom15081177

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop