GET THE APP

Multifaceted Roles of Carbon Anhydrase IX in Cancer Cell Proliferation, Survival, Metastasis and Therapy Resistance and Indication of Promising Novel Therapies by its Inhibitors
..

Molecular and Genetic Medicine

ISSN: 1747-0862

Open Access

Review Article - (2022) Volume 16, Issue 11

Multifaceted Roles of Carbon Anhydrase IX in Cancer Cell Proliferation, Survival, Metastasis and Therapy Resistance and Indication of Promising Novel Therapies by its Inhibitors

Mitsuru Sakitani*
*Correspondence: Mitsuru Sakitani, Director of the Institute CCC, Kobe, Hyogo, Japan, Email:
Director of the Institute CCC, Kobe, Hyogo, Japan

Received: 04-Nov-2022, Manuscript No. jmgm-22-79048; Editor assigned: 05-Nov-2022, Pre QC No. P-79048; Reviewed: 08-Nov-2022, QC No. Q-79048; Revised: 16-Nov-2022, Manuscript No. R-79048; Published: 23-Nov-2022 , DOI: 10.37421/1747-0862.2022.16.582
Citation: Sakitani, Mitsuru. “Multifaceted Roles of Carbon Anhydrase IX in Cancer Cell Proliferation, Survival, Metastasis and Therapy Resistance and Indication of Promising Novel Therapies by its Inhibitors.” J Mol Genet Med 16 (2022): 582.
Copyright: © 2022 Sakitani M. This is 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.

Abstract

Cancer cells of solid tumors suffer from decrease in blood supply and hypoxia. To compensate deterioration in the intra-tumorous environment, hypoxia inducible factors, HIF-1α, HIF-2α and HIF-3α, are induced and numerous target genes are activated by these HIFs. Of these, CA9 is highly important because of its involvement in oncogenesis of solid tumors and malignant lymphomas. The essential function of CA9 in cancer is regulation of intracellular phi and extracellular pHe by mitigating the decreased pH due to excessed glycolysis in cancer cell. In this review article, first, CA9’s multifaceted roles in cancer cell proliferation, survival and metastasis and therapy resistance were summarized in association with effects of CA9 inhibitors. Second, after clarifying the CA9 activation via HIF-1α in the KRAS/RAF/MEK/ERK and PI3E/AKT/mTOR signaling pathways, molecular targeted therapies by means of CA9 inhibitors against therapy resistant triple negative breast cancer (TNBC), pancreatic ductal adenocarcinoma (PDAC), intraductal papillary mucinous neoplasms of the pancreas (IPMN) and adult T-cell leukemia/lymphoma (ATL) were shown to be promising novel therapies.

Keywords

CA9 • HIF • KRAS • PI3K/AKT/mTOR • TNBC • PDAC • IPMN • ATL

Introduction

Cancer cells of solid tumors suffer from decrease in blood supply and hypoxia [1,2] and require adequate blood perfusion to obtain nutrients and oxygen for proliferation and survival. To compensate deterioration in the intratumorous environment, hypoxia inducible factors (HIFs), HIF-1α, HIF-2α and HIF-3α, are induced [3,4]. Under normoxia, HIF-1α is subjected to ubiquitinproteasomal degradation by von hippel-lindau tumor suppressor protein (pVHL) [5,6] or tumor suppressor protein p53 [7,8], whereas factor inhibiting HIF (FIH) inhibits HIF’s recruitment of co-activator CBP/p300 [9,10], leading to inhibition of its functions. However, under hypoxia, expression of HIF-1α is activated and HIF-1α is accumulated in cytoplasm. Then, HIF-1α is translocated to nucleus via the nuclear trans localization signals [11] and forms a heterodimer with a constitutively expressed HIF-1β (known as aryl hydrocarbon receptor nuclear translocator [ARNT]) [12,13]. The HIF-1α/HIF-1β complex directly binds to hypoxia response elements (HREs) of target genes [14] with recruitment of CREB-binding protein (CBP)/p300 transcriptional co-activator [11] and activates numerous target genes involved in energy metabolism (glucose transporters [GLUTs] 1/2/3, hexokinases 1/2, phosphofructokinase L, aldolase A, pyruvate kinase M, phosphorglycerate kinase 1, enolase 1 and lactate dehydrogenase A), metabolic adaptation/pH regulation (carbonic anhydrase IX [CA9] and CA12), cell proliferation and survival (insulin-like growth factor 2 [IGF2], IGF binding proteins 1/2/3, transforming growth factors [TGFs] α/β, epidermal growth factor [EGF] and c-myc), angiogenesis (vascular endothelial growth factor [VEGF], VEGF receptor-1, leptin, LDL-receptor-related protein 1 and adrenomedullin), erythropoiesis (erythropoietin, transferrin, transferrin receptor and ceruloplasmin) and vasomotor control (nitric oxide synthase, adrenomedullin, α1B-adrenergic receptor and endothelia-1) [1,15-17].

Of these target genes, CA9 is highly important because the involvement of CA9 in oncogenesis [18,19] and therapy resistance [20-23] has been indicated and CA9 has been recognized to be an important target for cancer therapies [24-26]. CAs are metalloenzymes that containing a zinc, catalyze the reversible hydration of carbon dioxide to bicarbonate and proton: CO2+ H2O ⟷ HCO3 - + H+. CAs are evolutionally grouped into six classes, α, β, γ, δ, ζ and η [27]. The human CAs belong to the α-class and consist of catalytically active 12 isoforms (CAs I–IV, CAs VA–VB, CAs VI–VII, CA IX and CAs XII–XIV) [28] while other 3 isoforms CA VIII, X and XI lack the metal ion within the active site. The essential function of CA9 in cancer is regulation of intracellular pHi and extracellular pHe by mitigating the decreased pH due to excessed glycolysis in cancer cells [19,28], leading to cancer progression [24-26,29]. In addition, molecular targeted therapies (MTTs) against CA9 by means of CA9 inhibitors are suggested to be effective against broad solid tumors and malignant lymphomas [2,24,26,30-33].

In this review, we show the current status about CA9 research on its multifaceted roles in cancer cell proliferation, survival, metastasis and therapy resistance and we indicate that molecular targeted therapies by means of CA9 inhibitors against therapy resistant triple negative breast cancer (TNBC), pancreatic ductal adenocarcinoma (PDAC), intraductal papillary mucinous neoplasms of the pancreas (IPMN) and adult T-cell leukemia/lymphoma (ATL) are promising novel therapies.

Overview of the regulation and functions of CA9

Under hypoxia in solid tumors, CA9 is induced by HIFs, in particular HIF-1α. In the promoter regions of CA9, there are several cis-acting elements [34,35]. Of these, HRE is the most potent activating element of CA9 transcription by binding of the HIF-1α/HIF-1β complex with recruitment of CBP/p300 [5,35,36]. In addition, there are other enhancer elements of binding sites for specificity protein (SP)-1 and SP-3 [35,37] and activator protein-1 (AP-1) [34,37] in (Figure 1). Activation of HIF-1α is also induced by various mechanisms [1,2,38]. Proteasomal degradation of HIF-1α and HIF-2α is inhibited by inactivation of pVHL [5,6] and p53 [7,8], leading to accumulation of the HIF-1α and HIF-2α protein, while inactivation of the factor inhibiting HIF (FIH) that inhibits recruitment of coactivator CBP/p300 [9,10] recovers the recruitment, resulting in activation of HIF-1α and HIF-2α. Transcription of the HIF-α mRNA is activated by components of the nuclear factor kappa B (NF-κB) signaling pathway such as 50-RelA [39-41], IKKβ [42] or TRAF6 [43]. In addition, translation of the HIF-1α mRNA is activated by the phosphatidylinositol 3-kinase (PI3K)/AKT/mechanistic target of rapamycin (mTOR) signaling pathway [17,44] including loss of inhibitory phosphatase and tensin homolog (PTEN) [45] and the KRAS/mitogen-activated protein kinase kinase (MEK)/extracellular signal-regulated kinase (ERK) signaling pathway [17,46-48]. Activation of CBP/p300 by ERK also activates the function of HIF- 1α [49]. In solid tumors under hypoxia, CA9 regulates intracellular (pHi) and extracellular (pHe) acidosis in the tumor microenvironment and promotes cancer progression by inducing cell proliferation [18,19], cell survival by apoptosis inhibition [50,51], cell migration, invasion and metastasis [26,29,52], therapy resistance [21,53,54] and tumorigenicity [55] as in (Figure 2).

molecular-genetic-medicine-signaling-effectors

Figure 1. Overview of CA9 regulation: Under normoxia, functions of HIF-1α are inhibited by pVHL, p53 or FIH, while activation of HIF-1α induced under hypoxia or by various factors such as the NF-κB, PI3K/AKT/mTOR and KRA/RAF/MER/ERK signaling effectors. Then, HIF-1α in cytoplasm is translocated to nucleus and forms a heterodimer with a constitutively expressed HIF-1β. The HIF-1α/HIF-1β complex directly binds to hypoxia HRE of CA9 with recruitment of CBP/p300, leading to activated transcription of CA9 mRNA. In addition, SP-1, SP-3 and AP-1 enhances transcription of CA9 mRNA. Abbreviations: CA9, carbon anhydrase IX; FIH, factor inhibiting HIF; HIF, hypoxia inducible factor; HRE, hypoxia response element; mTOR, mechanistic target of rapamycin; NF-κB, nuclear factor kappa B; PI3K, phosphatidylinositol 3-kinase; pVHL, von Hippel-Lindau tumor suppressor protein.

molecular-genetic-medicine-multifaceted-roles

Figure 2. Multifaceted roles of CA9 in cancer progression of solid tumors: CA9 regulates intracellular pHi and extracellular pHe and promotes cancer progression by inducing cell proliferation, cell survival by apoptosis inhibition, cell migration/invasion/ metastasis, therapy resistance, and tumorigenicity.

Cell proliferation

Activated expression of a membrane-localized CA9 induces alkalization of intracellular pHi and acidification of extracellular pHe [19,28]. Compensation of the pHi and pHe by CA9 drives cancer cell proliferation, survival and migration [18,19]. In fact, CA9 expression is recognized as a biomarker of poor prognosis in solid tumors [25,56]. There are numerous reports on the CA9 expression that is correlated with severity of malignancy and poor prognosis in solid tumors and lymphomas such as breast cancer [21-23,54,57,58], PDAC [59-61], renal cell carcinoma (RCC) [62,63], non-small cell lung cancer (NSCLC) [64,65], colorectal cancer (CRC) [66], hepatocellular carcinoma (HCC) [67], prostate cancer [20], cervical cancer [68], brain tumor [69], thymic cancer [70], classical Hodgkin lymphoma [71,72], B-cell lymphoma [2,73,74], T-cell lymphoma [75] and ATL [55].

Cell survival by apoptosis inhibition

In many cancers, apoptosis is suppressed and this condition is favorable for cancer cell survival [50,51]. Apoptotic signaling pathways essentially consist of the extrinsic death receptor-dependent pathway and the intrinsic mitochondrial apoptosis pathway [76-78]. The extrinsic death receptor-dependent pathway is activated by binding of ligands such as tumor necrosis factor (TNF), Fas ligand (FasL) or TNF-related apoptosis-inducing ligand (TRAIL) to their corresponding death receptors, i.e., type 1 TNF receptor (TNFR1), Fas (also known as CD95 or Apo-1), TRAIL receptor 1 (TRAILR1 or DR4) or TRAIL receptor 2 (TRAILR2 or DR5) [79]. The activated ligand-bound death receptors (TNFα-TNFR1, FASL-FAS, TRAIL-TRAILR1 or TRAIL-TRAILR2) that contain intracellular death domain (DD) [80] recruit and interact with adaptor proteins Fas-associated death domain (FADD) and TNF receptor-associated death domain (TRADD) via DD [79].

In addition to DD, these adaptor proteins FADD and TRADD have also death effector domain (DED) [81] and interact with the initiators pro-caspase-8 and pro-caspase-10 that contain DED. The death receptors, FADD/TADD and pro-caspase-8/-10 thus form death-inducing signaling complex (DISC) [82], which promote auto-activation of pro-caspase-8 and procaspase-10, resulting in caspase-8 and caspase-10 that in turn activate the effectors caspase-3, -6 and -7 [83], leading to apoptosis. The intrinsic mitochondrial apoptosis pathway [84,85] is activated by various external stresses (irradiation, chemotherapy or others) and internal stimuli (irreparable genetic damage, hypoxia, cytosolic Ca2+ elevation or oxidative stress) [78,85] and numerous proteins of the B-cell lymphoma 2 protein (Bcl-2) family are activated. Proteins of the Bcl-2 family contain one to four different Bcl-2 homology domains (BH1, BH2, BH3 and BH4) and anti-apoptotic proteins have all the four BH domains (Bcl-2, Bcl-xL, A1, Mcl-1 and Bcl-w), while pro-apoptotic proteins contain either multidomains BH1, BH2 and BH3 (Bax, Bak and Bok) or only BH3 (Bid, Bad, Bim, Bmf, Bik, Hir/DP5, Blk, Nip3, BNip3/Nix, Puma and Noxa) [78,86].

When the intrinsic mitochondrial apoptosis pathway is stimulated, activation of pro-apoptotic proteins is induced and the activated pro-apoptotic proteins suppress anti-apoptotic proteins, resulting in disruption of mitochondrial membrane outer membrane permeabilization (MOMP) [87] that promotes release of cytochrome c from the mitochondrial inter membrane space into cytosol. Then, cytochrome c interacts with apoptosis protease-activating factor 1 (Apaf- 1) and forms a complex known as apoptosome [88]. The apoptosome recruits initiator pro-caspase-9 and promotes its autoactivation into caspase-9 that in turn activates executors caspase-3, -6 and -7 and finally apoptosis of cell is induced [85].

In addition, tumor suppressor p53 is involved in both the extrinsic and intrinsic apoptosis pathways [78,89,90], whereas there are several inhibitors of apoptosis proteins (IAPs) that contain one to three baculoviral IAP repeat (BIR) domains and are defined as BIR-domain-containing proteins (BIRPs) [91,92]. In solid tumors, apoptosis is suppressed by various mechanisms such as increased expression of anti-apoptotic proteins, decreased expression of pro-apoptotic proteins, increased expression of IAPs, decreased expression of caspases, impaired death receptor signaling pathway or defect/mutation in p53 [77,78,93]. However, CA9 inhibitors abrogate apoptosis inhibition of cancer cells. For instance, the CA9 inhibitor sulphonamide induces decreased intracellular pHi, increased intracellular free radical (ROS), reduced mitochondrial membrane potential that increases the production of free radicals, increased expression of PARP (an indicator of DNA damage), cell cycle arrest, increased expression of the mRNA of caspase-3, -8 and -9, increased expression of pro-apoptotic proteins Bax and others, triggering of the apoptotic morphology and reduced viability and proliferation of cells [94]. Taken together, CA9 inhibitor reverses the apoptosis inhibition in cancer cells and induces reduction of cell proliferation. The similar results were obtained by other studies [68,95,96].

Cell migration, invasion and metastasis

CA9 plays also an important role in cancer cell migration and invasion in association with integrins and metalloproteases (MMPs) [26,29,97]. Various parts of CA9, including the proteoglycan-(PG)-like domain, intracellular domain and catalytic domains, are involved in the processes of cell adhesion and invasion of cancer cells [29,97]. CA9 is associated with the collagen and laminin receptors integrin β1, integrin α2, integrin α3, integrin α5 and integrin α6 [97]. These integrins promote focal adhesions that are early steps of cell mobility and migration toward metastasis. In addition, a matrix protease MMP14 in association with CA9 participates in matrix degradation and invadopodia formation [97,98], leading to cell invasion and metastasis. Proton (H+) excretion by CA9 and consequent extracellular acidification (pHe) are prerequisite for the MMP14 activity [29,97].

A pH regulatory protein Na+/H+ exchanger 1 (NHE1) also contributes to the pHe acidification to promote the MMP14-related matrix degradation [99,100]. CA9 inhibitors SLC-0111 and AA-06-05 impair cancer cell migration and invasion in the in vitro experiments using a breast cancer cell line (MDAMB- 231) and a lung cancer cell line (A549) [101]. The same results have been obtained by other studies by means of CA9 inhibitors sulfonamide (CAI17) [102] or VD11-4-2 [103] in the in vitro systems with breast cancer cell lines 4T1, 66cl4, 67NR, MDA-MB-231 [102] or MCF-7, MDA-MB-231 [103]. Suppression of cell invasion and metastasis by CA9 inhibition is also confirmed by the in vivo animal models [98,102,104]. Accordingly, CA9 depletion by shRNAs also induces inhibition of cancer cell migration and invasion [52,102,105].

Therapy resistance by phenotypic plasticity (CSC and EMT)

There are various mechanisms of therapy resistance [67,106,107]. Primary resistance is derived from existence of the clones that have already resistant mutations before treatment of MTT [108,109], while acquired resistant is induced by different mechanisms such as secondary mutations of the target molecule [110,111], activation of downstream signaling pathways [112,113], or activation of bypass signaling pathways [114,115]. Intra-tumoral heterogeneity (ITH) [116,117] due to coexistence of cancer stem cell (CSC) [118,119] or clones harboring resistant mutations [120,121], which are accelerated by epigenetic changes [122,123], is another cause of therapy resistance. Accordingly, importance of phenotypic plasticity in resistance has been also recognized [124,125]. Phenotypic plasticity is induced by dedifferentiation (reversion from differentiated cell to stem cell) or transdifferentiation (direct shift from one differentiated cell type to another) [126].

Therapy resistance by dedifferentiation (reversion to CSCs) [124,127] is observed in TNBC [128,129], acute myeloid leukemia [130] and neuroblastoma [131].

CSCs possess stem cell features, i.e., self-renewal and differentiation potential [132] and stemness is induced by various signaling pathways and transcription factors such as Wnt signaling, Notch signaling, Hb signaling, PI3K/AKT/mTOR signaling, NF-κB signaling, JAK/STAT signaling, SOX2, NANOG and HIF [126,127,129]. Fortunately, there are several reports [21-23,58] that therapy resistance by CSCs in TNBC were abrogated by small CA9 inhibitors (CAI017, U104, RC44 or acetazolamide) or CA9-specific siRNAs. These studies show potential of CA9 inhibitors against therapy resistant TNBC. However, exact mechanisms of these anti-CA9 treatments should be further clarified. Therapy resistance by transdifferentiation is exemplified by epithelical-mesenchymal transition (EMT) [125,133,134] and EMT is found in NSCLC [135-137], prostate cancer [138], hepatocellular carcinoma [139] or gastric cancer [140]. Transcription factors such as Zeb1, Twist1, Snail1 and Slug are involved in EMT [126,129,133]. Fiaschi [20] proved the inhibition of EMT and invasiveness of prostate cancer cell lines by means of genetic silencing of CA9 or pharmacological inhibition of CA9 with sulfonamides sulfamides inhibitors. These results are indicative of future therapeutic possibility of anti-CA9 treatment against NSCLC, HCC or gastric cancer.

In addition, trial reports on the CA9 inhibitors that enhance the sensitivity of combination therapy with conventional cytotoxic chemotherapy [21,141] or radiation therapy [22] also demonstrate a future therapeutic strategy of CA9 inhibitors.

The KRAS/RAF/MEK/ERK signaling pathway and bypass signaling

KRAS mutation is broadly observed in various cancers such as PDAC (67.61%), CRC (35.77%), NSCLC (20.42%) or others [142] and it causes activation of the downstream signaling effectors, leading to cancer progression. Concerning gene mutations of PDAC, KRAS mutation is observed in 93%, while other mutations of p53 (72%), CDKN2A (30%) and SMAD4 (32%) are also found [143]. Of these, KRAS mutation is the most important [61,144,145] because it is required for initiation and maintenance of PDAC [144,146]. The KRAS signaling pathways are further ramified into the RAF/MEK/ERK [147-150], the PI3K/AKT/mTOR [151,152], the RALGDS/RAL [153,154] and other signaling pathways (TIAM1/RAC, PLCε and others) [142,155]. The RAF/MEK/ ERK pathway is involved in oncogenesis of PDAC via CA9 activation [61], while the PI3K/AKT/mTOR pathway also activates CA9 via HIF [17,44], but the RALGDS/RAL pathway may not be associated with CA9 activation [153,154]. Interaction of the PI3K/AKT/mTOR with CA9 will be later discussed.

In addition to the KRAS/RAF/MEK1,2/ERK1,2 signaling pathway, there are several ERK signaling pathways that are not mediated by KRAS, i.e., the MEK4/JNK, the MEK5/ERK5 and the MEK6/p38 [156,157]. However, the KRAS/RAF/MEK1,2/ERK1,2 signaling pathway plays the central role in activation of CA9 via HIF-1α by mediating eukaryotic initiation factors (eIFs) [158,159]. In fact, the downstream effectors ERK1 and ERK2 stimulate translation of HIF-1α mRNA by activating several eIFs [17,46-48], resulting in CA9 activation.

In the KRAS-mutant tumors, exemplified by PDAC [61,144,145], the KRAS/RAF/MEK/ERK signaling effectors are thought to promote oncogenic progression. Thus, numerous inhibitors against the KRAS mutants and its downstream effectors have been developed [141,150,155,160,161]. However, therapeutic results by the inhibitors against KRAS [162,163], RAF [164] or MEK [165,166] for PDAC have not been satisfactory [61,162,163]. Certainly therapy resistant mutation of KRAS such as G12C is a cause of these unmet results [142,167], but activation of bypass signaling is a more serious problem in targeting the KRAS/RAF/MEK/ERK signaling pathway [145,161]. Of various bypass signaling pathways, the PI3K/AKT/mTOR pathway is important. Thus combination therapy of the KRAS/RAF/MEK/ERK and PI3K/AKT/mTOR inhibitors was pursued [168,169]. However, an early clinical trial of dualpathway inhibition revealed moderate efficacy and toxicity significantly more severe than single inhibition regimen [170], resulting in stagnation of the dualinhibition strategy.

In contrast, CA9 inhibition with SLC-0111 [59-61] or FC12-531A [59,60] induces reduction of PDAC cell growth in several models, including the in vitro cell line models [59-61], the in vivo animal models, the human cell-line derived xenograft models and the human patient-derived xenograft (PDX) models [61]. This PDAC cell killing is enhanced by HIF-1α inhibition [59,60], indicating that CA9 activation is mediated by HIF-1α activation. In addition, HIF-1α stabilization is induced by translation activation of HIF-1α mRNA in the KRAS/ MEK/ERK-activated PDAC cell lines [61], while transcription activation of HIF- 1α mRNA may not play a major role in CA9 activation in the KRAS mutant PDAC cells [61]. Certainly single administration of CA9 inhibitor is effective against PDAC cell growth inhibition, but combination therapy with HIF-1α inhibitor [59,60] or conventional cytotoxic chemotherapy using gemcitabine [61] shows more potent efficacy than the CA9 inhibitor single regimen. Efficacy and safety of CA9 inhibitor SLC-0111 has been scrutinized under clinical trials [59]. Since CA9 is activated by both the KRAS/RAF/MEK/ERK and PI3K/ AKT/mTOR signaling pathways, this CA9 inhibition strategy is expected to be effective also against bypass activation of the PI3K/AKT/mTOR signaling in KRAS mutant PDAC.

Furthermore, KRAS mutation is also frequently found in IPMN [171-174]. The frequency of KRAS mutation in IPMN ranges from 38% [175] through 50% [176] or 61% [177] to 81% [178]. Among other gene mutations in IPMN [176,179], RAF mutation (2.7%) [180], PI3K mutation (11%) [181], AKT mutation (8.3%) [182] and PTEN loss (36%) [182] are relevant to CA9 activation, because the KRAS and RAF mutations induce activation of the KRAS/RAF/MEK/ERK signaling pathway while the PI3K mutation, AKT mutation and PTEN loss are associated with activation of the PI3K/AKT/mTOR signaling pathway, both leading to CA9 activation via HIFs. In this regard, just as in PDAC, CA9 inhibitors may pave a path to novel molecular targeted therapies against IPMN.

The PI3K/AKT/mTOR signaling pathway and tumorigenicity

Since CA9 inhibitor effectively reduces cell growth of KRAS mutant PDAC, CA9 inhibitor strategy is expected to be also applicable to the solid tumors and lymphomas that are activated by the PI3K/AKT/mTOR signaling pathway via HIFs and CA9. Phosphorylation of phosphoinositides (PtdIns) is a first step in activation of the PI3K/AKT/mTOR signaling pathway [183-185]. PtdIns(4,5)P2 binds to the pleckstrin homology (PH) domain of the agonist-activated PI3K in the membrane. Then, PI3K phosphorylates PtdIns(4,5)P2 and generates PtdIns(3,4,5)P3. In turn, PtdIns(3,4,5) is produced into PtdIns(3,4)P2 by Srchomology 2 containing inositol polyphosphate 5-phosphatase 1 (SHIP1). Then, the AKT-activating PtdIns(3,4)P2 is converted into inactive PtdIns(3) P by inositol polyphosphate-4-phosphatase (INPP4) [186]. PTEN reverses PI3K function by dephosphorylating PtdIns(3,4,5)P3 into PtdIns(4,5)P2 [183,185,186-188].

When PI3K is activated by its agonists, AKT in cytoplasm is recruited to the membrane and is bound via its PH domain to PtdIns(3,4,5)P3 and PtdIns(4,5)P2 that are produced by agonist-stimulated PI3K and SHIP1, respectively. This binding of AKT to PtdIns(3,4,5)P3 and PtdIns(4,5)P2 induces conformational change of AKT for consequent phosphorylation of Thr308 and Ser473 [189,190]. Then, Thr308 is phosphorylated by phosphoinositidedependent kinase 1 (PDK1) [191,192], whereas Ser473 is phosphorylated by mTORC2 [193]. These phosphorylations activate AKT and activated AKT is then translocated through cytoplasm into nucleus and phosphorylates various substrates including mTORC1 [183,194]. mTORC1 then activates HIF-1α under normoxia [44,195,196] by stimulating translation of HIF-1α mRNA [17]. Translation of eukaryotic mRNA consists of four phases, i.e., initiation, elongation, termination and ribosome recycling [158,159].

Initiation is the rate-limiting phase [197] and starts from formation of preinitiation 43S complex consisting of 40S ribosome subunit, termary complex (of eIF2, GTP and Met-tRNAi), eIF1, eIF1A, eIF3 and eIF5 [158]. This pre-initiation 43S complex attaches to unwounded mRNA activated by the eIF4F complex (consisting of eIF4E [a mRNA-cap binding component], eIF4G [a scaffolding protein] and eIF4A [an ATP- dependent RNA helicase]) [198] with assistance of eIF4B and poly(A)-binding protein (PABP) [199]. Then 48S initiation complex is formed and recognition of the initiation codon starts [158,159]. On the one hand, mTORC1 stimulates phosphorylation of 4E-binding proteins (4E-BPs), which is grouped into 4E-BP1, 4E-BP2 and 4E-BP3. Phosphorylation of 4EBP1 induces its release from eIF4E, leading to association of eIF4E with eIF4G. This then induces assembly and activation of the mRNA-cap binding eIF4F complex (eIF4E, eIF4G and eIF4A) [158,159,198]. The activated eIF4F complex stimulates translation of mRNA. On the other hand, mTORC1 activates ribosomal S6 kinases (S6Ks). The relevant substrates of S6Ks in translation regulation are ribosomal protein S6 (rpS6), eIF4B, eukaryotic elongation factor 2 kinase (eEF2K) and programmed cell death 4 protein (PDCD4) [159], which finally stimulate translation of mRNA.

Taken together, the PI3K/AKT/mTOR signaling effectors activate translation of HIF-1α mRNA. Then, stimulated HIF-1α activates transcription of CA9 via HRE. The PI3K/AKT/mTOR signaling pathway, HIF-1α and CA9 are involved in oncogenesis of ATL [200,201], which is caused by human T-cell leukemia virus type 1 (HTLV-1) [202,203]. On the one hand, proliferation of ATL cells is suppressed by numerous inhibitors against the PI3K/AKT/mTOR signaling effectors, including PI3K inhibitors (idelalisib [204], NVP-BMK120 [205,206], CDUC907 [206] and copanlisib [207]), AKT inhibitor (MK2206 [208]), mTORC1 inhibitors (torin2 [209], ramapycin [210] and RAD001/everolimus [205,210,211]), mTORC1/mTORC2 dual inhibitor (torin2 [209], PP242 [210] and AZT8055/sapanisertib [210,212]), PI3K/mTORC1 dual inhibitor (NVPBEZ235 [205]), or combination therapy of PI3K inhibitor with JAK1/2 inhibitor [212] or NF-κB inhibitor [207]. These in vitro and in vivo experimental trials mean indirect evidence of the involvement of the PI3K/AKT/mTOR signaling pathway in oncogenesis of ATL.

On the other hand, expression of AKT and CA9 is correlated with tumorigenicity. Yamaguchi [208] obtained AKT-activated ST1-N1 subclone from ATL patient-derived ST1 cell line [213] by serial transplantation into immune deficient NOG mice and proved correlation of AKT expression with tumorigenicity. Inhibition of AKT by inhibitor MK2206 abrogated tumorigenicity of ST1-N1. Furthermore, Nasu [55] confirmed that the responsible effector of tumorigenicity in ST1 was CA9 and that CA9 was expressed on ATL primary cells in lymph nodes of all the four cases of ATL patients. Tomita [214] demonstrated high expression of AKT and HIF-1α in both ATL cell lines and primary ATL cells. Cell growth inhibition by HIF-1α depletion using siRNA was also shown. These data clearly indicate that CA9 plays a central role in ATL oncogenesis and that CA9 inhibitors are expected to be effective against ATL. In addition, NF-κB plays a plethora of roles in the multistep oncogenesis of ATL at its early stage [215-218] and activation of HIFs by NK-κB has been detected in various mechanisms [39-43]. As a logical consequence, linkage between NK-κB and CA9 via HIFs in ATL oncogenesis is quite possible. In sum, these abundant experimental data strongly suggest that administration of CA9 inhibitors can be a promising novel therapeutic strategy against therapy resistant ATL.

Conclusion

CA9 is involved in oncogenesis of solid tumors and malignant lymphomas by its multifaceted roles in cancer cell proliferation, survival and metastasis and therapy resistance. In consequence of its critical role in malignant progression, CA9 inhibitors have great potentialities targeting therapy resistant TNBC and PDAC. Furthermore, since CA9 is suspected to be a substantial effector in oncogenic processes, a therapeutic strategy by CA9 inhibitors against IPMN and ATL is quite promising. Thus, one should start the in vitro experimental therapeutic trials for IPMN and ATL cells.

Acknowledgments

I cordially acknowledge Professor Emeritus Kazuo Sugamura, MD, PhD, at Tohoku University for his encouragement and permission to incorporate into this review article the achievements of his research group in the Institute of Miyagi Cancer Center about the involvement of AKT and CA9 in tumorigenicity of ATL cell lines.

Conflict Of Interest

The author declares that there is no conflict of interests.

References

  1. Vaupel, Peter and Louis Harrison. "Tumor hypoxia: Causative factors, compensatory mechanisms and cellular response." Oncol 9 (2004): 4-9.
  2. Google Scholar, Crossref, Indexed at

  3. Matolay, Orsolya, and Gábor Méhes. "Sustain, Adapt and Overcome—Hypoxia associated changes in the progression of lymphatic neoplasia." Front Oncol 9 (2019): 1277.
  4. Google Scholar, Crossref, Indexed at

  5. Kaelin Jr, William G., and Peter J. Ratcliffe. "Oxygen sensing by metazoans: The central role of the HIF hydroxylase pathway." Mol Cell 30 (2008): 393-402.
  6. Google Scholar, Crossref, Indexed at

  7. Serocki, Marcin, Sylwia Bartoszewska, Anna Janaszak-Jasiecka and Renata J. Ochocka, et al. "miRNAs regulate the HIF switch during hypoxia: A novel therapeutic target." Angiogenesis 21 (2018): 183-202.
  8. Google Scholar, Crossref, Indexed at

  9. Wykoff, Charles C., Nigel JP Beasley, Peter H. Watson and Kevin J. Turner, et al. "Hypoxia-inducible expression of tumor-associated carbonic anhydrases." Cancer res 60 (2000): 7075-7083.
  10. Google Scholar, Crossref, Indexed at

  11. Bárdos JI, Ashcroft M (2004) Hypoxia-inducible factor-1 and oncogenic signalling. Bioassays 26: 262-9
  12. Google Scholar, Crossref, Indexed at

  13. Ravi, Rajani, Bijoyesh Mookerjee, Zaver M. Bhujwalla and Carrie Hayes Sutter, et al. "Regulation of tumor angiogenesis by p53-induced degradation of hypoxia-inducible factor 1α." Genes Dev 14 (2000): 34-44.
  14. Google Scholar, Crossref, Indexed at

  15. Kaluzová, Milota, Stefan Kaluz, Michael I. Lerman and Eric J. Stanbridge, et al. "DNA damage is a prerequisite for p53-mediated proteasomal degradation of HIF-1α in hypoxic cells and downregulation of the hypoxia marker carbonic anhydrase IX." Mol Cell Biol 24 (2004): 5757-5766.
  16. Google Scholar, Crossref, Indexed at

  17. Freedman, Steven J., Zhen-Yu J. Sun, Florence Poy and Andrew L. Kung, et al. "Structural basis for recruitment of CBP/p300 by hypoxia-inducible factor-1α." Proc Natl Acad Sci 99 (2002): 5367-5372.
  18. Google Scholar, Crossref, Indexed at

  19. Safran, Michal and William G. Kaelin. "HIF hydroxylation and the mammalian oxygen-sensing pathway." J Clin Investig 111 (2003): 779-783.
  20. Google Scholar, Crossref, Indexed at

  21. Kallio, Pekka J., Kensaku Okamoto, Sallyann O'Brien and Pilar Carrero, et al. "Signal transduction in hypoxic cells: Inducible nuclear translocation and recruitment of theCBP/p300 coactivator by the hypoxia-induciblefactor-1α." J EMBO 17 (1998): 6573-6586.
  22. Google Scholar, Crossref, Indexed at

  23. Mandl, Markus and Reinhard Depping. "Hypoxia-inducible aryl hydrocarbon receptor nuclear translocator (ARNT)(HIF-1β): is it a rare exception?." J Mol Med 20 (2014): 215-220.
  24. Google Scholar, Crossref, Indexed at

  25. Vorrink, Sabine U., and Frederick E. Domann. "Regulatory crosstalk and interference between the xenobiotic and hypoxia sensing pathways at the AhR-ARNT-HIF1α signaling node." Chem Biol Interact 218 (2014): 82-88.
  26. Google Scholar, Crossref, Indexed at

  27. Wenger, Roland H., Daniel P. Stiehl and Gieri Camenisch. "Integration of oxygen signaling at the consensus HRE." Science s STKE (2005): re12-re12.
  28. Google Scholar, Crossref, Indexed at

  29. Wenger, Roland H. "Cellular adaptation to hypoxia: O2‐sensing protein hydroxylases, hypoxia‐inducible transcription factors, and O2‐regulated gene expression." The FASEB journal 16 (2002): 1151-1162.
  30. Google Scholar, Crossref, Indexed at

  31. Balamurugan, Kuppusamy. "HIF‐1 at the crossroads of hypoxia, inflammation, and cancer." Int J Cancer 138 (2016): 1058-1066.
  32. Google Scholar, Crossref, Indexed at

  33. Masoud, Georgina N., and Wei Li. "HIF-1α pathway: role, regulation and intervention for cancer therapy." Acta Pharm Sin B B 5 (2015): 378-389.
  34. Google Scholar, Crossref, Indexed at

  35. Chiche, Johanna, Karine Ilc, Julie Laferriere and Eric Trottier, et al. "Hypoxia-inducible carbonic anhydrase IX and XII promote tumor cell growth by counteracting acidosis through the regulation of the intracellular pH.Cancer res 69 (2009): 358-368.
  36. Google Scholar, Crossref, Indexed at

  37. Swietach, Pawel, Shalini Patiar, Claudiu T. Supuran and Adrian L. Harris, et al. "The role of carbonic anhydrase 9 in regulating extracellular and intracellular ph in three-dimensional tumor cell growths." J Biol Chem 284 (2009): 20299-20310.
  38. Google Scholar, Crossref, Indexed at

  39. Fiaschi, Tania, Elisa Giannoni, Letizia Taddei and Paolo Cirri, et al. "Carbonic anhydrase IX from cancer-associated fibroblasts drives epithelial-mesenchymal transition in prostate carcinoma cells." Cell Cycle 12 (2013): 1791-1801.
  40. Google Scholar, Crossref, Indexed at

  41. Lock, F. E., P. C. McDonald, Y. Lou and I. Serrano, et al. "Targeting carbonic anhydrase IX depletes breast cancer stem cells within the hypoxic niche." Oncogene 32 (2013): 5210-5219.
  42. Google Scholar, Crossref, Indexed at

  43. Güttler, Antje, Katharina Theuerkorn, Anne Riemann and Henri Wichmann, et al. "Cellular and radiobiological effects of carbonic anhydrase IX in human breast cancer cells." Oncol Rep 41 (2019): 2585-2594.
  44. Google Scholar, Crossref, Indexed at

  45. Sarnella, Annachiara, Giuliana D’Avino, Billy Samuel Hill and Vincenzo Alterio, et al. "A novel inhibitor of carbonic anhydrases prevents hypoxia-induced TNBC cell plasticity." Int J Mol Sci 21 (2020): 8405.
  46. Google Scholar, Crossref, Indexed at

  47. Tafreshi, Narges K., Mark C. Lloyd, Marilyn M. Bui and Robert J. Gillies, et al. "Carbonic anhydrase IX as an imaging and therapeutic target for tumors and metastases." Carbonic anhyd applicat (2014): 221-254.
  48. Google Scholar, Crossref, Indexed at

  49. Mahon, Brian P., Melissa A. Pinard and Robert McKenna. "Targeting carbonic anhydrase IX activity and expression." Molecules 20 (2015): 2323-2348.
  50. Google Scholar, Crossref, Indexed at

  51. Mboge, Mam Y., Anusha Kota, Robert McKenna and Susan C. Frost, et al. "Biophysical, biochemical, and cell Based approaches used to decipher the role of carbonic anhydrases in cancer and to evaluate the potency of targeted inhibitors." Int J Med Chem 2018 (2018).
  52. Google Scholar, Crossref, Indexed at

  53. Capasso, Clemente and Claudiu T. Supuran. "An overview of the alpha-, beta-and gamma-carbonic anhydrases from Bacteria: Can bacterial carbonic anhydrases shed new light on evolution of bacteria?." J Enzyme Inhib Med Chem 30 (2015): 325-332.
  54. Google Scholar, Crossref, Indexed at

  55. Mboge, Mam Y., Brian P. Mahon, Robert McKenna and Susan C. Frost, et al. "Carbonic anhydrases: Role in pH control and cancer." Metabolites 8 (2018): 19.
  56. Google Scholar, Crossref, Indexed at

  57. McDonald, Paul C., Mridula Swayampakula and Shoukat Dedhar. "Coordinated regulation of metabolic transporters and migration/invasion by carbonic anhydrase IX." Metabolites 8 (2018): 20.
  58. Google Scholar, Crossref, Indexed at

  59. McDonald, Paul C., Jean-Yves Winum, Claudiu T. Supuran and Shoukat Dedhar, et al. "Recent developments in targeting carbonic anhydrase IX for cancer therapeutics." Oncotarget 3 (2012): 84.
  60. Google Scholar, Crossref, Indexed at

  61. Supuran, Claudiu T. "Inhibition of carbonic anhydrase IX as a novel anticancer mechanism." World journal of clinical oncology 3 (2012): 98.
  62. Google Scholar, Crossref, Indexed at

  63. Supuran, Claudiu T. "Carbonic anhydrase inhibition and the management of hypoxic tumors." Metabolites 7 (2017): 48.
  64. Google Scholar, Crossref, Indexed at

  65. McDonald, Paul C., Shawn C. Chafe, Claudiu T. Supuran and Shoukat Dedhar, et al. "Cancer therapeutic targeting of hypoxia induced carbonic anhydrase IX: From bench to bedside." Cancers 14 (2022): 3297.
  66. Google Scholar, Crossref, Indexed at

  67. Kaluz, Stefan, Milota Kaluzová, René Opavský and Silvia Pastoreková, et al. "Transcriptional regulation of the MN/CA 9 gene coding for the tumor-associated carbonic anhydrase IX: identification and characterization of a proximal silencer element." J Biol Chem 274 (1999): 32588-32595.
  68. Google Scholar, Crossref, Indexed at

  69. Kaluz, Stefan, Milota Kaluzová, Shu-Yuan Liao and Michael Lerman, et al. "Transcriptional control of the tumor-and hypoxia-marker carbonic anhydrase 9: A one transcription factor (HIF-1) show?." Biochimica et Biophysica Acta (BBA)-Reviews on Cancer 1795 (2009): 162-172.
  70. Google Scholar, Crossref, Indexed at

  71. Kaluz, Stefan, Milota Kaluzová and Eric J. Stanbridge. "Expression of the hypoxia marker carbonic anhydrase IX is critically dependent on SP1 activity. Identification of a novel type of hypoxia-responsive enhancer." Cancer  res 63 (2003): 917-922.
  72. Google Scholar, Crossref, Indexed at

  73. Kaluz, Stefan, Milota Kaluzová and Eric J. Stanbridge. "Regulation of gene expression by hypoxia: Integration of the HIF-transduced hypoxic signal at the hypoxia-responsive element." Clin Chim Acta 395 (2008): 6-13.
  74. Google Scholar, Crossref, Indexed at

  75. Kaluzova, Milota, Silvia Pastoreková, Eliska Svastová and Jaromír Pastorek, et al. "Characterization of the MN/CA 9 promoter proximal region: A role for specificity protein (SP) and activator protein 1 (AP1) factors." Biochem J 359 (2001): 669-677.
  76. Google Scholar, Crossref, Indexed at

  77. BelAiba, Rachida S., Steve Bonello, Christian Zähringer and Stefanie Schmidt, et al. "Hypoxia up-regulates hypoxia-inducible factor-1α transcription by involving phosphatidylinositol 3-kinase and nuclear factor κB in pulmonary artery smooth muscle cells." Mol Biol Cell 18= (2007): 4691-4697.
  78. Google Scholar, Crossref, Indexed at

  79. Bonello, Steve, Christian Zähringer, Rachida S. BelAiba and Talija Djordjevic, et al. "Reactive oxygen species activate the HIF-1α promoter via a functional NFκB site." Arterioscler Thromb Vasc Biol 27 (2007): 755-761.
  80. Google Scholar, Crossref, Indexed at

  81. Wenger, Roland H., Gieri Camenisch, Daniel P. Stiehl and Dorthe M. Katschinski, et al. "HIF prolyl-4-hydroxylase interacting proteins: Consequences for drug targeting." Curr Pharm Des 15 (2009): 3886-3894.
  82. Google Scholar, Crossref, Indexed at

  83. Rius, Jordi, Monica Guma, Christian Schachtrup and Katerina Akassoglou, et al. "NF-κB links innate immunity to the hypoxic response through transcriptional regulation of HIF-1α." Nature 453, (2008): 807-811.
  84. Google Scholar, Crossref, Indexed at

  85. Sun, Heng, Xue-Bing Li, Ya Meng and Li Fan, et al. "TRAF6 upregulates expression of HIF-1α and promotes tumor angiogenesis." Cancer Res 73 (2013): 4950-4959.
  86. Google Scholar, Crossref, Indexed at

  87. Blancher, Christine, John W. Moore, Naomi Robertson and Adrian L. Harris, et al. "Effects of ras and von Hippel-Lindau (VHL) gene mutations on hypoxia-inducible factor (HIF)-1α, HIF-2α, and vascular endothelial growth factor expression and their regulation by the phosphatidylinositol 3′-kinase/Akt signaling pathway.Cancer res 61 (2001): 7349-7355.
  88. Google Scholar, Crossref, Indexed at

  89. Zundel, Wayne, Cornelia Schindler, Daphne Haas-Kogan and Albert Koong, et al. "Loss of PTEN facilitates HIF-1-mediated gene expression." Genes Dev 14 (2000): 391-396.
  90. Google Scholar, Crossref, Indexed at

  91. Jing, Yi, Ling-Zhi Liu, Yue Jiang and Yingxue Zhu, et al. "Cadmium increases HIF-1 and VEGF expression through ROS, ERK, and AKT signaling pathways and induces malignant transformation of human bronchial epithelial cells." Toxicol Sci 125 (2012): 10-19.
  92. Google Scholar, Crossref, Indexed at

  93. Wan, Jun, and Wei Wu. "Hyperthermia induced HIF-1a expression of lung cancer through AKT and ERK signaling pathways." J Exp Clin Cancer Res 35 (2016): 1-11.
  94. Google Scholar, Crossref, Indexed at

  95. Xu, Xuewen, Kai You, and Renge Bu. "Proximal tubular development is impaired with downregulation of MAPK/ERK signaling, HIF-1α, and catalase by hyperoxia exposure in neonatal rats." Oxid Med Cell Longev 2019 (2019).
  96. Google Scholar, Crossref, Indexed at

  97. Sang, Nianli, Daniel P. Stiehl, Jolene Bohensky and Irene Leshchinsky, et al. "MAPK signaling up-regulates the activity of hypoxia-inducible factors by its effects on p300." J Biol Chem 278 (2003): 14013-14019.
  98. Google Scholar, Crossref, Indexed at

  99. Mohammad, Ramzi M., Irfana Muqbil, Leroy Lowe and Clement Yedjou, et al. "Broad targeting of resistance to apoptosis in cancer." Semin Cancer Biol 35 Academic Press, 2015.
  100. Google Scholar, Crossref, Indexed at

  101. Morana, Ornella, Will Wood and Christopher D. Gregory. "The apoptosis paradox in cancer." Int J Mol Sci 23 (2022): 1328.
  102. Google Scholar, Crossref, Indexed at

  103. Csaderova, Lucia, Michaela Debreova, Peter Radvak and Matej Stano, et al. "The effect of carbonic anhydrase IX on focal contacts during cell spreading and migration." Front Physiol 4 (2013): 271.
  104. Google Scholar, Crossref, Indexed at

  105. Wouters, An, Bea Pauwels, Filip Lardon and Jan B. Vermorken, et al. "Implications of in vitro research on the effect of radiotherapy and chemotherapy under hypoxic conditions." Oncologist 12 (2007): 690-712.
  106. Google Scholar, Crossref, Indexed at

  107. Tan, E. Y., M. Yan, L. Campo and C. Han, et al. "The key hypoxia regulated gene CAIX is upregulated in basal-like breast tumours and is associated with resistance to chemotherapy." Br J Cancer 100 (2009): 405-411.
  108. Google Scholar, Crossref, Indexed at

  109. Nasu, Kentaro, Kazunori Yamaguchi, Tomoka Takanashi and Keiichi Tamai, et al. "Crucial role of carbonic anhydrase IX in tumorigenicity of xenotransplanted adult T‐cell leukemia‐derived cells." Cancer science 108 (2017): 435-443.
  110. Google Scholar, Crossref, Indexed at

  111. Zamanova, Sabina, Ahmed M. Shabana, Utpal K. Mondal and Marc A. Ilies, et al. "Carbonic anhydrases as disease markers." Expert Opin Ther Pat 29 (2019): 509-533.
  112. Google Scholar, Crossref, Indexed at

  113. Trastour, Cynthia, Emmanuel Benizri, Francette Ettore and Alain Ramaioli, et al. "HIF‐1α and CA IX staining in invasive breast carcinomas: Prognosis and treatment outcome." Int J Cancer 120 (2007): 1451-1458.
  114. Google Scholar, Crossref, Indexed at

  115. Zandberga, Elīna, Pawel Zayakin, Artūrs Ābols and Dārta Pūpola, et al. "Depletion of carbonic anhydrase IX abrogates hypoxia-induced overexpression of stanniocalcin-1 in triple negative breast cancer cells." Cancer Biol Ther 18 (2017): 596-605.
  116. Google Scholar, Crossref, Indexed at

  117. Logsdon, Derek P., Michelle Grimard, Meihua Luo and Safi Shahda, et al. "Regulation of HIF1α under hypoxia by APE1/Ref-1 impacts CA9 expression: Dual targeting in patient-derived 3D pancreatic cancer models dual-targeting APE1/Ref-1 and CA9 in hypoxic PDAC cells." Mol Cancer Ther 15 (2016): 2722-2732.
  118. Google Scholar, Crossref, Indexed at

  119. Logsdon, Derek P., Fenil Shah, Fabrizio Carta and Claudiu T. Supuran, et al. "Blocking HIF signaling via novel inhibitors of CA9 and APE1/Ref-1 dramatically affects pancreatic cancer cell survival." Sci Rep 8 (2018): 1-14.
  120. Google Scholar, Crossref, Indexed at

  121. McDonald, Paul C., Shawn C. Chafe, Wells S. Brown and Saeed Saberi, et al. "Regulation of pH by carbonic anhydrase 9 mediates survival of pancreatic cancer cells with activated KRAS in response to hypoxia." Gastroenterology 157 (2019): 823-837.
  122. Google Scholar, Crossref, Indexed at

  123. Leibovich, Bradley C., Yuri Sheinin, Christine M. Lohse and R. Houston et al. "Carbonic anhydrase IX is not an independent predictor of outcome for patients with clear cell renal cell carcinoma." J Clin Oncol 25 (2007): 4757-4764.
  124. Google Scholar, Crossref, Indexed at

  125. Genega, Elizabeth M., Musie Ghebremichael, Robert Najarian and Yineng Fu, et al. "Carbonic anhydrase IX expression in renal neoplasms: Correlation with tumor type and grade." Am J Clin Pathol 134 (2010): 873-879.
  126. Google Scholar, Crossref, Indexed at

  127. Giatromanolaki, Alexandra, Michael I. Koukourakis, Efthimios Sivridis and Jaromir Pastorek, et al. "Expression of hypoxia-inducible carbonic anhydrase-9 relates to angiogenic pathways and independently to poor outcome in non-small cell lung cancer." Cancer res 61 (2001): 7992-7998.
  128. Google Scholar, Crossref, Indexed at

  129. Ilie, M., N. M. Mazure, Véronique Hofmanm and R. E. Ammadi et al. "High levels of carbonic anhydrase IX in tumour tissue and plasma are biomarkers of poor prognostic in patients with non-small cell lung cancer." Br J Cancer 102 (2010): 1627-1635.
  130. Google Scholar, Crossref, Indexed at

  131. Kivela, Antti J., Seppo Parkkila, Juha Saarnio and Tuomo J. Karttunen, et al. "Expression of von Hippel-Lindau tumor suppressor and tumor-associated carbonic anhydrases IX and XII in normal and neoplastic colorectal mucosa." World J Gastroenterol 11 (2005): 2616.
  132. Google Scholar, Crossref, Indexed at

  133. Huang, Lihua, and Liwu Fu. "Mechanisms of resistance to EGFR tyrosine kinase inhibitors." Acta Pharm Sin 5 (2015): 390-401.
  134. Google Scholar, Crossref, Indexed at

  135. Koyuncu, Ismail, Ataman Gonel, Abdurrahim Kocyigit and Ebru Temiz, et al. "Selective inhibition of carbonic anhydrase-IX by sulphonamide derivatives induces pH and reactive oxygen species-mediated apoptosis in cervical cancer HeLa cells." J Enzyme Inhib Med Chem 33 (2018): 1137-1149.
  136. Google Scholar, Crossref, Indexed at

  137. Järvelä, Sally, Seppo Parkkila, Helena Bragge and Marketta Kähkönen, et al. "Carbonic anhydrase IX in oligodendroglial brain tumors." BMC cancer 8 (2008): 1-9.
  138. Google Scholar, Crossref, Indexed at

  139. Ohtaki, Yoichi, Kimihiro Shimizu, Reika Kawabata-Iwakawa and Navchaa Gombodorj, et al. "Carbonic anhydrase 9 expression is associated with poor prognosis, tumor proliferation, and radiosensitivity of thymic carcinomas." Oncotarget 10 (2019): 1306.
  140. Google Scholar, Crossref, Indexed at

  141. Matolay, Orsolya, Lívia Beke, Andrea Gyurkovics and Mónika Francz, et al. "Quantitative analysis of carbonic anhydrase IX uncovers hypoxia-related functional differences in classical hodgkin lymphoma subtypes." Int J Mol Sci 20 (2019): 3463.
  142. Google Scholar, Crossref, Indexed at

  143. Méhes, Gábor, Orsolya Matolay, Livia Beke and Marianna Czenke, et al. "Hypoxia‐related carbonic anhydrase IX expression is associated with unfavourable response to first‐line therapy in classical Hodgkin's lymphoma." Histopathology 74 (2019): 699-708.
  144. Google Scholar, Crossref, Indexed at

  145. Stewart, M., K. Talks, R. Leek and H. Turley, et al. "Expression of angiogenic factors and hypoxia inducible factors HIF 1, HIF 2 and CA IX in non‐hodgkin's lymphoma." Histopathology 40 (2002): 253-260.
  146. Google Scholar, Crossref, Indexed at

  147. Chen, Liu Qi, Christine M. Howison, Catherine Spier and Alison T. Stopeck, et al. "Assessment of carbonic anhydrase IX expression and extracellular pH in B-cell lymphoma cell line models." Leuk Lymphoma 56 (2015): 1432-1439.
  148. Google Scholar, Crossref, Indexed at

  149. Lounnas, Nadia, Célia Rosilio, Marielle Nebout and Didier Mary, et al. "Pharmacological inhibition of carbonic anhydrase XII interferes with cell proliferation and induces cell apoptosis in T-cell lymphomas." Cancer Lett 333 (2013): 76-88.
  150. Google Scholar, Crossref, Indexed at

  151. Jin, Zhaoyu, and Wafik S. El-Deiry. "Overview of cell death signaling pathways." Cancer Biol Ther 4 (2005): 147-171.
  152. Google Scholar, Crossref, Indexed at

  153. Goldar, Samira, Mahmoud Shekari Khaniani, Sima Mansoori Derakhshan and Behzad Baradaran, et al. "Molecular mechanisms of apoptosis and roles in cancer development and treatment." Asian Pac J Cancer Prev 16 (2015): 2129-2144.
  154. Google Scholar, Crossref, Indexed at

  155. Pistritto, Giuseppa, Daniela Trisciuoglio, Claudia Ceci and Alessia Garufi, et al. "Apoptosis as anticancer mechanism: Function and dysfunction of its modulators and targeted therapeutic strategies." Aging 8 (2016): 603.
  156. Google Scholar, Crossref, Indexed at

  157. Guicciardi, Maria Eugenia and Gregory J. Gores. "Life and death by death receptors." The FASEB Journal 23 (2009): 1625-1637.
  158. Google Scholar, Crossref, Indexed at

  159. Park, Hyun Ho, Yu-Chih Lo, Su-Chang Lin and Liwei Wang, et al. "The death domain superfamily in intracellular signaling of apoptosis and inflammation." Ann Rev Immunol 25 (2007).
  160. Google Scholar, Crossref, Indexed at

  161. Khosravi-Far, Roya. "Death receptor signals to the mitochondria." Cancer Biol Ther 3 (2004): 1051-1057.
  162. Google Scholar, Crossref, Indexed at

  163. Boatright, Kelly M., Martin Renatus, Fiona L. Scott and Sabina Sperandio, et al. "A unified model for apical caspase activation." Mol Cell 11 (2003): 529-541.
  164. Google Scholar, Crossref, Indexed at

  165. Salvesen, Guy S., and Vishva M. Dixit. "Caspase activation: The induced-proximity model." Proc Natl Acad Sci U S A 96 (1999): 10964-10967.
  166. Google Scholar, Crossref, Indexed at

  167. Green, Douglas R., and Guido Kroemer. "The pathophysiology of mitochondrial cell death." Science 305 (2004): 626-629.
  168. Google Scholar, Crossref, Indexed at

  169. Kroemer, Guido, Lorenzo Galluzzi and Catherine Brenner. "Mitochondrial membrane permeabilization in cell death." Physiol Rev 87 (2007): 99-163.
  170. Google Scholar, Crossref, Indexed at

  171. Chipuk, J. E., L. Bouchier-Hayes and D. R. Green. "Mitochondrial outer membrane permeabilization during apoptosis: The innocent bystander scenario." Cell Death Differ 13 (2006): 1396-1402.
  172. Google Scholar, Crossref, Indexed at

  173. Von Ahsen, O., N. J. Waterhouse, T. Kuwana and D. D. Newmeyer, et al. "The ‘harmless’ release of cytochrome c." Cell Death Differ 7 (2000): 1192-1199.
  174. Google Scholar, Crossref, Indexed at

  175. Yuan, Shujun and Christopher W. Akey. "Apoptosome structure, assembly, and procaspase activation." Structure 21 (2013): 501-515.
  176. Google Scholar, Crossref, Indexed at

  177. Vousden, Karen H. "p53: Death star." Cell 103 (2000): 691-694.
  178. Google Scholar, Crossref, Indexed at

  179. Chipuk, Jerry E., Tomomi Kuwana, Lisa Bouchier-Hayes and Nathalie M. Droin, et al. "Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis." Science 303 (2004): 1010-1014.
  180. Google Scholar, Crossref, Indexed at

  181. Deveraux, Quinn L and John C. Reed. "IAP family proteins—suppressors of apoptosis." Genes Dev 13 (1999): 239-252.
  182. Google Scholar, Crossref, Indexed at

  183. Verhagen, Anne M., Elizabeth J. Coulson and David L. Vaux. "Inhibitor of apoptosis proteins and their relatives: IAPs and other BIRPs." Genome Biol 2 (2001): 1-10.
  184. Google Scholar, Crossref, Indexed at

  185. Wong, Rebecca SY. "Apoptosis in cancer: from pathogenesis to treatment." J Exp Clin Cancer 30 (2011): 1-14.
  186. Google Scholar, Crossref, Indexed at

  187. Temiz, Ebru, Ismail Koyuncu, Mustafa Durgun and Murat Caglayan, et al. "Inhibition of carbonic anhydrase IX promotes apoptosis through intracellular PH level alterations in cervical cancer cells." International Journal of Molecular Sciences 22 (2021): 6098.
  188. Google Scholar, Crossref, Indexed at

  189. Koyuncu, Ismail, Ataman Gonel, Mustafa Durgun and Abdurrahim Kocyigit, et al. "Assessment of the antiproliferative and apoptotic roles of sulfonamide carbonic anhydrase IX inhibitors in hela cancer cell line." J Enzyme Inhib Med Chem 34 (2019): 75-86.
  190. Google Scholar, Crossref, Indexed at

  191. Li, Zan, Li Jiang, Shan Hwu Chew and Tasuku Hirayama, et al. "Carbonic anhydrase 9 confers resistance to ferroptosis/apoptosis in malignant mesothelioma under hypoxia." Redox Biol 26 (2019): 101297.
  192. Google Scholar, Crossref, Indexed at

  193. Swayampakula, M., P. C. McDonald, M. Vallejo and E. Coyaud, et al. "The interactome of metabolic enzyme carbonic anhydrase IX reveals novel roles in tumor cell migration and invadopodia/MMP14-mediated invasion." Oncogene 36 (2017): 6244-6261.
  194. Google Scholar, Crossref, Indexed at

  195. Debreova, Michaela, Lucia Csaderova, Monika Burikova and Lubomira Lukacikova, et al. "CAIX regulates invadopodia formation through both a pH-dependent mechanism and interplay with actin regulatory proteins." J Mol Sci 20 (2019): 2745.
  196. Google Scholar, Crossref, Indexed at

  197. Beaty, Brian T., Yarong Wang, Jose Javier Bravo-Cordero and Ved P. Sharma, et al. "Talin regulates moesin–NHE-1 recruitment to invadopodia and promotes mammary tumor metastasis." J Cell Biol 205 (2014): 737-751.
  198. Google Scholar, Crossref, Indexed at

  199. Amith, Schammim Ray, Sunny Fong, Shairaz Baksh and Larry Fliegel, et al. "Na+/H+ exchange in the tumour microenvironment: does NHE1 drive breast cancer carcinogenesis?." Int J Dev Biol 59 (2015).
  200. Google Scholar, Crossref, Indexed at

  201. Ciccone, Valerio, Arianna Filippelli, Andrea Angeli and Claudiu T. Supuran, et al. "Pharmacological inhibition of CA-IX impairs tumor cell proliferation, migration and invasiveness." Int J Mol Sci 21 (2020): 2983.
  202. Google Scholar, Crossref, Indexed at

  203. Lou, Yuanmei, Paul C. McDonald, Arusha Oloumi and Stephen Chia, et al. "Targeting tumor hypoxia: Suppression of breast tumor growth and metastasis by novel carbonic anhydrase IX inhibitors targeted inhibition of CAIX suppresses breast cancer progression targeted inhibition of CAIX suppresses breast cancer progression." Cancer res 71 (2011): 3364-3376.
  204. Google Scholar, Crossref, Indexed at

  205. Janoniene, Agne, Linas Mazutis, Daumantas Matulis and Vilma Petrikaite, et al. "Inhibition of carbonic anhydrase IX suppresses breast cancer cell motility at the single-cell level." Int J Mol Sci 22 (2021): 11571.
  206. Google Scholar, Crossref, Indexed at

  207. Ward, Carol, James Meehan, Peter Mullen and Claudiu Supuran, et al. "Evaluation of carbonic anhydrase IX as a therapeutic target for inhibition of breast cancer invasion and metastasis using a series of in vitro breast cancer models." Oncotarget 6 (2015): 24856.
  208. Google Scholar, Crossref, Indexed at

  209. Radvak, Peter, Marko Repic, Eliska Svastova and Martina Takacova, et al. "Suppression of carbonic anhydrase IX leads to aberrant focal adhesion and decreased invasion of tumor cells." Oncol Rep 29 (2013): 1147-1153.
  210. Google Scholar, Crossref, Indexed at

  211. Stewart, Erin L., Samuel Zhixing Tan, Geoffrey Liu and Ming-Sound Tsao, et al. "Known and putative mechanisms of resistance to EGFR targeted therapies in NSCLC patients with EGFR mutations—a review." Transl Lung Cancer 4 (2015): 67.
  212. Google Scholar, Crossref, Indexed at

  213. Aggarwal, Charu. "Targeted therapy for lung cancer: Present and future." Ann Palliat Med 3 (2014): 229-235.
  214. Google Scholar, Crossref, Indexed at

  215. Ikeda, Koei, Hiroaki Nomori, Takeshi Mori and Jiichiro Sasaki, et al. "Novel germline mutation: EGFR V843I in patient with multiple lung adenocarcinomas and family members with lung cancer." Ann Thorac Surg 85 (2008): 1430-1432.
  216. Google Scholar, Crossref, Indexed at

  217. Girard, Nicolas, Emil Lou, Christopher G. Azzoli and Rekha Reddy, et al. "Analysis of genetic variants in never-smokers with lung cancer facilitated by an Internet-based blood collection protocol: A preliminary report." Clin Cancer Res 16 (2010): 755-763.
  218. Google Scholar, Crossref, Indexed at

  219. Balak, Marissa N., Yixuan Gong, Gregory J. Riely and Romel Somwar, et al. "Novel D761Y and common secondary T790M mutations in epidermal growth factor receptor–mutant lung adenocarcinomas with acquired resistance to kinase inhibitors." Clin Cancer Res 12 (2006): 6494-6501.
  220. Google Scholar, Crossref, Indexed at

  221. Sasaki, Takaaki, Jussi Koivunen, Atsuko Ogino and Masahiko Yanagita, et al. "A Novel ALK secondary mutation and EGFR signaling cause resistance to ALK kinase inhibitors resistance mechanisms in ALK kinase inhibitors." Cancer res 71 (2011): 6051-6060.
  222. Google Scholar, Crossref, Indexed at

  223. Engelman, Jeffrey A., Toru Mukohara, Kreshnik Zejnullahu and Eugene Lifshits, et al. "Allelic dilution obscures detection of a biologically significant resistance mutation in EGFR-amplified lung cancer." J Clin Invest 116 (2006): 2695-2706.
  224. Google Scholar, Crossref, Indexed at

  225. Yamasaki, Fumiyuki, Mary J. Johansen, Dongwei Zhang and Savitri Krishnamurthy, et al. "Acquired resistance to erlotinib in A-431 epidermoid cancer cells requires down-regulation of MMAC1/PTEN and up-regulation of phosphorylated Akt." Cancer res 67 (2007): 5779-5788.
  226. Google Scholar, Crossref, Indexed at

  227. Sharifnia, Tanaz, Victor Rusu, Federica Piccioni and Mukta Bagul, et al. "Genetic modifiers of EGFR dependence in non-small cell lung cancer." Proc Natrl Acad Sci U S A 111 (2014): 18661-18666.
  228. Google Scholar, Crossref, Indexed at

  229. Lovly, Christine M. "Combating acquired resistance to tyrosine kinase inhibitors in lung cancer." Am Soc Clin Oncol Educ Book 35 (2015): e165-e173.
  230. Google Scholar, Crossref, Indexed at

  231. Shackleton, Mark, Elsa Quintana, Eric R. Fearon and Sean J. Morrison, et al. "Heterogeneity in cancer: Cancer stem cells versus clonal evolution." Cell 138 (2009): 822-829.
  232. Google Scholar, Crossref, Indexed at

  233. Gay, L., A. M. Baker and T. A. Graham. "Tumour Cell Heterogeneity. F1000Res 5." F1000 Faculty Rev 238 (2016).
  234. Google Scholar, Crossref, Indexed at

  235. Ciurea, Marius Eugen, Ada Maria Georgescu, Stefana Oana Purcaru and Stefan-Alexandru Artene, et al. "Cancer stem cells: Biological functions and therapeutically targeting." Int J Mol Sci 15 (2014): 8169-8185.
  236. Google Scholar, Crossref, Indexed at

  237. Szumna, Klaudia Daria, Faustyna Piędel, Patryk Jasielski and Sylwia Grosman, et al. "Cancer stem cells as a new promising approach of efficient oncological treatment-the review of literature." J Educ Health Sport 10 (2020): 115-120.
  238. Google Scholar, Crossref, Indexed at

  239. Bai, Hua, Zhijie Wang, Yuyan Wang and Minglei Zhuo, et al. "Detection and clinical significance of intratumoral EGFR mutational heterogeneity in Chinese patients with advanced non-small cell lung cancer." PloS one 8 (2013): e54170.
  240. Google Scholar, Crossref, Indexed at

  241. Remon, Jordi and Margarita Majem. "EGFR mutation heterogeneity and mixed response to EGFR tyrosine kinase inhibitors of non small cell lung cancer: A clue to overcoming resistance." Transl Lung Cancer Res 2 (2013): 445.
  242. Google Scholar, Crossref, Indexed at

  243. Guo, M., Y. Peng, A. Gao and C. Du, et al. "Epigenetic heterogeneity in cancer". Biomark Res 7 (2019).
  244. Google Scholar, Crossref, Indexed at

  245. Beyes, Sven, Naiara Garcia Bediaga and Alessio Zippo. "An epigenetic perspective on intra-tumour heterogeneity: Novel insights and new challenges from multiple fields." Cancers 13 (2021): 4969.
  246. Google Scholar, Crossref, Indexed at

  247. Marjanovic, Nemanja D., Robert A. Weinberg and Christine L. Chaffer. "Cell plasticity and heterogeneity in cancer." Clin chem 59 (2013): 168-179.
  248. Google Scholar, Crossref, Indexed at

  249. Yuan, Salina, Robert J. Norgard and Ben Z. Stanger. "Cellular plasticity in cancer cancer cells change identity during tumor progression." Cancer discov 9 (2019): 837-851.
  250. Google Scholar, Crossref, Indexed at

  251. Gupta, Piyush B., Ievgenia Pastushenko, Adam Skibinski and Cedric Blanpain, et al. "Phenotypic plasticity: Driver of cancer initiation, progression, and therapy resistance." Cell stem cell 24 (2019): 65-78.
  252. Google Scholar, Crossref, Indexed at

  253. Huang, Tianzhi, Xiao Song, Dandan Xu and Deanna Tiek, et al. "Stem cell programs in cancer initiation, progression, and therapy resistance." Theranostics 10 (2020): 8721.
  254. Google Scholar, Crossref, Indexed at

  255. Li, Xiaoxian, Michael T. Lewis, Jian Huang and Carolina Gutierrez, et al. "Intrinsic resistance of tumorigenic breast cancer cells to chemotherapy." J Natl Cancer Inst 100 (2008): 672-679.
  256. Google Scholar, Crossref, Indexed at

  257. Fultang, Norman, Madhuparna Chakraborty and Bela Peethambaran. "Regulation of cancer stem cells in triple negative breast cancer." Cancer Drug Resist 4 (2021): 321.
  258. Google Scholar, Crossref, Indexed at

  259. Marchand, Tony and Sandra Pinho. "Leukemic stem cells: From leukemic niche biology to treatment opportunities." Front Immunol 12 (2021): 775128.
  260. Google Scholar, Crossref, Indexed at

  261. Yabo, Yahaya A., Simone P. Niclou and Anna Golebiewska. "Cancer cell heterogeneity and plasticity: A paradigm shift in glioblastoma." Neuro Oncol 24 (2022): 669-682.
  262. Google Scholar, Crossref, Indexed at

  263. Reya, Tannishtha, Sean J. Morrison, Michael F. Clarke and Irving L. Weissman, et al. "Stem cells, cancer, and cancer stem cells." nature 414 (2001): 105-111.
  264. Google Scholar, Crossref, Indexed at

  265. Du, Bowen and Joong Sup Shim. "Targeting epithelial–mesenchymal transition (EMT) to overcome drug resistance in cancer." Molecules 21 (2016): 965.
  266. Google Scholar, Crossref, Indexed at

  267. Liao, Tsai‐Tsen and Muh‐Hwa Yang. "Revisiting epithelial‐mesenchymal transition in cancer metastasis: The connection between epithelial plasticity and stemness." Mol Oncol 11 (2017): 792-804.
  268. Google Scholar, Crossref, Indexed at

  269. Byers, Lauren Averett, Lixia Diao, Jing Wang and Pierre Saintigny, et al. "An epithelial–mesenchymal transition gene signature predicts resistance to EGFR and PI3K inhibitors and identifies Axl as a therapeutic target for overcoming EGFR inhibitor resistance EMT predicts EGFR and PI3K inhibitor resistance in NSCLC." Clin Cancer Res 19 (2013): 279-290.
  270. Google Scholar, Crossref, Indexed at

  271. Jakobsen, Kristine Raaby, Christina Demuth, Boe Sandahl Sorensen and Anders Lade Nielsen, et al. "The role of epithelial to mesenchymal transition in resistance to epidermal growth factor receptor tyrosine kinase inhibitors in non-small cell lung cancer." Transl Lung Cancer Res 5 (2016): 172.
  272. Google Scholar, Crossref, Indexed at

  273. Zhu, Xuan, Lijie Chen, Ling Liu and Xing Niu, et al. "EMT-mediated acquired EGFR-TKI resistance in NSCLC: mechanisms and strategies." Frontiers in Oncology 9 (2019): 1044.
  274. Google Scholar, Crossref, Indexed at

  275. Wade, Cameron A., and Natasha Kyprianou. "Profiling prostate cancer therapeutic resistance."Int J Mol Sci 19 (2018): 904.
  276. Google Scholar, Crossref, Indexed at

  277. Hyuga, Satoshi, Hiroshi Wada, Hidetoshi Eguchi and Toru Otsuru, et al. "Expression of carbonic anhydrase IX is associated with poor prognosis through regulation of the epithelial‑mesenchymal transition in hepatocellular carcinoma." Int J Oncol 51 (2017): 1179-1190.
  278. Google Scholar, Crossref, Indexed at

  279. Peng, Zhao, Chen-Xiao Wang, Er-Hu Fang and Guo-Bin Wang, et al. "Role of epithelial-mesenchymal transition in gastric cancer initiation and progression." World J Gastroenterol 20 (2014): 5403.
  280. Google Scholar, Crossref, Indexed at

  281. Andreucci, Elena, Jessica Ruzzolini, Silvia Peppicellia and Francesca Bianchini, et al. "The carbonic anhydrase IX inhibitor SLC-0111 sensitises cancer cells to conventional chemotherapy." J Enzyme Inhib Med Chem 34 (2019): 117-123.
  282. Google Scholar, Crossref, Indexed at

  283. Huang, Lamei, Zhixing Guo, Fang Wang and Liwu Fu, et al. "KRAS mutation: From undruggable to druggable in cancer." Signal Transdut Target Ther 6 (2021): 1-20.
  284. Google Scholar, Crossref, Indexed at

  285. Raphael, Benjamin J., Ralph H. Hruban, Andrew J. Aguirre and Richard A. Moffitt, et al. "Integrated genomic characterization of pancreatic ductal adenocarcinoma." Cancer cell 32 (2017): 185-203.
  286. Google Scholar, Crossref, Indexed at

  287. Hezel, Aram F., Alec C. Kimmelman, Ben Z. Stanger and Nabeel Bardeesy, et al. "Genetics and biology of pancreatic ductal adenocarcinoma." Genes Dev 20 (2006): 1218-1249.
  288. Google Scholar, Crossref, Indexed at

  289. Eser, S., A. Schnieke, G. Schneider and D. Saur, et al. "Oncogenic KRAS signalling in pancreatic cancer." Br J Cancer 111 (2014): 817-822.
  290. Google Scholar, Crossref, Indexed at

  291. Collins, Meredith A., Filip Bednar, Yaqing Zhang and Jean-Christophe Brisset, et al. "Oncogenic Kras is required for both the initiation and maintenance of pancreatic cancer in mice." J Clin Invest 122 (2012): 639-653.
  292. Google Scholar, Crossref, Indexed at

  293. Sprang, Stephen R. "How ras works: Structure of a rap–raf complex." Structure 3 (1995): 641-643.
  294. Google Scholar, Crossref, Indexed at

  295. Herrmann, Christian, and Nicolas Nassar. "Ras and its effectors." Progress in biophysics and Mol Biol 66 (1996): 1-41.
  296. Google Scholar, Crossref, Indexed at

  297. McKay, M. M., and D. K. Morrison. "Integrating signals from RTKs to ERK/MAPK." Oncogene 26 (2007): 3113-3121.’
  298. Google Scholar, Crossref, Indexed at

  299. Dillon, Martha, Antonio Lopez, Edward Lin and Dominic Sales, et al. "Progress on Ras/MAPK signaling research and targeting in blood and solid cancers." Cancers 13 (2021): 5059.
  300. Google Scholar, Crossref, Indexed at

  301. Castellano, Esther and Julian Downward. "RAS interaction with PI3K: More than just another effector pathway." Genes Cancer 2 (2011): 261-274.
  302. Google Scholar, Crossref, Indexed at

  303. Krygowska, Agata Adelajda and Esther Castellano. "PI3K: A crucial piece in the RAS signaling puzzle." Cold Spring Harb Perspect Med 8 (2018): a031450.
  304. Google Scholar, Crossref, Indexed at

  305. Lim, Kian-Huat, Antonio T. Baines, James J. Fiordalisi and Michail Shipitsin, et al. "Activation of RalA is critical for Ras-induced tumorigenesis of human cells." Cancer Cell 7 (2005): 533-545.
  306. Google Scholar, Crossref, Indexed at

  307. Kashatus, David F. "Ral GTPases in tumorigenesis: Emerging from the shadows." Exp Cell Res 319 (2013): 2337-2342.
  308. Google Scholar, Crossref, Indexed at

  309. Roman, Marta, Elizabeth Hwang and E. Alejandro Sweet-Cordero. "Synthetic Vulnerabilities in the KRAS Pathway." Cancers 14 (2022): 2837.
  310. Google Scholar, Crossref, Indexed at

  311. Johnson, Gary L and Razvan Lapadat. "Mitogen-activated protein kinase pathways mediated by ERK, JNK, and p38 protein kinases." Science 298 (2002): 1911-1912.
  312. Google Scholar, Crossref, Indexed at

  313. Dhillon, Amardeep Singh, Suzanne Hagan, O. Rath and W. Kolch, et al. "MAP kinase signalling pathways in cancer." Oncogene 26 (2007): 3279-3290.
  314. Google Scholar, Crossref, Indexed at

  315. Ali, Muhammad Umar, Muhammad Saif Ur Rahman, Zhenyu Jia and Cao Jiang, et al. "Eukaryotic translation initiation factors and cancer." Tumour Biol 39 (2017): 1010428317709805.
  316. Google Scholar, Crossref, Indexed at

  317. Hao, Peiqi, Jiaojiao Yu, Richard Ward and Yin Liu, et al. "Eukaryotic translation initiation factors as promising targets in cancer therapy." Cell Commun Signal 18 (2020): 1-20.
  318. Google Scholar, Crossref, Indexed at

  319. Aguirre, Andrew J and William C. Hahn. "Synthetic lethal vulnerabilities in KRAS-mutant cancers." Cold Spring Harb Perspect Med 8 (2018): a031518.
  320. Google Scholar, Crossref, Indexed at

  321. Hou, Pingping and Y. Alan Wang. "Conquering oncogenic KRAS and its bypass mechanisms." Theranostics 12 (2022): 5691.
  322. Google Scholar, Crossref, Indexed at

  323. Zeitouni, Daniel, Yuliya Pylayeva-Gupta, Channing J. Der and Kirsten L. Bryant, et al. "KRAS mutant pancreatic cancer: no lone path to an effective treatment." Cancers 8 (2016): 45.
  324. Google Scholar, Crossref, Indexed at

  325. Waters, Andrew M and Channing J. Der. "KRAS: the critical driver and therapeutic target for pancreatic cancer." Cold Spring Harb Perspect Med 8 (2018): a031435.
  326. Google Scholar, Crossref, Indexed at

  327. Peng, Sheng-Bin, James R. Henry, Michael D. Kaufman and Wei-Ping Lu, et al. "Inhibition of RAF isoforms and active dimers by LY3009120 leads to anti-tumor activities in RAS or BRAF mutant cancers." Cancer cell 28 (2015): 384-398.
  328. Google Scholar, Crossref, Indexed at

  329. Tolcher, A. W., J. C. Bendell, K. P. Papadopoulos and H. A. Burris III, et al. "A phase IB trial of the oral MEK inhibitor trametinib (GSK1120212) in combination with everolimus in patients with advanced solid tumors." Ann Concol 26 (2015): 58-64.
  330. Google Scholar, Crossref, Indexed at

  331. Grilley-Olson, J. E., P. L. Bedard, A. Fasolo and M. Cornfeld, et al. "A phase Ib dose-escalation study of the MEK inhibitor trametinib in combination with the PI3K/mTOR inhibitor GSK2126458 in patients with advanced solid tumors." Invest New Drugs 34 (2016): 740-749.
  332. Google Scholar, Crossref, Indexed at

  333. Bannoura, Sahar F., Md Uddin, Misako Nagasaka and Farzeen Fazili et al. "Targeting KRAS in pancreatic cancer: New drugs on the horizon." Cancer Metastasis Rev 40 (2021): 819-835.
  334. Google Scholar, Crossref, Indexed at

  335. McCubrey, James A., Linda S. Steelman, William H. Chappell and Stephen L. Abrams, et al. "Ras/Raf/MEK/ERK and PI3K/PTEN/Akt/mTOR cascade inhibitors: How mutations can result in therapy resistance and how to overcome resistance." Oncotarget 3 (2012): 1068.
  336. Google Scholar, Crossref, Indexed at

  337. Temraz, Sally, Deborah Mukherji and Ali Shamseddine. "Dual inhibition of MEK and PI3K pathway in KRAS and BRAF mutated colorectal cancers." Int J Mol Sci 16 (2015): 22976-22988.
  338. Google Scholar, Crossref, Indexed at

  339. Shimizu, Toshio, Anthony W. Tolcher, Kyriakos P. Papadopoulos and Muralidhar Beeram, et al. "The clinical effect of the dual-targeting strategy involving PI3K/AKT/mTOR and RAS/MEK/ERK pathways in patients with advanced cancer clinical effect of dual PI3K and MAPK pathways inhibitions." Clin Cancer Res 18 (2012): 2316-2325.
  340. Google Scholar, Crossref, Indexed at

  341. Tanaka, Masao, Suresh Chari, Volkan Adsay and Fernandez-Del Carlos Castillo, et al. "International consensus guidelines for management of intraductal papillary mucinous neoplasms and mucinous cystic neoplasms of the pancreas.Pancreatology 6 (2006): 17-32.
  342. Google Scholar, Crossref, Indexed at

  343. Tanaka, Masao, Carlos Fernández-del Castillo, Volkan Adsay and Suresh Chari, et al. "International consensus guidelines 2012 for the management of IPMN and MCN of the pancreas." Pancreatology 12 (2012): 183-197.
  344. Google Scholar, Crossref, Indexed at

  345. Sakitani M, Taki J., "Subtypes of intraductal papillary mucinous neoplasms of the pancreas and its prognosis with reference to early palliative care integrated with oncology" (2015):14: 1-34.
  346. Google Scholar, Crossref, Indexed at

  347. Tanaka, Masao, Carlos Fernández-del Castillo, Terumi Kamisawa and Jin Young Jang, et al. "Revisions of international consensus fukuoka guidelines for the management of IPMN of the pancreas." Pancreatology 17 (2017): 738-753.
  348. Google Scholar, Crossref, Indexed at

  349. Jang, Jin-Young, Yoon-Chan Park, Yoon Sup Song and Seung Eun Lee, et al. "Increased K-ras mutation and expression of S100A4 and MUC2 protein in the malignant intraductal papillary mucinous tumor of the pancreas." J Hepatobiliary Pancreat Surg 16 (2009): 668-674.
  350. Google Scholar, Crossref, Indexed at

  351. Amato, Eliana, Marco dal Molin, Andrea Mafficini and Jun Yu, et al. "Targeted next‐generation sequencing of cancer genes dissects the molecular profiles of intraductal papillary neoplasms of the pancreas." J Pathol 233 (2014): 217-227.
  352. Google Scholar, Crossref, Indexed at

  353. Lee, Ju-Han, Younghye Kim, Jung-Woo Choi and Young-Sik Kim, et al. "KRAS, GNAS and RNF43 mutations in intraductal papillary mucinous neoplasm of the pancreas: A meta-analysis." Springerplus 5 (2016): 1-12.
  354. Google Scholar, Crossref, Indexed at

  355. Z'graggen, Kaspar, Jaime A. Rivera, Carolyn C. Compton and Michael Pins, et al. "Prevalence of activating K-ras mutations in the evolutionary stages of neoplasia in intraductal papillary mucinous tumors of the pancreas." Ann Surg 226 (1997): 491.
  356. Google Scholar, Crossref, Indexed at

  357. Paini, Marina, Stefano Crippa, Stefano Partelli and Filippo Scopelliti, et al. "Molecular pathology of intraductal papillary mucinous neoplasms of the pancreas." World J Gasgtroenterol 20 (2014): 10008.
  358. Google Scholar, Crossref, Indexed at

  359. Schönleben, Frank, Wanglong Qiu, Karl C. Bruckman and Nancy T. Ciau, et al. "BRAF and KRAS gene mutations in intraductal papillary mucinous neoplasm/carcinoma (IPMN/IPMC) of the pancreas." Cancer Lett 249 (2007): 242-248.
  360. Google Scholar, Crossref, Indexed at

  361. Schönleben, Frank, Wanglong Qiu, Nancy T. Ciau and Daniel J. Ho, et al. "PIK3CA mutations in intraductal papillary mucinous neoplasm/carcinoma of the pancreas." Clin Cancer Res 12 (2006): 3851-3855.
  362. Google Scholar, Crossref, Indexed at

  363. Garcia-Carracedo, Dario, Andrew T. Turk, Stuart A. Fine and Nathan Akhavan, et al. "Loss of PTEN expression is associated with poor prognosis in patients with intraductal papillary mucinous neoplasms of the pancreas PTEN loss is associated with poor prognosis in IPMN." Clin Cancer Res 19 (2013): 6830-6841.
  364. Google Scholar, Crossref, Indexed at

  365. Vanhaesebroeck, Bart, and Dario R. Alessi. "The PI3K–PDK1 connection: More than just a road to PKB." Biochem J 346 (2000): 561-576.
  366. Google Scholar, Crossref, Indexed at

  367. Falkenburger, Björn H., Jill B. Jensen, Eamonn J. Dickson and Byung‐Chang Suh, et al. "Symposium review: Phosphoinositides: lipid regulators of membrane proteins." J Physiol 588 (2010): 3179-3185.
  368. Google Scholar, Crossref, Indexed at

  369. Kerr, William G. "Inhibitor and activator: Dual functions for SHIP in immunity and cancer." Ann N Y Acad Sci 1217 (2011): 1-17.
  370. Google Scholar, Crossref, Indexed at

  371. Pedicone, Chiara, Shea T. Meyer, John D. Chisholm and William G. Kerr, et al. "Targeting SHIP1 and SHIP2 in Cancer." Cancers 13 (2021): 890.
  372. Google Scholar, Crossref, Indexed at

  373. Stambolic, Vuk, Akira Suzuki, José Luis De La Pompa and Greg M. Brothers, et al. "Negative regulation of PKB/Akt-dependent cell survival by the tumor suppressor PTEN." Cell 95 (1998): 29-39.
  374. Google Scholar, Crossref, Indexed at

  375. Milella, Michele, Italia Falcone, Fabiana Conciatori and Ursula Cesta Incani, et al. "PTEN: Multiple functions in human malignant tumors." Front Oncol 5 (2015): 24.
  376. Google Scholar, Crossref, Indexed at

  377. Calleja, Véronique, Damien Alcor, Michel Laguerre and Jongsun Park, et al. "Intramolecular and intermolecular interactions of protein kinase B define its activation in vivo." PLoS biology 5 (2007): e95.
  378. Google Scholar, Crossref, Indexed at

  379. Calleja, Véronique, Michel Laguerre, Peter J. Parker and Banafshé Larijani, et al. "Role of a novel PH-kinase domain interface in PKB/Akt regulation: Structural mechanism for allosteric inhibition." PLoS Biol 7 (2009): e1000017.
  380. Google Scholar, Crossref, Indexed at

  381. Alessi, Dario R., Stephen R. James, C. Peter Downes and Andrew B. Holmes, et al. "Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and activates protein kinase Bα." Curr Biol 7 (1997): 261-269.
  382. Google Scholar, Crossref, Indexed at

  383. Stokoe, David, Leonard R. Stephens, Terry Copeland and Piers RJ Gaffney, et al. "Dual role of phosphatidylinositol-3, 4, 5-trisphosphate in the activation of protein kinase B." Science 277 (1997): 567-570.
  384. Google Scholar, Crossref, Indexed at

  385. Sarbassov, Dos D., David A. Guertin, Siraj M. Ali and David M. Sabatini, et al. "Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex." Science 307 (2005): 1098-1101.
  386. Google Scholar, Crossref, Indexed at

  387. Manning, Brendan D and Alex Toker. "AKT/PKB signaling: Navigating the network." Cell 169 (2017): 381-405.
  388. Google Scholar, Crossref, Indexed at

  389. Zhong, Hua, Kelly Chiles, David Feldser and Erik Laughner, et al. "Modulation of hypoxia-inducible factor 1α expression by the epidermal growth factor/phosphatidylinositol 3-kinase/PTEN/AKT/FRAP pathway in human prostate cancer cells: implications for tumor angiogenesis and therapeutics." Cancer Res 60 (2000): 1541-1545.
  390. Google Scholar, Crossref, Indexed at

  391. Jiang, Bing-Hua, Guoqiang Jiang, Jenny Z. Zheng and Zhimin Lu et al. "Phosphatidylinositol 3-kinase signaling controls levels of hypoxia-inducible factor." Cell Growth Diff (2001): 363-369.
  392. Google Scholar, Crossref, Indexed at

  393. Verma, Manasvi, Junhong Choi, Kyle A. Cottrell and Zeno Lavagnino, et al. "A short translational ramp determines the efficiency of protein synthesis." Nat Commun 10 (2019): 1-15.
  394. Google Scholar, Crossref

  395. Roux, Philippe P., and Ivan Topisirovic. "Signaling pathways involved in the regulation of mRNA translation." Mol Cell Biol 38 (2018): e00070-18.
  396. Google Scholar, Crossref, Indexed at

  397. Sonenberg, Nahum and Alan G. Hinnebusch. "Regulation of translation initiation in eukaryotes: Mechanisms and biological targets." Cell 136 (2009): 731-745.
  398. Google Scholar, Crossref, Indexed at

  399. Takatsuki Kiyoshi, Uchiyama T, Sagawa K and Yodoi J. "Adult T cell leukemia in Japan." (1977): 73-77
  400. Google Scholar, Crossref, Indexed at

  401. Uchiyama, Takashi, Junji Yodoi, Kimitaka Sagawa and Kiyoshi Takatsuki, et al. "Adult T-cell leukemia: Clinical and hematologic features of 16 cases." Blood 50 (1977): 481-492.
  402. Google Scholar, Crossref, Indexed at

  403. Hinuma, Yorio, Kinya Nagata, Masao Hanaoka and Masuyo Nakai et al. "Adult T-cell leukemia: Antigen in an ATL cell line and detection of antibodies to the antigen in human sera." Proc Natl Acad Sci 78 (1981): 6476-6480.
  404. Google Scholar, Crossref, Indexed at

  405. Seiki, Motoharu, Seisuke Hattori, Yoko Hirayama and Mitsuaki Yoshida, et al. "Human adult T-cell leukemia virus: Complete nucleotide sequence of the provirus genome integrated in leukemia cell DNA." Proc Natl Acad Sci 80 (1983): 3618-3622.
  406. Google Scholar, Crossref, Indexed at

  407. Katsuya, Hiroo, Lucy Cook, Aileen G. Rowan and Yorifumi Satou, et al. "Phosphatidylinositol 3-kinase-δ (PI3K-δ) is a potential therapeutic target in adult T-cell leukemia-lymphoma." Biomark Res 6 (2018): 1-4.
  408. Google Scholar, Crossref, Indexed at

  409. Ishikawa, Chie, Masachika Senba and Naoki Mori. "Effects of NVP‑BEZ235, a dual phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor, on HTLV-1-infected T-cell lines." Oncol Lett 15 (2018): 5311-5317.
  410. Google Scholar, Crossref, Indexed at

  411. Ishikawa, Chie and Naoki Mori. "The role of CUDC‐907, a dual phosphoinositide‐3 kinase and histone deacetylase inhibitor, in inhibiting proliferation of adult T‐cell leukemia.Eur J Haematolo 105 (2020): 763-772.
  412. Google Scholar, Crossref, Indexed at

  413. Daenthanasanmak, Anusara, Richard N. Bamford, Makoto Yoshioka and Shyh-Ming Yang, et al. "Triple combination of BET plus PI3K and NF-κB inhibitors exhibit synergistic activity in adult T-cell leukemia/lymphoma." Blood Adv 6 (2022): 2346-2360.
  414. Google Scholar, Crossref, Indexed at

  415. Yamaguchi, Kazunori, Tomoka Takanashi, Kentaro Nasu and Keiichi Tamai, et al. "Xenotransplantation elicits salient tumorigenicity of adult T‐cell leukemia‐derived cells via aberrant AKT activation." Cancer Sci 107 (2016): 638-643.
  416. Google Scholar, Crossref, Indexed at

  417. Watanabe, Tatsuro, Akemi Sato, Naomi Kobayashi-Watanabe and Naoko Sueoka-Aragane, et al. "Torin2 potentiates anticancer effects on adult T-cell leukemia/lymphoma by inhibiting mammalian target of rapamycin." Anticancer Res 36 (2016): 95-102.
  418. Google Scholar, Crossref, Indexed at

  419. Kawata, Takahito, Kohei Tada, Masayuki Kobayashi and Takashi Sakamoto, et al. "Dual inhibition of the mTORC 1 and mTORC 2 signaling pathways is a promising therapeutic target for adult T‐cell leukemia." Cancer Sci 109 (2018): 103-111.
  420. Google Scholar, Crossref, Indexed at

  421. Darwiche, Nadine, Ansam Sinjab, Ghada Abou‐Lteif and Mirella Bou Chedid, et al. "Inhibition of mammalian target of rapamycin signaling by everolimus induces senescence in adult T‐cell leukemia/lymphoma and apoptosis in peripheral T‐cell lymphomas." Int J Cancer 129 (2011): 1006-1017.
  422. Google Scholar, Crossref, Indexed at

  423. Daenthanasanmak, Anusara, Yuquan Lin, Meili Zhang and Bonita R. Bryant, et al. "Enhanced efficacy of JAK1 inhibitor with mTORC1/C2 targeting in smoldering/chronic adult T cell leukemia." Transl Oncol 14 (2021): 100913.
  424. Google Scholar, Crossref, Indexed at

  425. Yamada, Yasuaki, Kazuyuki Sugahara, Kazuto Tsuruda and Kazuo Nohda, et al. "Fas-resistance in ATL cell lines not associated with HTLV-I or FAP-1 production." Cancer Lett 147 (1999): 215-219.
  426. Google Scholar, Crossref, Indexed at

  427. Tomita, Mariko, Gregg L. Semenza, Canine Michiels and Takehiro Matsuda, et al. "Activation of hypoxia-inducible factor 1 in human T-cell leukaemia virus type 1-infected cell lines and primary adult T-cell leukaemia cells." Biochem J 406 (2007): 317-323.
  428. Google Scholar, Crossref, Indexed at

  429. Azran, Inbal, Yana Schavinsky-Khrapunsky and Mordechai Aboud. "Role of tax protein in human T-cell leukemia virus type-I leukemogenicity." Retrovirology 1 (2004): 1-24.
  430. Google Scholar, Crossref, Indexed at

  431. Boxus, Mathieu, Jean-Claude Twizere, Sébastien Legros and Jean-François Dewulf, et al. "The HTLV-1 tax interactome." Retrovirology 5 (2008): 1-24.
  432. Google Scholar, Crossref, Indexed at

  433. Harhaj, Edward William and Chou‐Zen Giam. "NF‐κB signaling mechanisms in HTLV‐1‐induced adult T‐cell leukemia/lymphoma." FEBSJ 285 (2018): 3324-3336.
  434. Google Scholar, Crossref, Indexed at

  435. Sakitani M. “Tax-dependent and tax-independent activation of NF-κB in adult T-cell leukemia/lymphoma (ATL).” Vox Propria 19(2020): 37-79.
  436. Google Scholar, Crossref, Indexed at

Google Scholar citation report
Citations: 3919

Molecular and Genetic Medicine received 3919 citations as per Google Scholar report

Molecular and Genetic Medicine peer review process verified at publons

Indexed In

 
arrow_upward arrow_upward