Next Article in Journal
Application of Electronic Health Record Text Mining: Real-World Tolerability, Safety, and Efficacy of Adjuvant Melanoma Treatments
Previous Article in Journal
Generalization of Deep Learning in Digital Pathology: Experience in Breast Cancer Metastasis Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Protein Kinase C (PKC) Isozymes as Diagnostic and Prognostic Biomarkers and Therapeutic Targets for Cancer

1
Center for Advanced Medical Innovation, Kyushu University, 3-1-1 Maidashi, Higashi-ku, Fukuoka 812-8582, Japan
2
Department of Urology, Graduate School of Medical Sciences, Kyushu University, Maidashi, Higashi-ku, Fukuoka 812-8582, Japan
3
Division of Biopharmaceutics and Pharmacokinetics, National Cerebral and Cardiovascular Center Research Institute, 6-1 Shinmachi, Kishibe, Suita, Osaka 564-8565, Japan
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(21), 5425; https://doi.org/10.3390/cancers14215425
Submission received: 17 October 2022 / Revised: 2 November 2022 / Accepted: 2 November 2022 / Published: 3 November 2022
(This article belongs to the Section Cancer Biomarkers)

Abstract

:

Simple Summary

Protein kinase C (PKC) isozymes play key roles in the proliferation, differentiation, survival, migration, invasion, apoptosis, and anticancer drug resistance of cancer cells. PKC isozymes are attractive therapeutic targets for cancer and have great potential as diagnostic and prognostic biomarkers for diagnosing cancers and for predicting disease-free survival and survival rates, respectively. This review discusses the potential of PKC isozymes as diagnostic and prognostic biomarkers and therapeutic targets for cancer.

Abstract

Protein kinase C (PKC) is a large family of calcium- and phospholipid-dependent serine/threonine kinases that consists of at least 11 isozymes. Based on their structural characteristics and mode of activation, the PKC family is classified into three subfamilies: conventional or classic (cPKCs; α, βI, βII, and γ), novel or non-classic (nPKCs; δ, ε, η, and θ), and atypical (aPKCs; ζ, ι, and λ) (PKCλ is the mouse homolog of PKCι) PKC isozymes. PKC isozymes play important roles in proliferation, differentiation, survival, migration, invasion, apoptosis, and anticancer drug resistance in cancer cells. Several studies have shown a positive relationship between PKC isozymes and poor disease-free survival, poor survival following anticancer drug treatment, and increased recurrence. Furthermore, a higher level of PKC activation has been reported in cancer tissues compared to that in normal tissues. These data suggest that PKC isozymes represent potential diagnostic and prognostic biomarkers and therapeutic targets for cancer. This review summarizes the current knowledge and discusses the potential of PKC isozymes as biomarkers in the diagnosis, prognosis, and treatment of cancers.

1. Introduction

Protein kinase-mediated phosphorylation of serine (S), threonine (T), and/or tyrosine (Y) residues in target proteins is involved in the activation or inactivation of intracellular signal transduction pathways. Protein kinase C (PKC) is a family of calcium- and phospholipid-dependent serine/threonine kinases. The PKC family consists of at least 11 isozymes and is classified into three subfamilies based on their structural characteristics and mode of activation: conventional or classic (cPKCs; α, βI, βII, and γ), novel or non-classic (nPKCs; δ, ε, η, and θ), and atypical (aPKCs; ζ, ι, and λ) (PKCλ is the mouse homolog of PKCι) PKC isozymes [1,2].
All PKCs consist of a regulatory and catalytic (kinase) domain. The regulatory region is divided into an autoinhibitory domain (pseudosubstrate) and two membrane-targeting domains (C1 and C2). The C1 and C2 domains bind to diacylglycerol (DAG) and Ca2+, respectively. The C3 and C4 domains in the catalytic region bind to ATP and its target substrate, respectively. The C1 domain mediates DAG-dependent translocation of cPKCs and nPKCs, but not of aPKCs, which contain a single C1 domain. cPKCs contain the calcium-sensitive C2 domain and bind to Ca2+, whereas nPKCs (contain an atypical C2-like domain) and aPKCs (without the C2 domain) do not. A phosphatidylserine (PS)-binding domain is not found in all PKCs, but PS, either alone or in combination with DAG and Ca2+, is essential for the phosphorylation of the target substrate [1,2]. The consensus phosphorylation site motifs for PKCs are (R/K)X(S/T), (R/K)(R/K)X(S/T), (R/K)XX(S/T), (R/K)X(S/T)XR/K, and (R/K)XX(S/T)XR/K, which clearly show that PKC substrates are typically rich in basic amino acids (arginine (R) and/or lysine (K)) [3]. PKC isozyme-specific substrates and their design methods have been extensively reviewed in previous articles [3,4,5].
PKC isozymes play key roles in the proliferation, differentiation, survival, migration, invasion, apoptosis, and anticancer drug resistance of cancer cells. Because of their high potential as therapeutic targets, many natural and synthetic PKC inhibitors have been developed and tested in clinical trials for cancer treatment (for review, see [6,7]). Furthermore, PKC isozymes also have great potential as diagnostic and prognostic biomarkers for diagnosing cancers and for predicting disease-free survival and survival rates (Figure 1). This review discusses the potential of PKC isozymes as diagnostic and prognostic biomarkers and therapeutic targets for cancer.

2. PKC Isozymes as Prognostic Biomarkers or Therapeutic Targets for Cancer

2.1. Bladder Cancer

Among the PKC isozymes, PKCα, βI, βII, δ, ε, η, and ζ have been observed in bladder cancer cells and tissues. PKCβI, βII, δ, and η are found mainly in early-stage bladder cancer, but their levels are reduced as cancer progresses. PKCα and ζ levels increase with increasing cancer stage [8,9,10].
In a large-scale multi-omics analysis, elevated expression of PKCα protein was associated with poor prognosis in patients with bladder cancer, in addition to increased expression of beclin, epidermal growth factor receptor (EGFR), annexin-1, and AXL proteins and downregulation of Src protein [11]. A previous study demonstrated that PKCα/β has a critical role in phospholipase Cε-mediated bladder cancer cell invasion and migration [12], and cell proliferation [13]. Furthermore, the expression of PKCα and nuclear factor kappa-B (NF-κB) in bladder cancer cells positively correlated with poor prognosis [14]. PKCα induced cellular resistance to apoptosis by stimulating NF-κB activation [14,15].
High PKCα activity, high netrin-1 expression, and low UNC5B expression enhanced the tolerance of bladder cancer cells to cisplatin, whereas the opposite expression pattern increased their sensitivity to cisplatin treatment [16]. Overexpression of tripartite motif 29 (TRIM29) upregulated the levels of cell survival-related proteins (e.g., cyclin and Bcl family) and inhibited cisplatin-mediated cell apoptosis in bladder cancer cells. However, its expression was downregulated following treatment with the PKC inhibitor staurosporine or the NF-κB inhibitor BAY 11-7082. These results indicate that TRIM29 inhibits drug-induced apoptosis in bladder cancer via the PKC/NF-κB signaling pathway [17]. Moreover, in patients treated with the anticancer drug adriamycin, high PKCα level is associated with a shorter recurrence-free period and higher drug resistance than low PKCα level [18]. However, PKCα inhibition induces apoptosis in bladder cancer cells by enhancing the activities of caspase-3 and poly (ADP-ribose) polymerase (PARP) [19]. These studies suggest that PKCα activity in bladder cancer may be a biomarker for poor prognosis and anticancer drug resistance and that PKCα inhibition may be a useful therapeutic option for bladder cancer.
In contrast, loss of aPKC (PKCι and ζ) expression in superficial bladder cancer is associated with a high recurrence rate and poor survival [20]. Treatment with the aPKC inhibitors ζ-Stat and 5-amino-1-2,3-dihydroxy-4-(methylcyclopentyl)-1H-imidazole-4-carboxamide (ICA-1), together with rapamycin, blocked bladder cancer progression [21].

2.2. Blood and Bone Marrow Cancers

Blood and bone marrow cancers can be divided into three major types: multiple myeloma (MM), leukemia, and lymphoma.

2.2.1. MM

MM is a type of bone marrow cancer. Very few studies have examined the role or function of PKC isozymes in MM. PKCβ has attracted immense attention as a therapeutic target in MM [22,23]; however, in clinical trials, treatment with the oral inhibitor enzastaurin showed no clinical benefit in patients with MM [24].

2.2.2. Leukemia

Based on the cell of origin, leukemia is classified as lymphocytic (lymphoid or lymphoblastic) or myeloid (myelogenous or myeloblastic) types and further divided into four types: acute lymphocytic leukemia (ALL), chronic lymphocytic leukemia (CLL), acute myeloid leukemia (AML), and chronic myeloid leukemia (CML).

ALL and CLL

Expression of PKCβ, γ, δ, and ζ was found in all patients with CLL, and that of PKCα, ε, and ι was variable, whereas PKCθ was not expressed [25]. Activated PKCα/βII (Thr638/641) was higher in patients with differentiated B-cell CLL compared to that in healthy controls [26]. The interaction of Rack1 and PKCα, but not PKCβ, was observed in two T-cell ALL-derived cell lines (Jurkat and CCRF-CEM). PKCα inhibition increased apoptosis in Rack1-overexpressing T-cell ALL cells following treatment with chemotherapeutic drugs [27]. In ALL, overexpression of PKCα did not affect cell proliferation, cell cycle, or activation of mitogen-activated protein kinases (MAPKs), but increased chemoresistance through Bcl-2 activation [28]. These studies suggest that PKCα may be closely associated with increased chemoresistance in lymphocytic leukemia.
Among the PKC isozymes, PKCβ is considered to be a useful therapeutic target for lymphocytic leukemia as it participates in cell survival and proliferation [29,30,31], resistance to apoptosis [32], and chemoresistance induced by stromal cells, which are key components of the lymphocytic leukemia microenvironment [30,33]. However, PKCβ-specific inhibitors have failed to show significant clinical benefits in patients with lymphocytic leukemia [7].
In addition, PKCε [34] and PKCδ have been reported to mediate leukemic cell survival [35] and cell sensitivity to anticancer drugs induced by PKCζ overexpression [36]. Furthermore, links between PKCδ and Notch2 [37] and PKCθ and Notch1 signaling in leukemic cells [38] and aPKCλ/ι-mediated transformation of B-cell progenitors (can generate B-cell ALL) by BCR-ABL [39] have been reported.

AML and CML

PKCα activation is associated with poor survival in patients with AML [40]. PKCα activation also enhanced resistance to chemotherapy in AML cells through Bcl-2 phosphorylation [41] and extracellular-signal-regulated kinase 1/2 (ERK1/2) and Akt activation [42]. PKCα inhibition enhanced selenite-induced apoptosis of the acute promyelocytic leukemia cell line NB4 [42].
PKCε was found to be markedly overexpressed in patients with AML and positively correlated with reduced complete remission, disease-free survival, and enhanced resistance to the chemotherapeutic agent daunorubicin through P-glycoprotein (P-gp)-mediated drug efflux [43]. PKCε overexpression protects AML cells from mitochondrial reactive oxygen species (ROS)-inducing agents. However, PKCε deletion reduced patient-derived AML cell survival and disease onset in an AML mouse model [44].
There was a significant association between reduced PKCδ levels and relapse in patients with AML [45]. PKCδ appears to be involved in stimulating anticancer drug-mediated apoptosis through caspase-3 activation [46,47], phosphorylation of eukaryotic initiation factor-α [45], and downregulation of heterogeneous nuclear ribonucleoprotein K [48].
A recent phase III trial showed that treatment with midostaurin (also known as PKC412; CGP 41251) with standard chemotherapy significantly prolonged overall and event-free survival in patients with mutant FLT3-positive AML [49]. Although midostaurin was originally developed as a PKC inhibitor, its clinical benefits are mainly achieved via tyrosine kinase inhibition [50]. Midostaurin has been approved by the FDA for the treatment of newly diagnosed adult patients with mutant FLT3-positive AML and adult patients with systemic mastocytosis with associated hematological neoplasm or mast cell leukemia, which is an aggressive subtype of AML [7].
The addition of the PKC inhibitor staurosporine increases the sensitivity of imatinib-resistant CML to imatinib by inducing G2/M phase arrest through PKCα-dependent CDC23 downregulation [51]. PKCβ overexpression in CML cells also enhances resistance to imatinib through arachidonate 5-lipoxygenase (Alox5) signaling. Alox5 levels were increased in both bone marrow biopsies and CD34+ cells derived from patients with imatinib-resistant CML. In contrast, prolonged survival was observed in CML mice treated with imatinib in combination with the PKCβ inhibitor LY333531 [52]. PKCη was upregulated in samples from patients with CML with BCR-ABL-independent imatinib resistance or CML stem cells, leading to sustained RAF/MEK/ERK signaling following imatinib treatment. Combined treatment with imatinib and the MEK inhibitor trametinib prolonged survival in mouse models of BCR-ABL-independent imatinib-resistant CML [53]. In addition, aPKCλ/ι may be a potential therapeutic target for treating tyrosine kinase inhibitor (TKI)-resistant CML [39].

Myelodysplastic Syndromes (MDSs)

MDSs are a heterogeneous group of hematopoietic stem cell disorders and frequently evolve into AML [54,55]. Nuclear translocation of PKCα induced erythropoiesis in patients with low-risk MDS following treatment with the immunomodulatory drug lenalidomide [56]. Furthermore, the PI-PLCβ1/cyclin D3/PKCα signaling pathway was associated with iron-induced oxidative stress and ROS production in MDS patients [57].

2.2.3. Lymphoma

Lymphoma begins in the T or B cells of the lymphatic system and is classified into two major subtypes: Hodgkin and non-Hodgkin lymphoma (NHL). PKC isozyme analysis using reactive lymphoid tissues, human B-cell lymphoma, and human lymphoma cell lines revealed that PKCα, βII, γ, and δ were expressed in B-cell malignancies. Compared to other types of lymphomas, Burkitt’s lymphomas overexpress PKCα. In Burkitt’s lymphoma, the overall survival was higher in PKCγ-positive cases than in PKCγ-negative cases [58]. PKCζ, but not cPKC, is involved in the regulation of telomerase activity in Burkitt’s lymphoma cells [59].
In follicular lymphomas, PKCβII is overexpressed, mainly in the mantle and marginal zones. PKCβII expression was also found in most angioimmunoblastic T-cell lymphomas, lymphoblastic T-cell lymphomas, and marginal zone/mucosa-associated lymphoid tissue lymphomas, although the pattern of expression was very heterogeneous. However, PKCβII expression was not observed in Hodgkin’s disease or anaplastic large-cell lymphoma [60]. Higher PKCβII expression was noted in human immunodeficiency virus-infected patients than in uninfected patients with diffuse large B-cell lymphoma (DLBCL), which is the most common subtype of NHL [61]. In DLBCL, higher PKCβ expression was found in the activated B-cell-like subtype than in the germinal center B-cell-like subtype, and its elevated levels were associated with worse survival in both subtypes [62]. PKCβII expression in DLBCL was correlated with poor overall and progression-free survival in patients treated with cyclophosphamide, doxorubicin (DOX), vincristine, and prednisolone [63]. PKCβII expression was associated with worse 5-year event-free and overall survival in patients with nodal DLBCL, especially in patients with low-risk International Prognostic Index [64,65,66]. Based on these reports, PKCβII is regarded as a marker for poor prognosis and a chemotherapeutic target for lymphoid malignancies.
In lymphoma, PKCδ activation stimulates anticancer drug-mediated apoptosis through caspase-3 activation [67,68], JNK activation [69], or phosphorylation and activation of lysosomal acidic sphingomyelinase [70]. The PKCζ/mammalian target of rapamycin (mTOR) pathway may also be a therapeutic target for rituximab-mediated treatment of follicular lymphoma [71].

2.3. Brain Cancer (Glioblastoma)

Glioblastoma is a high-grade astrocytoma and the most malignant type of brain tumor. Astrocytoma malignancies are positively correlated with progesterone receptor (PR) and PKCα levels as well as with the intracellular colocalization of these proteins. Patients with astrocytoma grades III and IV with low expression of PGR and PRKCA mRNA showed higher survival than those with high expression [72]. Treatment with mTOR inhibitors (rapamycin, temsirolimus, torin-1, and PP242) reduces glioblastoma progression by reducing invasion, migration, and matrix metalloproteinase (MMP) activity (MMP2 and MMP9) through the reduction of PKCα and NF-κB signaling pathways [73]. Furthermore, PKCα/phosphoinositide 3-kinase (PI3K) signaling pathways increase astrocytoma invasion by downregulating low-density lipoprotein receptor-related protein [74]. Overexpression of the long noncoding RNA TCONS_00020456, which targets the Smad2/PKCα axis, reduced glioma cell proliferation, migration, and invasion and inhibited epithelial–mesenchymal transformation and glioma progression in vivo [75]. Activation of the lysophosphatidic acid receptor LPA1 induces PKCα translocation to the nucleus, inhibits the LPA1/PKCα axis, and reduces glioblastoma growth and progression [76,77].
Although PKCα is a therapeutic target for glioblastoma, a previous study showed no clinical benefits in patients with high-grade gliomas following treatment with the antisense oligonucleotide aprinocarsen directed against PKCα [78]. However, a recent study suggested that combination therapy with JAK2 (AZD1480) and a PKCα inhibitor (erlotinib or osimertinib) induced apoptosis of glioblastoma, which is the most malignant and aggressive form of astrocytoma, in both flank and in patient-derived orthotopic xenograft models, indicating that PKCα and JAK2 may be therapeutic targets for glioblastoma [79]. Interestingly, in vitro experiments using U87MG cells showed that loss of PKCα proteins inhibited cell growth or survival, but the same effects were not obtained by inhibiting PKCα activity, indicating that ATP-competitive inhibitors of PKCα may have little or no therapeutic effect in glioblastoma [80].
PKCι is associated with cell proliferation [81,82,83], survival [84], invasion [81,83], apoptosis [82], and anticancer resistance [85] in glioblastomas. PKCι is overexpressed and activated in patient-derived glioblastoma stem-like cells compared to normal neural stem cells and normal brain lysates [86]. Glioblastoma cell proliferation depends on the PI3K/PKCι/CDK7/CDK2 pathway [82], and cell survival depends on the PI3K/PDK1/PKCι/BAD pathway [87]. Moreover, elevated PKCι level increases resistance to cisplatin in glioblastoma cells by suppressing GMFβ/p38 MAPK signaling [85] and induces glioblastoma motility by coordinating the formation of a single leading-edge lamellipod [81]. These results demonstrated that PKCι may be an important therapeutic target for glioblastoma.
Elevated PKCι activation in glioblastoma cells increased their sensitivity to PKCι inhibitors, but low PKCι activation resulted in both Src activation and sensitivity to Src inhibitors. The combination of PKCι and Src inhibitors prolonged survival beyond that of either drug alone [84]. The combination of PKCι inhibitors ICA-1 and temozolomide also decreased the invasion of glioblastoma cell lines and reduced glioblastoma growth and volume in mice [83]. Furthermore, combined treatment with ICA-1 and tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) stimulated caspase-3-mediated apoptosis in glioblastoma cells by downregulating PKCι and c-Jun [88].
As mentioned above, PKCα activation leads to increased proliferation and decreased apoptosis of glioblastoma cells. However, activated PKCδ has the opposite effect resulting in decreased proliferation and increased apoptosis [89,90,91] through Bcl-2 phosphorylation [91] or Akt (also called PKB) inhibition [90].
PKCε overexpression was found in primary pediatric anaplastic astrocytoma (grade III) tumor samples as well as in glioblastoma multiforme (grade IV) and gliosarcoma tumor samples, but not in pilocytic astrocytomas (grade I) [92]. PKCε inhibition decreased the expression of Beclin1, Atg5, and PI3K in glioma cells and increased the expression of the autophagy-related proteins mTOR and Bcl-2. PKCε knockdown also reduced the adhesion of glioblastoma cells by decreasing total focal adhesion kinase (FAK) protein levels and its phosphorylation [93].
In addition to PKCδ and ε as potential therapeutic targets for glioblastoma, PKCζ may be a therapeutic target for glioblastoma cell migration and invasion [94], and PKCη may be a therapeutic target for glioblastoma proliferation [95,96].

2.4. Breast Cancer

Several PKC isozymes, including PKCα, β, δ, ε, ζ, η, ζ, θ, and λ, have been identified in breast cancer. PKCα, βI, and βII levels in breast cancer specimens and PKCβII levels in HER2-positive cancers are higher than those in adjacent normal breast tissues [97]. A previous study reported enhanced PKCε levels in high histologic grade and human epidermal growth factor receptor-2 (HER2/ErbB2)-positive, estrogen receptor (ER)-negative, and PR-negative breast cancers [98], whereas another study suggested that PKCε is downregulated in all cancer stages, molecular subtypes, metastatic and nonmetastatic groups, and patients with or without anticancer drug treatment compared to healthy controls [99].
PKCα is closely associated with poor survival in patients with breast cancer and increased anticancer resistance. In fact, poorer survival was observed in patients with PKCα-positive breast cancer than in those with PKCα-negative breast cancer. PKCα levels are positively associated with estrogen and PR negativity, cancer grade, and proliferative activity [100]. In endocrine-resistant and triple-negative breast cancer cell lines, PKCα plays a key role in maintaining their migratory and invasive phenotypes through FOXC2-mediated repression of p120-catenin [101].
PKCα may also be used as a marker for estrogen resistance because of the positive association between high PKCα levels and enhanced resistance to antiestrogen hormonal therapy (e.g., tamoxifen) [102,103]. PKCα levels are positively correlated with triple-negative breast cancers that are characterized by a lack of ER, PR, or ErbB2 expression [104,105], and there is an inverse relationship between PKCα levels and ERα expression [106]. Furthermore, PKCα showed relatively higher basal activity in drug-resistant MCF-7/ADR cells than in drug-sensitive MCF-7 cells. Inhibition of PKCα activity improves intracellular accumulation of DOX in MCF-7/ADR cells [107].
Overexpression of the Notch1 receptor and its ligand Jagged-1 is associated with poor survival in patients with ErbB-positive breast cancer [108,109] and increased trastuzumab resistance [110]. However, activated PKCα attenuates Jagged-1-mediated Notch1 activity in ErbB2-positive breast cancer and restores trastuzumab resistance, suggesting that PKCα activity may be a potential prognostic marker for low Notch activity and increased trastuzumab sensitivity in ErbB2-positive breast cancer [111].
Combined treatment with a PKC inhibitor and all-trans-retinoic acid (ATRA) reduced the growth, self-renewal, and frequency of cancer stem cells (CSCs) in a retinoic acid receptor (RAR) signaling-dependent manner. Low PKCα and high RAR levels were associated with significantly increased relapse-free survival (RFS) in patients with ER-negative breast cancer [112]. However, another study reported that PKCα overexpression promotes RARα expression levels in breast cancer cells following ATRA treatment, and increased RARα leads to ATRA sensitization through AP1 trans-repression [113].
ErbB2 entry into the endocytic recycling compartment stimulated by PKCα and PKCδ [114], or PKCδ-mediated Src activation, promotes ErbB2-induced mammary tumorigenesis [115]. Furthermore, human breast CSCs efficiently formed tumor xenografts in nude mice; however, their tumorigenesis was markedly reduced by PKCδ inhibition. In the mesenchymal CSC-like MCF10C cell line (M3), which is derived from MCF10A (M1) cells, PKCδ inhibition blocked tumor spheroid formation [116]. These data indicate that PKCδ is associated with mammary tumorigenesis and may be a predictive marker.
In contrast, in breast cancer samples from patients, high PKCδ and PKCα expression was correlated with endocrine responsiveness and ER negativity, respectively. A longer duration of endocrine response is observed in patients with a PKCδ(+)/PKCα(−) than the PKCδ(+)/PKCα(+) phenotype, indicating that PKCδ may be useful for predicting the response to antiestrogen therapy [117]. Interestingly, AD198 (a DOX analog)-induced apoptosis is PKCδ-dependent [118], but PKCδ in normal murine mammary gland cells increased resistance against AD198-mediated cell death through Akt and NF-κB survival pathways [119].
The high PKCζ group exhibits poorer prognosis, including advanced clinical stage, more lymph node involvement, larger tumor size, and lower disease-free and overall survival rates, compared to the low PKCζ group [120]. Moreover, PKCζ levels are higher in invading tissues than in non-invading tissues and are more abundant in ductal tissues than in lobular tissues. Its invasive behavior is induced through the Ras-related C3 botulinum toxin substrate 1 (Rac1) and Ras homolog gene family member A (RhoA) pathways. These results suggest that PKCζ may be used as an indicator of bladder cancer invasion [121] and a prognostic marker for breast cancer.
PKCθ-induced phosphorylation of Fra-1 stimulates the migration of breast cancer cells, and phosphorylated Fra-1 expression is enriched at the invasive front of human breast cancer cells. Furthermore, PKCθ is positively associated with MMP1 mRNA expression in human breast cancer samples [122]. PKCθ is enriched in circulating tumor cells in patients with triple-negative breast cancer brain metastases. Nuclear PKCθ-positive phenotype, together with cell surface vimentin-positive and ABCB5-positive phenotypes, a CSC-like marker associated with therapeutic resistance, is found in a higher proportion in brain metastases of patients with breast cancer than in primary breast tumors, indicating an association between PKCθ and cancer metastasis [123]. Enhanced PKCθ levels in triple-negative breast cells activate growth factor-independent growth, anoikis resistance, and migration [124]. Therefore, PKCθ upregulation may be used as a marker for predicting migratory and invasive behaviors in breast cancer cells.
Pal’s group reported that PKCη may serve as a potential biomarker for breast cancer malignancy because of higher expression of PKCη in malignant cells than in non-tumorigenic or pre-malignant cells, and they also reported a positive correlation between PKCη levels and increased breast cancer cell growth or clonogenic survival [125]. Increased PKCη expression in post-chemotherapy biopsies of patients with advanced and aggressive breast cancers was correlated with poor survival, showing that PKCη may also be an indicator of poor survival and a predictor of the effectiveness of anticancer treatment in patients with breast cancer [126,127].
Patients with late-stage (stage III–IV) breast cancer with high PKCλ, c-Met, and ALDH1A3 levels showed a poorer prognosis than patients with low PKCλ, c-Met, and ALDH1A3 levels. Treatment with the c-Met inhibitor foretinib and PKCλ inhibitor auranofin significantly suppressed cell viability and tumor-sphere formation mediated by ALDH1-positive breast CSCs in late-stage basal-like breast cancer. These results suggest that c-Met and PKCλ cooperatively induce poor prognosis in breast cancer [128,129]. Similarly, PKCλ and GLO1 cooperatively promote cell survival in ALDH1-positive breast CSCs, but their inhibition decreases cell viability and tumor-sphere formation [130].
High PKCε levels are associated with shorter disease-free survival in patients with ER-negative breast cancer than in those with ER-positive breast cancer. Although a correlation between PKCε and claudin 1, which is activated by the ERK signaling pathway, was identified in ER-negative cancer, claudin 1 levels are not a prognostic indicator of disease recurrence or survival [131]. Moreover, PKCε-induced activation of TRIM47 stimulates NF-κB signaling, resulting in enhanced breast cancer proliferation and resistance to endocrine therapy [132]. PKCε overexpression in MCF-7 cells increases cell survival by inhibiting apoptosis and inducing autophagy [133]. These results indicate that PKCε is a prognostic marker and therapeutic target in breast cancer.

2.5. Colorectal (Colon) Cancer (CRC)

PKCα is involved in cell proliferation, migration, and survival [134] and enhances drug resistance [135] in colon cancer. In colon cancer SW620 cells, PKCα stimulates TF/VIIa/PAR2-induced cell proliferation, migration, and survival through its downstream signaling pathways, ERK1/2/NF-κB [134] and ERK1/2/c-Jun/AP-1 [136]. Mitotic checkpoint kinase Mps1 (also known as TTK) activates the PKCα/ERK1/2 pathway but inhibits the PI3K/Akt pathway, resulting in the promotion of cell proliferation in colon cancer HT-29 and SW480 cells [137].
PKCα activation inhibited DOX-induced apoptosis in HCT15/DOX cells through scavenging of ROS and inhibition of PARP cleavage, whereas siRNA-mediated PKCα knockdown induced apoptosis [135]. Furthermore, PKCα inhibition enhanced resveratrol-induced apoptosis of HT-29 cells [138].
In contrast, the anticancer action of PKCα has been reported in CRC cells. PKCα increased IL12/GM-CSF-mediated M1 polarization of tumor-associated macrophages (TAMs) through the MKK3/6-P38 signaling pathway [139]. Furthermore, PKCα activation inhibited β-catenin-induced transcription and expression of cyclin D1 and c-myc, which are known targets of β-catenin, resulting in the reduced growth of CRC cells [140]. PKCα-deficient ApcMin/+ mice developed a more aggressive histopathological phenotype and had higher mortality than PKCα+/+ or PKCα+/– mice [141]. PKCα downregulation is observed at a higher frequency in tissues from advanced CRC stages than in the corresponding normal mucosa [142]. A PKCα mutation was found in CRC samples, but PKCα activation triggered CRC cell death [143]. Interestingly, low PKCα and high Kirsten rat sarcoma viral oncogene homolog (KRAS) expression are associated with a relatively poor prognosis in patients with CRC. PKCα expression in patients decreased in the following order: poorly differentiated < moderately differentiated < well-differentiated adenocarcinoma. However, KRAS levels are correlated with the degree of CRC differentiation [144]. These studies suggest that PKCα may be a potential drug target for CRC treatment. However, further studies are needed to clarify the role of PKCα in CRC cells.
Combined treatment with an atypical PKC inhibitor (ICA-I or ζ-Stat) and thymidylate synthase inhibitor 5-FU synergistically reduced the viability of CRC cells and induced apoptosis and DNA damage [145]. Enhanced PKCζ expression was found in human CRC tissues and cells and correlated with reduced AMPK activation and increased mTOR complex 1 (mTORC1) activation. Silencing of PKCζ inhibited HT-29 cell proliferation via AMPK activation [146]. However, activation of PKCζ inhibited TRAIL-induced apoptosis by regulating survivin levels [147]. Furthermore, PKC-ζ activation increased abnormal growth, proliferation, and migration of metastatic LOVO colon cancer cells via the PKC-ζ/Rac1/Pak1/β-catenin pathway [148]. Phosphorylated PKCζ/λ expression was also higher in colorectal adenocarcinomas than in adenomas. PKCζ/λ overexpression is associated with tumorigenesis in colorectal adenocarcinoma, but PKCζ/λ downregulation is associated with poor prognosis [149]. These results indicate that PKCζ is a useful target for the treatment of CRC.
Dowling’s group reported that PKCβII acts as a tumor suppressor in CRC and that decreased PKCβII level is associated with poor survival outcomes [150]. However, Spindler’s group reported that an increased level of PKCβII is associated with poor prognosis [151]. PKCβI and PKCβII activation increases CRC carcinogenesis and proliferation rates [152]. In COLO205-S cells, PKCβ inhibition increased cell apoptosis through the inactivation of Akt and glycogen synthase kinase-3β (GSK3β) [153].
PKCδ suppresses CRC growth through the activation of p21Waf1/Cip1 and p53 [154] but inhibits 5-FU-induced CRC apoptosis [155]. Moreover, PKCδ activation induces CRC cell motility and metastasis via enhanced B7-H4, which plays an important role in cancer growth and immunosuppression. Enhanced expression of PKCδ and B7-H4 is associated with moderate/poor differentiation, lymph node metastasis, and advanced Dukes’ stage [156]. In addition, the activation of PKCδ/NF-κB signaling increases CRC growth, whereas its inhibition results in CRC apoptosis through extrinsic/intrinsic pathways [157]. Increased nuclear translocation of PKCδ in CRC is also associated with worse prognosis [158].
Furthermore, Du’s group suggested that PKCι may serve as a novel therapeutic target for CRC because its inhibition reduces epithelial–mesenchymal transition (EMT), migration, and invasion of CRC cells by suppressing the Rac1-JNK pathway [159]. PKCλ/ι is a key regulator of the interferon pathway. Low PKCλ/ι levels correlate with enhanced interferon signaling and good prognosis in patients with CRC [160].

2.6. Gastric (Stomach) Cancer

PKCα is overexpressed in gastric cancer cells and tissues [161,162,163]. PKCα protein overexpression is significantly correlated with age, histologic type, tumor differentiation, depth of invasion, angiolymphatic invasion, pathologic stage, and distant metastasis in gastric cancer [161,162]. Furthermore, PKCα levels were higher in the vincristine-resistant human gastric cancer cell line SGC7901/VCR than in the non-vincristine-resistant cell line SGC7901. PKCα, but not PKCβI, βII, or γ, plays a role in multidrug resistance of SGC7901/VCR cells [163,164]. In HER2-negative advanced gastric cancer, PHD finger protein 8 (PHF8) positively correlates with PKCα, and high PHF8 and PKCα levels are significantly associated with poor clinical outcome [165].
Patients with gastric cancer with high PKCι levels showed lower overall survival compared to those with low PKCι levels [166]. Overexpression of circular RNA of PKCι is positively correlated with poor prognosis in patients with gastric cancer. In vitro experiments revealed that its overexpression promotes proliferation and invasion and reduces apoptosis of gastric cancer cells [167]. Moreover, stathmin 1 expression was significantly associated with gender and poorly differentiated gastric cancer. Furthermore, stathmin 1 expression was significantly correlated with activation-induced cytidine deaminase and PKCι levels [168]. The recurrence of gastric cancer following curative gastrectomy was increased in patients with PKCλ/ι overexpression [169].
These results suggest that PKCα and PKCι may serve as potential prognostic indicators and therapeutic targets for gastric cancer.

2.7. Head and Neck Squamous Cell Carcinoma (HNSCC)

HNSCC develops in the mucosal epithelium of the oral cavity, pharynx, and larynx [170], and several PKC isozymes, such as PKCα, β, γ, ε, θ, ι, and ζ are found in HNSCC [171,172,173].
PKCα overexpression occurs more frequently in younger (≤45 years) than older (>45 years) patients with oral tongue SCC (OTSCC). PKCα upregulation is associated with a negative history of alcohol and tobacco consumption. Both overall survival and disease-free survival are impaired in young patients with PKCα overexpression [174]. Furthermore, CC-chemokine receptor 7 and PKCα overexpression in HNSCC are significantly correlated with both cervical lymph node metastasis and clinical stage [175]. Another study also suggested that high PKCα expression is associated with a significantly higher probability of recurrence or death [176].
High levels of autophagy-suppressive circPARD3 are associated with malignant progression and poor prognosis in patients with laryngeal SCC (LSCC). CircPARD3 inhibits autophagy and promotes LSCC cell proliferation, migration, invasion, and chemoresistance through the PKCι/Akt/mTOR pathway [173]. In oral SCC (OSCC), PKCλ/ι expression is positively correlated with malignancy and progression-free survival [177].
High nuclear expression of PKCθ [178] or PKCβII [172] was significantly associated with poor overall survival and rapid recurrence in patients with OSCC, indicating that their nuclear expression can be a potential prognostic marker in patients with OSCC. Furthermore, the expression of CXCR-4, PKCδ, and CD133 is high in poorly differentiated and lymph node metastasis-positive cases of OSCC. CXCR4+/CD133+ and CXCR4+/PKCδ+ double-positive cases show poor survival [179].

2.8. Liver Cancer (Hepatocellular Carcinoma)

PKCα levels were higher in biopsy and surgical specimens of hepatocellular carcinoma (HCC) than in adjacent non-cancerous liver tissues [180]. PKCα expression correlated with tumor size and TNM stage. Patients with high PKCα expression showed shorter survival rates than those with low PKCα expression [181]. Inhibition of PKCα expression reduced several migration/invasion-related genes (e.g., MMP1, u-PA, u-PAR, and FAK) in both HA22T/VGH and SK-Hep-1 cell lines. Furthermore, PKCα inhibition decreased cyclin D1 levels and increased the levels of p53 and p21WAF1/CIP1, resulting in a decreased growth rate of HCC [182]. Enhanced expression of the retinoblastoma protein (RB)-binding transcription factor E2F1 transactivates cell-cycle-related factors and promotes HCC proliferation by activating PKCα [183]. Furthermore, PKCα stimulates dual oxidase 2 (DUOX2)-mediated ROS generation at the post-transcriptional level. DUOX2 inhibition blocked PKCα-induced activation of the Akt/MAPK pathways, as well as HCC cell proliferation, migration, and invasion [184]. A recent study reported that PKCα induces immune evasion and anti-PD1 tolerance by stimulating the zinc finger protein 64/macrophage colony-stimulating factor axis that transforms macrophages to the M2 phenotype to drive immune escape and anti-PD1 tolerance [185].
Suppression of the PKCδ/p38 MAPK pathway induced NF-κB-mediated inhibition of HCC progression [186] and attenuated phosphorylation of heat shock protein 27 that correlates with HCC progression [187]. Blockage of the PKCδ/p38 MAPK/nuclear factor erythroid 2-related factor (Nrf2) pathway also reduced the expression of heme oxygenase-1, which inhibits HCC cell death [188]. In addition, PKCδ triggers HCC progression by increasing mitochondrial ROS generation and HSP60 oxidation and inhibiting RAF kinase inhibitor protein, a negative regulator of MAPK [189]. Hypoxia induces HIF-2α-mediated activation of CUB domain-containing protein 1 (CDCP1) and phosphorylation of PKCδ, which is a downstream factor of CDCP1, leading to stimulation of HCC cell invasion. In fact, CDCP1 expression increases progressively with HCC tumor grade and is negatively correlated with disease-free survival [190]. These studies indicate that PKCδ is a potential prognostic biomarker for HCC.
PKCλ/ι is regarded as a tumor suppressor in HCC. PKCλ/ι levels negatively correlate with HCC histological tumor grade. PKCλ/ι inhibition promotes HCC progression by inducing autophagy, ROS production, and Nrf2 activation [191,192]. Furthermore, PKCβII and PKCθ are downregulated in HCC tissues. Reduced levels of PKCβII and PKCθ are associated with HBV infection and HCC grade, respectively [193]. PKCβ expression was found to be lower in the liver tissues of patients with HCC than in non-tumorous liver tissues [194]. However, another study reported that PKCβ is upregulated in HCC cell lines. Its upregulation increases the migration and invasion of HCC cells [195]. In addition, PKCη expression is downregulated in HCC tissues, and this reduction is associated with poor long-term survival of patients with HCC [196].

2.9. Lung Cancer

The two main types of lung cancers are small-cell lung cancer (SCLC) and non-small-cell lung cancer (NSCLC), which are further divided into adenocarcinoma, SCC, and large-cell carcinoma. The roles of PKC isozymes vary in SCLC and NSCLC. For example, following DOX treatment, NSCLC cells showed increased resistance to DOX through PKCα-mediated phosphorylation of Ral-interacting protein (RLIP76) compared to SCLC cells. Depletion of PKCα results in higher growth inhibition in NSCLC cells than in SCLC cells [197]. Phorbol 12-myristate 13-acetate (PMA), also known as 12-O-tetradecanoylphorbol 13-acetate (TPA), induces JNK activation in NSCLC, but not SCLC cells. The absence of JNK activation in PMA-treated SCLC cells was related to the absence of PKCε [198].
PKCα is highly expressed in NSCLCs, and its expression is higher in adenocarcinomas than in SCCs [199]. High PKCα/Rab37/tissue inhibitor of metalloproteinase-1 (TIMP1) expression profile correlated with worse progression-free survival in patients with lung cancer. PKCα-mediated Rab37 phosphorylation stimulated lung cancer cell motility [200]. In lung adenocarcinomas with EGFR mutation, PKCα activation plays a key role in the activation of the Akt/mTORC1 signaling pathway, which is involved in cell survival, growth, and proliferation [201]. PKCα also impairs TRAIL-induced apoptosis in H1299 NSCLC cells by activating the GSK3β/NF-κB pathway, whereas TRIM21 inhibits the activation of NF-κB by GSK3β [202]. Blocking the PKCα/ERK1/2 axis suppresses the proliferation and metastasis of human lung adenocarcinoma A549 cells [203]. In addition, erlotinib-resistant NSCLC cell line H1650-M3 showed substantial upregulation of PKCα and downregulation of PKCδ. Conversely, pharmacological inhibition or RNA interference-mediated depletion of PKCα sensitized H1650-M3 cells to erlotinib [204].
Based on these results, PKCα is regarded as a potential therapeutic target for NSCLC; however, treatment with PKCα-targeted inhibitors has yielded unsatisfactory clinical results [7,205]. Furthermore, a recent study suggested that among the PKC isozymes, high expression of PKCα and the phosphorylation state of PKCα, β, and δ showed the strongest positive correlation with RFS in patients with operable lung adenocarcinomas [206]. Hill’s group also demonstrated that PKCα suppresses KRAS-mediated lung tumor formation by activating the p38 MAPK/TGFβ pathway [207].
In A549 cells, 12-deoxyphorbol esters induce growth arrest and apoptosis by activation of the PKCδ/PKD/ERK pathway [208]. PKCδ activation induced morphological changes and migration of A549 cells by increasing tumor necrosis factor-α (TNF-α)-induced claudin-1 expression [209]. The PKCδ/midkine axis induces hypoxic proliferation and differentiation of A549 cells [210]. Moreover, suppression of the EGFR/PKCδ/NF-κB pathway induced imipramine-triggered anti-NSCLC effects in both in vitro and in vivo models [211]. Interaction of PKCδ with procollagen-lysine,2-oxoglutarate 5-dioxygenase 3 (PLOD3) activates caspase-2 and -4-dependent apoptosis through endoplasmic reticulum stress-induced inositol-requiring enzyme 1α activation and downstream unfolded protein response pathway [212]. Resistance to EGFR TKIs has been observed in EGFR-mutant NSCLC, and nuclear translocation of PKCδ is associated with the response of patients with NSCLC to TKIs. Combined inhibition of PKCδ and EGFR results in a marked regression of resistant NSCLC tumors with EGFR mutations [213]. These results show that PKCδ is involved in cell survival, antiapoptosis, and anticancer drug resistance in NSCLC and thus represents a potential therapeutic target for NSCLC.
Higher expression of PKCε was detected in primary human NSCLC tissue than in the normal lung epithelium [214]. PKCε plays an important role in KRAS-mediated tumorigenesis. Induction of lung tumorigenesis by the carcinogen benzo[a]pyrene, which induces mutations in KRAS, was markedly reduced in PKCε-knockout mice [215]. Moreover, PKCɛ is required for NSCLC cell survival and tumor growth. Depletion and inhibition of PKCɛ result in elevated expression of proapoptotic proteins of the Bcl-2 family, caspase recruitment domain-containing proteins, and tumor necrosis factor ligands/receptor superfamily members [216]. Enhanced PKCɛ expression increases XIAP and Bcl-xL levels and anticancer drug resistance in SCLC cells [217]. These results indicate that PKCɛ is an attractive target for lung cancer therapy.
In addition, there was a positive relationship between PKCι expression and c-Myc/GLUT1 signaling in NSCLC. High co-expression of PKCι and GLUT1 is associated with worse prognosis in patients with NSCLC [218]. Poor prognosis and survival in NSCLC are also positively correlated with PKCη expression [219].
Smoking is the most important risk factor for lung cancer. PKCε is involved in smoke-induced activation of tumor necrosis factor-convertase and hyperproliferation of lung cells [220]. High expression of PKCα, β, and δ showed the strongest positive correlation with RFS, depending on the molecular subtype; smoking; and mutational status of EGFR, KRAS, and TP53 [206]. In an experiment using the carcinogen nitrosamine 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK), which is produced by the nitrosation of nicotine, PKCι activation enhanced the survival and chemoresistance of human lung cancer cells by increasing NNK-induced Bad phosphorylation [221].

2.10. Ovarian Cancer

In ovarian cancer, PKCα upregulation is positively correlated with anticancer drug resistance via activation of the PKCα/ERK1/2 or PKCα/JNK signaling pathways [222,223]. Furthermore, increased expression of Wnt family member 5A (Wnt5a) correlates with enhanced metastasis of ovarian cancer via increased vasculogenic capacity, motility, and invasiveness. Wnt5a enhanced vasculogenic mimicry, EMT, migration, and invasiveness of ovarian cancer cells in a PKCα-dependent manner, and inhibition of PKCα blocked these effects [224]. The PKCα/CARMA3 axis plays an important role in the lysophosphatidic acid-induced invasion of ovarian cancer cells [225]. The expression of PKCα, PKCε, and P-gp is higher in epithelial ovarian cancer tissue than in normal, benign, and borderline epithelial ovarian cancer tissues. They were also more highly expressed in the recurrent carcinoma tissues than in patients with initial treatment and were related to poor survival and prognosis in patients with epithelial ovarian cancer [226].
PKCι activation is positively correlated with histopathological grading, International Federation of Gynecology and Obstetrics (FIGO) stage, and poor survival in patients with ovarian cancer [227]. Similarly, Zhang’s group reported PKCι overexpression in most ovarian carcinomas evaluated and a positive correlation between increased PKCι expression and tumor stage or grade [228]. PKCι protein is markedly increased or mislocalized and associated with decreased progression-free survival in epithelial ovarian cancers. In a Drosophila in vivo epithelial tissue model, increased PKCι levels resulted in defects in apical-basal polarity, cyclin E expression, and proliferation [229]. siRNA-mediated PKCι silencing led to apoptosis in PKCι-amplified ovarian cancer cells, but not in those without PKCι amplification [230]. The PKCι/angiomotin/Yes-associated protein 1 (YAP1) signaling pathway plays a critical role in ovarian cancer prognosis. PKCι inhibition reduces YAP1 nuclear localization and ovarian cancer growth [231,232]. There was also a positive correlation between PKCι and TNF-α expression. Increased levels of TNF-α and YAP1 promote immune suppression by inhibiting the infiltration of cytotoxic T cells [231]. These results suggest that PKCι may be a therapeutic target and prognostic biomarker for ovarian cancer.
High levels of PKCζ are associated with poor prognosis in human ovarian carcinomas [233]. In ovarian cancer, PKCζ has proapoptotic functions and participates in cell invasion and migration. The PKCζ inhibitor ζ-Stat decreased the invasive behavior of ovarian cancer cells by decreasing the activation of cytosolic Ect2 and Rac1 [234].

2.11. Pancreatic, Bile Duct, and Gallbladder Cancer

2.11.1. Pancreatic Cancer

PKCα activation is associated with increased survival, proliferation, migration, and resistance in pancreatic cancer. Hydrophobic motif phosphorylation in PKCα (Ser-657) improves survival in patients with pancreatic adenocarcinoma [235]. Transient receptor potential cation channel subfamily M member 2 (TRPM2) levels were increased in patients with pancreatic ductal adenocarcinoma with increasing tumor stage and showed a negative correlation with overall and progression-free survival time. TRPM2 directly activates PKCα by calcium or indirectly activates PKCε and PKCδ by increasing DAG, leading to activation of the downstream MAPK/MEK pathway [236]. TRAIL-induced apoptosis in pancreatic cancer cells is stimulated by the inhibition of the PKCα/AKT cascade [237]. Chow’s group demonstrated that the TGFβ/PKCα/PTEN pathway is key for the proliferation and metastasis of pancreatic cancer cells [238]. Furthermore, autophagy activation promotes cell survival, proliferation, invasion, and migration in pancreatic cancer [239]. Autophagy activation is dependent on the transcription factor p8, which responds to endoplasmic reticulum stress via the p53/PKCα axis [240]. Several in vivo and in vitro studies suggest that PKCα inhibitors may be of potential therapeutic value against human pancreatic cancers [241,242,243,244]; however, there are no reports of clinical trials using PKCα inhibitors.
High expression of PKCι is associated with poor prognosis in patients with pancreatic cancer [245,246]. High PKCι expression led to increased pancreatic cancer cell growth and migration via the PI3K/AKT and Wnt/β-catenin [246] or Rac1-MEK/ERK1/2 [247] pathways. PKCι is upregulated and activated in pancreatic cancers with mutated KRAS, resulting in increased dephosphorylation and nuclear translocation of YAP1. These changes promote the growth of pancreatic cancer [248]. Inhibition of PKCι alone [249] or in combination with other inhibitors (e.g., specificity protein 1 (Sp1) inhibitor) [250] reduced cell growth and metastasis and induced apoptosis in pancreatic cancer cells. These studies suggest that PKCι can be a promising therapeutic target for pancreatic cancer.
PKCζ activation is positively associated with poor prognosis in patients with pancreatic cancer. It is associated with invasive and metastatic phenotypes of pancreatic adenocarcinoma cells [251]. PKCζ inhibition efficiently reduced pancreatic cancer cell growth and metastasis [249]. PKCζ is a useful immunohistochemical marker for detecting the reverse polarity of invasive micropapillary carcinoma (IMPC) cells. The presence of an IMPC component of <20% was not associated with worse prognosis in patients with pancreatic ductal adenocarcinoma [252].
Enhanced PKCδ expression induces a more malignant phenotype of human ductal pancreatic cancer [253] and is associated with poor survival in patients with pancreatic cancer [254]. Furthermore, PKCδ activation in pancreatic cancer cells increases the expression of MUC1-C oncoprotein, which is associated with the progression of pancreatic cancer [255]. MIST1, a transcription factor, is downregulated in pancreatic cancers [256]. Pancreatic ductal adenocarcinoma showed decreased MIST1 expression, combined with increased nuclear PKCδ accumulation. PKCδ activation increased pancreatic acinar cell dedifferentiation in the absence of MIST1 [257]. Interestingly, following radiotherapy, dying pancreatic cancer cells stimulate the proliferation of living cancer cells via caspase-3/7-dependent PKCδ activation and its downstream Akt/p38 MAPK axis [258]. In addition, PKCδ inhibition may be useful in treating pancreatic cancer with distinct stem-like properties (cancer stem-like cells) [259,260].
PKCθ activation is positively correlated with PKCδ activation and poor survival in patients with pancreatic cancer [254]. In pancreatic cancer cells, MAP4K3 knockdown cells failed to phosphorylate PKCθ, and inhibition of PKCθ activity suppressed insulin-like growth factor-1-mediated cell growth and viability, indicating that the MAP4K3/PKCθ axis may be a therapeutic target for pancreatic cancer [261].

2.11.2. Bile Duct and Gallbladder Cancer

Cholangiocarcinoma (CCA) is a malignant bile duct cancer with a poor prognosis and a low 5-year survival rate (7–20%) [262]. PKCι expression was higher in CCA tissues than in benign bile duct tissues. PKCι expression is positively correlated with cell differentiation and invasion but negatively correlated with E-cadherin expression [263]. PKCι, Snail, and infiltrated immunosuppressive cells are upregulated and associated with poor prognosis in CCA. Although PKCι does not directly interact with Snail, it facilitates EMT and immunosuppression by regulating Snail. PKCι phosphorylates Sp1, and upregulation of phosphorylated Sp1 in CCA tissues is associated with poor prognosis in patients with CCA. Phosphorylated Sp1 regulates Snail expression through the enhanced binding of Sp1 to the Snail promoter [264]. Furthermore, high expression of the adapter protein 14-3-3ζ and PKC-ι was associated with poor prognosis in patients with CCA, and they synergistically induced EMT via the GSK3β/Snail pathway [265]. Therefore, PKCι may be a potential therapeutic target for CCA.
Gallbladder cancer is a rare malignancy with poor prognosis owing to its late diagnosis and rapid progression [266]. PKCι is upregulated and correlates with poor prognosis in patients with gallbladder cancer. PKCι stimulates the aPKCι/Keap1/Nrf2 axis to enhance gallbladder cancer cell growth and drug resistance [267]. Activation of the ASPP2/PKCι/GLI1 cascade promotes cell invasion and metastasis and enhances macrophage recruitment in gallbladder cancer via chemokine ligands (e.g., CCL2 and CCL5) and cytokines (e.g., TNF-α) [268]. Furthermore, PKCϵ was upregulated in peripheral blood samples and stem cells of patients with gallbladder cancer [269]. PKCϵ increased anticancer drug resistance in gallbladder cancer by upregulating MDR1/P-gp [270]. PKCϵ silencing inhibited anticancer drug resistance, proliferation, and colony formation rate and increased apoptosis of gallbladder cancer stem cells [269].

2.12. Prostate Cancer

Higher levels of PKCα, β, ε, and η have been detected in malignant prostate tissues than in benign tissues [271]. Moreover, increased PKCα and ζ; decreased PKCβ; and absence of PKCγ, δ, and θ expression were observed in early prostate cancer specimens [272]. However, some studies have reported enhanced PKCδ expression in both low- and high-grade prostate cancer [273,274].
Enhanced PKCα and β activation promotes prostate cancer cell proliferation and growth [275,276], and inhibition of PKCα and β induces apoptosis [276,277]. PKCα activation also increases anticancer drug resistance in prostate cancer cells by increasing Ser70-phosphorylated Bcl-2 and total Bcl-2 protein [278]. In contrast, PKCα activation reduced ATM and increased radiation-mediated apoptosis of androgen-sensitive human prostate cancer cells by stimulating ceramide synthase [279].
PKCδ mediates anticancer drug-induced apoptosis in prostate cancer. For example, apoptosis of prostate cancer cells induced by cystine dimethyl ester [280], PMA [281,282], moracin D [283], and paclitaxel [284] depends on PKCδ activity. PKCδ inhibition represents a potential strategy for treating prostate CSC [116].
PKCε overexpression is positively correlated with prostate cancer development [285,286]. PKCε-mediated signal transducer and activator of transcription-3 (Stat3) Ser727 phosphorylation through integration with the MAPK cascade (RAF-1, MEK1/2, and ERK1/2) is essential for prostate cancer cell invasion [287]. Moreover, PKCε activation has been linked to PTEN loss in prostate tumorigenesis via the CXCL13-CXCR5 pathway [286]. PKCε inhibition also led to significant downregulation of proliferative and metastatic genes, such as C/EBPβ (CCAAT/enhancer binding protein β), CRP (C-reactive protein), CMK, EGFR, CD64, Jun B, and gp130 [288].
aPKCs (PKCζ and λ/ι) are involved in cell growth, invasion, migration, and apoptosis in prostate cancer. PKCζ expression is positively correlated with poor overall survival in prostate cancer [289]. Inhibitors of PKCζ and λ/ι are therapeutic molecules for prostate cancer [290,291,292,293]. For example, treatment with the aPKC inhibitors 2-acetyl-1,3-cyclopentanedione (ACPD) and ICA-1 significantly decreased malignant cell proliferation and induced apoptosis [290,291]. Furthermore, inhibition of aPKCs attenuates prostate cancer cell metastasis by downregulating vimentin expression [293]. PKCζ inhibition also prevents CXCL12-driven cell migration [292]. Treatment-emergent neuroendocrine prostate cancer (NEPC) is a lethal form of castration-resistant prostate cancer [294]. Interestingly, in NEPC, PKCλ/ι downregulation stimulates serine biosynthesis through the mTORC1/ATF4/PHGDH axis and DNA methylation, resulting in enhanced NEPC differentiation and growth. However, inhibition of DNA methyltransferase activity blocks NEPC differentiation and growth induced by PKCλ/ι downregulation [295]. In addition, two PKCι single nucleotide polymorphisms, rs546950 and rs4955720, are associated with prostate cancer risk in Iranian [296] and Eastern Chinese populations [297]. These results suggest that aPKCs may be potential targets for the prevention and/or treatment of prostate cancer.

2.13. Renal Cell Carcinoma (RCC)

RCCs can be classified into four types: clear cells (70–80%), papillary (10–20%), chromophobe (5%), and collecting duct (1%) [298,299]. Expression of PKCα, βI, βII, δ, ε, η, ζ, and ι, but not PKCγ and θ, was observed in patients with clear cell RCC (ccRCC) [300]. Another study reported a relationship between PKCζ, RCC grade, and poor patient survival [301]. Increased PKCη (3 times) and PKCζ (20%) levels were observed in grade 3 and 4 versus grade 1 and 2 ccRCCs [300]. However, PKCα level was decreased in ccRCC versus normal tissue [300,302]. In another study, PKCβI, βII, δ, and ε were expressed in ccRCCs, whereas PKCα, βI, βII, η, and ι were expressed in oncocytoma, a benign kidney tumor [302].
PKCδ activation induces migration of ccRCC cells by stimulating CDCP1 [303] or β1 integrin and FAK [304]. High CDCP1 activation is associated with poor prognosis in patients [303].
PKCε also induces RCC proliferation by regulating β1 integrin [305]. PKCε expression positively correlated with Fuhrman grade and T stage in ccRCC. Inhibition of PKCε activation in the ccRCC cell line 769P inhibited cell growth, migration, and invasion, and it sensitized cells to anticancer drugs by increasing caspase-3 activity [306]. PKCε depletion suppressed the sorting and cancer stem-like phenotype of 769P side population cells by decreasing the ABCB1 transporter and the PI3K/Akt, Stat3, and MAPK/ERK pathways [307]. Moreover, PKCε-mediated claudin-4 phosphorylation induces the EMT phenotype and invasive and metastatic abilities in RCC cells [308]. Another study showed that PKCα and PKCε activation increases the invasive potential of RCC [309]. These results indicate that PKCε may be a potential therapeutic target for RCC.

2.14. Skin Cancer

Skin cancers are classified as melanoma and non-melanoma skin cancer (NMSC). The main types of NMSC are basal cell carcinoma (BCC) and SCC [310]. PKCα, δ, ε, ζ, and λ/ι are expressed in melanoma cells [311,312]. PKCβI and βII are expressed exclusively in normal melanocytes or epidermal melanocytes but are downregulated in melanoma cells and benign and malignant melanocytic lesions [311,313,314]. The loss of PKCβ is important for melanoma cell growth [315].

2.14.1. Melanoma

PKCα is overexpressed in melanoma tumor samples and is associated with poor overall survival [316]. PKCα is regarded as a potential therapeutic target for melanoma because it increases melanoma cell invasion by activating the AKT/ERK1/2 axis [317] or, in an αvβ3-dependent manner [318], increases cell proliferation by enhancing the G1 to S transition [319], and it increases melanoma vascularization in a vascular endothelial growth factor receptor-1 (VEGFR1)-independent manner [320].
In melanoma, PKCδ is associated with proapoptotic responses through JNK activation [321] or by inhibition of PKCα/PLD1/AKT signaling [319]. However, another study demonstrated that PKCδ inhibition reduced uveal melanoma cell growth through p53 reactivation [322]. In a recent phase I study, treatment with the PKC inhibitor AEB071 (also known as sotrastaurin) was well tolerated and showed modest clinical activity in patients with metastatic uveal melanoma [323].
PKCζ and ι are also regarded as therapeutic targets for melanoma. In melanoma cells, the aPKC/AKT/NF-κB and PKCι/Par6/RhoA pathways are involved in cell proliferation and increased EMT, respectively. Inhibition of both PKCζ and PKCι reduces EMT and induces apoptosis in melanoma cells [324]. However, PKCι is more involved in melanoma malignancy than PKCζ. Treatment with ICA-1 (PKCι-specific inhibitor) and ζ-Stat (PKCζ-specific inhibitor) reduced melanoma cell proliferation and induced apoptosis, whereas ICA-1 also reduced cell migration and invasion [325].
PKCε-mediated activation of activating transcription factor-2 (ATF2) regulates the migration and invasion of melanoma cells via cellular protein fucosylation. Activated PKCε and ATF2 were observed in advanced-stage melanomas and correlated with decreased cellular protein fucosylation, attenuated cell adhesion, and increased cell motility [326,327]. Furthermore, PKCε is involved in metabotropic glutamate receptor-1-mediated ERK1/2 phosphorylation, resulting in enhanced melanomagenesis and metastasis [328,329].

2.14.2. NMSC

PKCδ plays a protective role in SCC by downregulating p63 and suppressing cell proliferation [330] or by inducing apoptosis in SCC cells [331]. PKCε is involved in ultraviolet radiation (UVR)-induced SCC development. Following UVR treatment, the clonogenicity of isolated keratinocytes increased in PKCε-overexpressing transgenic mice [332]. The PKCε–Stat3 and PKCε–ERK1/2 interactions were also increased in SCC elicited following repeated UVR exposure. PKCε-mediated activation of Stat3 and ERK1/2 increased SCC development [333,334]. In addition, Hedgehog-dependent BCC growth is stimulated by activation of the mTOR/aPKC [335] or aPKC/histone deacetylase axes [336].

2.15. Thyroid Carcinoma

The expression of phosphorylated PKCδ along with that of cytokeratin 18, Stat1, HMG-1, p-p70 S6 kinase, Raf-B, glutamine synthetase, and HDAC1 was upregulated in papillary thyroid carcinoma [337]. PKCε expression is reduced in papillary thyroid carcinomas [338]. In anaplastic and follicular thyroid cancer cell lines, PMA treatment stimulates the translocation of PKCα, βI, and δ. PKCδ deletion reduces the PMA-induced antiproliferative effect by inducing cell cycle arrest in the G1/S phase [339]. The expression and localization of PKCβII and PKCδ were observed in medullary thyroid carcinomas. PKCβII inhibition by enzastaurin reduced cell proliferation and survival by inducing caspase-mediated apoptosis and blocking the stimulatory effect of IGF-I on calcitonin secretion [340]. Furthermore, mutated PKCα has been found in pituitary and thyroid tumors [341] and follicular thyroid carcinoma [342,343]. D294G, but not A294G, is a loss-of-function mutation [341,343].

3. PKC Isozymes as Diagnostic Biomarkers for Cancer

3.1. PKC Isozymes as Diagnostic Immunohistochemical Biomarkers

Compared to normal tissues, overexpression of PKC isozymes in cancer tissues can be used as a diagnostic immunohistochemical biomarker for specific cancer types. For example, higher PKCζ expression was found in invasive ductal carcinoma than in healthy breast tissue [121]. Furthermore, PKCι was significantly upregulated in ovarian cancer compared to normal ovarian tissue. There was a positive correlation between PKCι expression and tumor stage or grade [228]. DOG1 and PKCθ are overexpressed in KIT-negative gastrointestinal stromal tumors, indicating that DOG1 and/or PKCθ may be used in the diagnosis of KIT-negative GISTs [344,345,346]. As mentioned in OTSCC, PKCα was significantly overexpressed in young patients (≤45 years) compared to older patients (>45 years). PKCα overexpression was associated with poor overall and disease-free survival as well as with no alcohol and tobacco consumption. These results indicate that PKCα overexpression may be a novel diagnostic molecular marker for early-onset alcohol- and tobacco-negative high-risk OTSCC [174].

3.2. PKC Isozymes as Diagnostic Biomarkers in Body Fluids

Diagnostic cancer biomarkers in body fluids (e.g., blood, urine, feces, or saliva) offer several advantages, such as simple and non-invasive sample collection methods that are less painful in patients, when compared to diagnostic immunohistochemical biomarkers using tissue samples. PKC isozymes are detectable in body fluids as they are secreted by cancer cells [347,348,349].
High levels of activated PKCα have been observed in blood samples collected from cancer-bearing mice [347,348] and patients with lung cancer [350]. However, very low levels of activated PKCα were found in blood samples obtained from healthy mice [347,348] and humans [269]. Furthermore, despite the lack of identification of PKC isozyme, higher serum levels of PKC as well as FAK, MR-1, and Src were identified in patients with AML than in controls [351]. Expression of PKCε was significantly reduced in the blood of patients with cervical cancer compared to that in healthy controls [352].
PKCα expression negatively correlates with urinary microRNA (miR)-15a in patients with ccRCC. Increased miR-15a levels were determined in the urine of patients with RCC but were nearly undetectable in oncocytoma, other tumors, and urinary tract inflammation [302]. PKCε downregulation was closely related to miR-31 upregulation [353]. Urinary levels of miR-31 are higher in oncocytomas than in ccRCCs [354]. Recently, our group reported that high levels of activated PKCα were observed in urine samples collected from orthotopic xenograft mice bearing human bladder cancer cells compared with urine samples from normal mice [355]. In urine samples from patients with ccRCC, PKCα levels increased with increasing regression rate. However, PKCι levels were increased in urine samples from patients with oncocytoma but reduced in samples from patients with ccRCC [356].
In addition, increased fecal PKCβII mRNA levels and decreased fecal ζ mRNA levels were found in samples collected from colon cancer-bearing rats compared with those from normal rats [357].

4. Summary and Overall Conclusions

PKC isozymes represent potential therapeutic targets in cancer (Table 1). Several natural and synthetic PKC inhibitors have been developed and used in clinical trials. However, most clinical trials using PKC inhibitors with or without other anticancer agents have failed to show significant clinical benefits [7]. Despite these unfavorable results, the fact remains that PKC isozymes constitute attractive therapeutic targets for cancer, and satisfactory clinical results with PKC inhibitors may be obtained when combined with other inhibitors of cancer-related signaling pathways (e.g., TKIs) [7].
Many studies have shown positive relationships between PKC isozymes and poor disease-free survival and survival rates, poor survival following anticancer treatment, and enhanced recurrence (Table 1). Furthermore, several groups have reported differential expression of PKC isozymes by cancer type, for example, PKCθ overexpression in KIT-negative GISTs [344,345,346] or PKCα overexpression in OTSCC [170]. Therefore, PKC isozymes hold great potential as prognostic and diagnostic biomarkers.
PKC-based cancer diagnosis has been performed mainly using tissue samples collected from patients with cancer. Inactivated PKC isozymes are present in the cytosol; however, following activation, PKC isozymes translocate from the cytosol to the inner cell membrane. Several studies have suggested that activated PKC isozymes present in the extracellular space are released into the bloodstream and urine [347,348,349,355]. These studies indicate that PKC isozymes in body fluids (e.g., blood, urine, feces, or saliva) may be potential diagnostic biomarkers for cancer. However, there are very few reports based on PKC isozymes in body fluids. Furthermore, the mechanism by which PKC isozymes are released into bodily fluids remains unclear.

Author Contributions

Conceptualization, T.K. and J.-H.K.; writing-original draft preparation, T.K. and J.-H.K.; writing-review and editing, J.I., M.E. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the Japan Agency for Medical Research and Development (AMED; grant number: JP22he0122023) and Japan Society for the Promotion of Science (JSPS) KAKENHI (grant number: 22H03976).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Steinberg, S.F. Structural basis of protein kinase C isoform function. Physiol. Rev. 2008, 88, 1341–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Newton, A.C. Protein kinase C: Perfectly balanced. Crit. Rev. Biochem. Mol. Biol. 2018, 5, 208–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kang, J.H.; Toita, R.; Kim, C.W.; Katayama, Y. Protein kinase C (PKC) isozyme-specific substrates and their design. Biotechnol. Adv. 2012, 30, 1662–1672. [Google Scholar] [CrossRef] [PubMed]
  4. Jaken, S. Protein kinase C isozymes and substrates. Curr. Opin. Cell Biol. 1996, 8, 168–173. [Google Scholar] [CrossRef]
  5. Hofmann, J. The potential for isoenzyme-selective modulation of protein kinase C. FASEB J. 1997, 11, 649–669. [Google Scholar]
  6. Gonelli, A.; Mischiati, C.; Guerrini, R.; Voltan, R.; Salvadori, S.; Zauli, G. Perspectives of protein kinase C (PKC) inhibitors as anti-cancer agents. Mini Rev. Med. Chem. 2009, 9, 498–509. [Google Scholar] [CrossRef]
  7. Kawano, T.; Inokuchi, J.; Eto, M.; Murata, M.; Kang, J.H. Activators and inhibitors of protein kinase C (PKC): Their applications in clinical trials. Pharmaceutics 2021, 13, 1748. [Google Scholar]
  8. Langzam, L.; Koren, R.; Gal, R.; Kugel, V.; Paz, A.; Farkas, A.; Sampson, S.R. Patterns of protein kinase C isoenzyme expression in transitional cell carcinoma of bladder. Relation to degree of malignancy. Am. J. Clin. Pathol. 2001, 116, 377–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Varga, A.; Czifra, G.; Tállai, B.; Németh, T.; Kovács, I.; Kovács, L.; Bíró, T. Tumor grade-dependent alterations in the protein kinase C isoform pattern in urinary bladder carcinomas. Eur. Urol. 2004, 46, 462–465. [Google Scholar] [CrossRef]
  10. Kang, J.H.; Inokuchi, J.; Kawano, T.; Murata, M. Protein kinase Cα as a therapeutic target in cancer. In Protein Kinase C: Emerging Roles and Therapeutic Potential; Pierce, D.N., Ed.; Nova Science Publishers, Inc.: New York, NY, USA, 2018; pp. 25–47. [Google Scholar]
  11. Xu, W.; Anwaier, A.; Ma, C.; Liu, W.; Tian, X.; Palihati, M.; Hu, X.; Qu, Y.; Zhang, H.; Ye, D. Multi-omics reveals novel prognostic implication of SRC protein expression in bladder cancer and its correlation with immunotherapy response. Ann. Med. 2021, 53, 596–610. [Google Scholar] [CrossRef]
  12. Du, H.F.; Ou, L.P.; Yang, X.; Song, X.D.; Fan, Y.R.; Tan, B.; Luo, C.L.; Wu, X.H. A new PKCα/β/TBX3/E-cadherin pathway is involved in PLCε-regulated invasion and migration in human bladder cancer cells. Cell Signal. 2014, 26, 580–593. [Google Scholar] [CrossRef] [PubMed]
  13. Ling, Y.; Chunli, L.; Xiaohou, W.; Qiaoling, Z. Involvement of the PLCε/PKCα pathway in human BIU-87 bladder cancer cell proliferation. Cell Biol. Int. 2011, 35, 1031–1036. [Google Scholar] [CrossRef] [PubMed]
  14. Zheng, J.; Kong, C.; Yang, X.; Cui, X.; Lin, X.; Zhang, Z. Protein kinase C-α (PKCα) modulates cell apoptosis by stimulating nuclear translocation of NF-kappa-B p65 in urothelial cell carcinoma of the bladder. BMC Cancer 2017, 17, 432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Zhang, X.; Zhang, J.; Zhang, H.; Liu, Y.; Yin, L.; Liu, X.; Li, X.; Yu, X.; Yao, J.; Zhang, Z.; et al. Exploring the five different genes associated with PKCα in bladder cancer based on gene expression microarray. J. Cell Mol. Med. 2021, 25, 1759–1770. [Google Scholar] [PubMed]
  16. Liu, J.; Li, J. PKCα and Netrin-1/UNC5B positive feedback control in relation with chemical therapy in bladder cancer. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 1712–1717. [Google Scholar] [PubMed]
  17. Tan, S.T.; Liu, S.Y.; Wu, B. TRIM29 overexpression promotes proliferation and survival of bladder cancer cells through NF-κB signaling. Cancer Res. Treat. 2016, 48, 1302–1312. [Google Scholar] [CrossRef]
  18. Kong, C.; Zhu, Y.; Liu, D.; Yu, M.; Li, S.; Li, Z.; Sun, Z.; Liu, G. Role of protein kinase C-α in superficial bladder carcinoma recurrence. Urology 2005, 65, 1228–1232. [Google Scholar] [CrossRef]
  19. Jiang, Z.; Kong, C.; Zhang, Z.; Zhu, Y.; Zhang, Y.; Chen, X. Reduction of protein kinase C α (PKC-α) promote apoptosis via down-regulation of Dicer in bladder cancer. J. Cell. Mol. Med. 2015, 19, 1085–1093. [Google Scholar]
  20. Namdarian, B.; Wong, E.; Galea, R.; Pedersen, J.; Chin, X.; Speirs, R.; Humbert, P.O.; Costello, A.J.; Corcoran, N.M.; Hovens, C.M. Loss of APKC expression independently predicts tumor recurrence in superficial bladder cancers. Urol. Oncol. 2013, 31, 649–655. [Google Scholar]
  21. Patel, R.; Islam, S.A.; Bommareddy, R.R.; Smalley, T.; Acevedo-Duncan, M. Simultaneous inhibition of atypical protein kinase-C and mTOR impedes bladder cancer cell progression. Int. J. Oncol. 2020, 56, 1373–1386. [Google Scholar] [CrossRef] [Green Version]
  22. Neri, A.; Marmiroli, S.; Tassone, P.; Lombardi, L.; Nobili, L.; Verdelli, D.; Civallero, M.; Cosenza, M.; Bertacchini, J.; Federico, M.; et al. The oral protein-kinase Cβ inhibitor enzastaurin (LY317615) suppresses signalling through the AKT pathway, inhibits proliferation and induces apoptosis in multiple myeloma cell lines. Leuk. Lymphoma 2008, 49, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  23. Podar, K.; Raab, M.S.; Zhang, J.; McMillin, D.; Breitkreutz, I.; Tai, Y.T.; Lin, B.K.; Munshi, N.; Hideshima, T.; Chauhan, D.; et al. Targeting PKC in multiple myeloma: In vitro and in vivo effects of the novel, orally available small-molecule inhibitor enzastaurin (LY317615.HCl). Blood 2007, 109, 1669–1677. [Google Scholar] [CrossRef] [PubMed]
  24. Jourdan, E.; Leblond, V.; Maisonneuve, H.; Benhadji, K.A.; Hossain, A.M.; Nguyen, T.S.; Wooldridge, J.E.; Moreau, P. A multicenter phase II study of single-agent enzastaurin in previously treated multiple myeloma. Leuk. Lymphoma. 2014, 55, 2013–2017. [Google Scholar] [CrossRef]
  25. Alkan, S.; Huang, Q.; Ergin, M.; Denning, M.F.; Nand, S.; Maududi, T.; Paner, G.P.; Ozpuyan, F.; Izban, K.F. Survival role of protein kinase C (PKC) in chronic lymphocytic leukemia and determination of isoform expression pattern and genes altered by PKC inhibition. Am. J. Hematol. 2005, 79, 97–106. [Google Scholar] [CrossRef] [PubMed]
  26. Frezzato, F.; Accordi, B.; Trimarco, V.; Gattazzo, C.; Martini, V.; Milani, G.; Bresolin, S.; Severin, F.; Visentin, A.; Basso, G.; et al. Profiling B cell chronic lymphocytic leukemia by reverse phase protein array: Focus on apoptotic proteins. J. Leukoc. Biol. 2016, 100, 1061–1070. [Google Scholar] [CrossRef] [Green Version]
  27. Lei, J.; Li, Q.; Gao, Y.; Zhao, L.; Liu, Y. Increased PKCα activity by Rack1 overexpression is responsible for chemotherapy resistance in T-cell acute lymphoblastic leukemia-derived cell line. Sci. Rep. 2016, 6, 33717. [Google Scholar] [CrossRef]
  28. Jiffar, T.; Kurinna, S.; Suck, G.; Carlson-Bremer, D.; Ricciardi, M.R.; Konopleva, M.; Andreeff, M.; Ruvolo, P.P. PKCα mediates chemoresistance in acute lymphoblastic leukemia through effects on Bcl2 phosphorylation. Leukemia 2004, 18, 505–512. [Google Scholar] [CrossRef] [Green Version]
  29. Lutzny, G.; Kocher, T.; Schmidt-Supprian, M.; Rudelius, M.; Klein-Hitpass, L.; Finch, A.J.; Dürig, J.; Wagner, M.; Haferlach, C.; Kohlmann, A.; et al. Protein kinase C-β-dependent activation of NF-κB in stromal cells is indispensable for the survival of chronic lymphocytic leukemia B cells in vivo. Cancer Cell 2013, 23, 77–92. [Google Scholar] [CrossRef] [Green Version]
  30. El-Gamal, D.; Williams, K.; LaFollette, T.D.; Cannon, M.; Blachly, J.S.; Zhong, Y.; Woyach, J.A.; Williams, E.; Awan, F.T.; Jones, J.; et al. PKC-β as a therapeutic target in CLL: PKC inhibitor AEB071 demonstrates preclinical activity in CLL. Blood 2014, 124, 1481–1491. [Google Scholar] [CrossRef] [Green Version]
  31. Handl, S.; von Heydebrand, F.; Voelkl, S.; Oostendorp, R.A.J.; Wilke, J.; Kremer, A.N.; Mackensen, A.; Lutzny-Geier, G. Immune modulatory effects of Idelalisib in stromal cells of chronic lymphocytic leukemia. Leuk. Lymphoma 2021, 62, 2679–2689. [Google Scholar] [CrossRef]
  32. Zum Büschenfelde, C.M.; Wagner, M.; Lutzny, G.; Oelsner, M.; Feuerstacke, Y.; Decker, T.; Bogner, C.; Peschel, C.; Ringshausen, I. Recruitment of PKC-βII to lipid rafts mediates apoptosis-resistance in chronic lymphocytic leukemia expressing ZAP-70. Leukemia 2010, 24, 141–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Amigo-Jiménez, I.; Bailón, E.; Aguilera-Montilla, N.; Terol, M.J.; García-Marco, J.A.; García-Pardo, A. Bone marrow stroma-induced resistance of chronic lymphocytic leukemia cells to arsenic trioxide involves Mcl-1 upregulation and is overcome by inhibiting the PI3Kδ or PKCβ signaling pathways. Oncotarget 2015, 6, 44832–44848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Loi, T.H.; Dai, P.; Carlin, S.; Melo, J.V.; Ma, D.D.F. Pro-survival role of protein kinase Cε in Philadelphia chromosome positive acute leukemia. Leuk. Lymphoma 2016, 57, 411–418. [Google Scholar] [CrossRef] [PubMed]
  35. Ringshausen, I.; Schneller, F.; Bogner, C.; Hipp, S.; Duyster, J.; Peschel, C.; Decker, T. Constitutively activated phosphatidylinositol-3 kinase (PI-3K) is involved in the defect of apoptosis in B-CLL: Association with protein kinase Cδ. Blood 2002, 100, 3741–3748. [Google Scholar] [CrossRef] [PubMed]
  36. Hartsink-Segers, S.A.; Beaudoin, J.J.; Luijendijk, M.W.; Exalto, C.; Pieters, R.; Den Boer, M.L. PKCζ and PKMζ are overexpressed in TCF3-rearranged paediatric acute lymphoblastic leukaemia and are associated with increased thiopurine sensitivity. Leukemia 2015, 29, 304–311. [Google Scholar] [CrossRef] [Green Version]
  37. Hubmann, R.; Düchler, M.; Schnabl, S.; Hilgarth, M.; Demirtas, D.; Mitteregger, D.; Hölbl, A.; Vanura, K.; Le, T.; Look, T.; et al. NOTCH2 links protein kinase Cδ to the expression of CD23 in chronic lymphocytic leukaemia (CLL) cells. Br. J. Haematol. 2010, 148, 868–878. [Google Scholar] [CrossRef]
  38. Giambra, V.; Jenkins, C.R.; Wang, H.; Lam, S.H.; Shevchuk, O.O.; Nemirovsky, O.; Wai, C.; Gusscott, S.; Chiang, M.Y.; Aster, J.C.; et al. NOTCH1 promotes T cell leukemia-initiating activity by RUNX-mediated regulation of PKC-θ and reactive oxygen species. Nat. Med. 2012, 18, 1693–1698. [Google Scholar] [CrossRef]
  39. Nayak, R.C.; Hegde, S.; Althoff, M.J.; Wellendorf, A.M.; Mohmoud, F.; Perentesis, J.; Reina-Campos, M.; Reynaud, D.; Zheng, Y.; Diaz-Meco, M.T.; et al. The signaling axis atypical protein kinase C λ/ι-Satb2 mediates leukemic transformation of B-cell progenitors. Nat. Commun. 2019, 10, 46. [Google Scholar] [CrossRef]
  40. Kurinna, S.; Konopleva, M.; Palla, S.L.; Chen, W.; Kornblau, S.; Contractor, R.; Deng, X.; May, W.S.; Andreeff, M.; Ruvolo, P.P. Bcl2 phosphorylation and active PKCα are associated with poor survival in AML. Leukemia 2006, 20, 1316–1319. [Google Scholar] [CrossRef] [Green Version]
  41. Ruvolo, P.P.; Deng, X.; Carr, B.K.; May, W.S. A functional role for mitochondrial protein kinase Cα in Bcl2 phosphorylation and suppression of apoptosis. J. Biol. Chem. 1998, 273, 25436–25442. [Google Scholar] [CrossRef] [Green Version]
  42. Li, Z.S.; Shi, K.J.; Guan, L.Y.; Jiang, Q.; Yang, Y.; Xu, C.M. Downregulation of protein kinase Cα was involved in selenite-induced apoptosis of NB4 cells. Oncol. Res. 2010, 19, 77–83. [Google Scholar] [CrossRef] [PubMed]
  43. Nicholson, R.; Menezes, A.C.; Azevedo, A.; Leckenby, A.; Davies, S.; Seedhouse, C.; Gilkes, A.; Knapper, S.; Tonks, A.; Darley, R.L. Protein kinase Cε overexpression is associated with poor patient outcomes in AML and promotes daunorubicin resistance through p-glycoprotein-mediated drug efflux. Front. Oncol. 2022, 12, 840046. [Google Scholar] [CrossRef] [PubMed]
  44. Di Marcantonio, D.; Martinez, E.; Sidoli, S.; Vadaketh, J.; Nieborowska-Skorska, M.; Gupta, A.; Meadows, J.M.; Ferraro, F.; Masselli, E.; Challen, G.A.; et al. Protein kinase Cε is a key regulator of mitochondrial redox homeostasis in acute myeloid leukemia. Clin. Cancer Res. 2018, 24, 608–618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Ozpolat, B.; Akar, U.; Tekedereli, I.; Alpay, S.N.; Barria, M.; Gezgen, B.; Zhang, N.; Coombes, K.; Kornblau, S.; Lopez-Berestein, G. PKCδ regulates translation initiation through PKR and eIF2α in response to retinoic acid in acute myeloid leukemia cells. Leuk. Res. Treat. 2012, 2012, 482905. [Google Scholar] [CrossRef] [PubMed]
  46. Song, M.G.; Gao, S.M.; Du, K.M.; Xu, M.; Yu, Y.; Zhou, Y.H.; Wang, Q.; Chen, Z.; Zhu, Y.S.; Chen, G.Q. Nanomolar concentration of NSC606985, a camptothecin analog, induces leukemic-cell apoptosis through protein kinase Cδ-dependent mechanisms. Blood 2005, 105, 3714–3721. [Google Scholar] [CrossRef] [Green Version]
  47. Yan, H.; Wang, Y.C.; Li, D.; Wang, Y.; Liu, W.; Wu, Y.L.; Chen, G.Q. Arsenic trioxide and proteasome inhibitor bortezomib synergistically induce apoptosis in leukemic cells: The role of protein kinase Cδ. Leukemia 2007, 21, 1488–1495. [Google Scholar] [CrossRef]
  48. Gao, F.H.; Wu, Y.L.; Zhao, M.; Liu, C.X.; Wang, L.S.; Chen, G.Q. Protein kinase C-δ mediates down-regulation of heterogeneous nuclear ribonucleoprotein K protein: Involvement in apoptosis induction. Exp. Cell Res. 2009, 315, 3250–3258. [Google Scholar] [CrossRef]
  49. Stone, R.M.; Mandrekar, S.J.; Sanford, B.L.; Laumann, K.; Geyer, S.; Bloomfield, C.D.; Thiede, C.; Prior, T.W.; Döhner, K.; Marcucci, G.; et al. Midostaurin plus chemotherapy for acute myeloid leukemia with a FLT3 mutation. N. Engl. J. Med. 2017, 377, 454–464. [Google Scholar] [CrossRef]
  50. Weisberg, E.; Boulton, C.; Kelly, L.M.; Manley, P.; Fabbro, D.; Meyer, T.; Gilliland, D.G.; Griffin, J.D. Inhibition of mutant FLT3 receptors in leukemia cells by the small molecule tyrosine kinase inhibitor PKC412. Cancer Cell 2002, 1, 433–443. [Google Scholar] [CrossRef] [Green Version]
  51. Ma, D.; Wang, P.; Fang, Q.; Yu, Z.; Zhou, Z.; He, Z.; Wei, D.; Yu, K.; Lu, T.; Zhang, Y.; et al. Low-dose staurosporine selectively reverses BCR-ABL-independent IM resistance through PKC-α-mediated G2/M phase arrest in chronic myeloid leukaemia. Artif. Cells Nanomed. Biotechnol. 2018, 46 (Suppl. 3), S208–S216. [Google Scholar] [CrossRef] [Green Version]
  52. Ma, D.; Liu, P.; Wang, P.; Zhou, Z.; Fang, Q.; Wang, J. PKC-β/Alox5 axis activation promotes Bcr-Abl-independent TKI-resistance in chronic myeloid leukemia. J. Cell. Physiol. 2021, 236, 6312–6327. [Google Scholar] [CrossRef] [PubMed]
  53. Ma, L.; Shan, Y.; Bai, R.; Xue, L.; Eide, C.A.; Ou, J.; Zhu, L.J.; Hutchinson, L.; Cerny, J.; Khoury, H.J.; et al. A therapeutically targetable mechanism of BCR-ABL-independent imatinib resistance in chronic myeloid leukemia. Sci. Transl. Med. 2014, 6, 252ra121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Gangat, N.; Patnaik, M.M.; Tefferi, A. Myelodysplastic syndromes: Contemporary review and how we treat. Am. J. Hematol. 2016, 91, 76–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Cazzola, M. Myelodysplastic syndromes. N. Engl. J. Med. 2020, 383, 1358–1374. [Google Scholar] [CrossRef]
  56. Poli, A.; Ratti, S.; Finelli, C.; Mongiorgi, S.; Clissa, C.; Lonetti, A.; Cappellini, A.; Catozzi, A.; Barraco, M.; Suh, P.G.; et al. Nuclear translocation of PKC-α is associated with cell cycle arrest and erythroid differentiation in myelodysplastic syndromes (MDSs). FASEB J. 2018, 32, 681–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Cappellini, A.; Mongiorgi, S.; Finelli, C.; Fazio, A.; Ratti, S.; Marvi, M.V.; Curti, A.; Salvestrini, V.; Pellagatti, A.; Billi, A.M.; et al. Phospholipase C beta1 (PI-PLCbeta1)/Cyclin D3/protein kinase C (PKC) alpha signaling modulation during iron-induced oxidative stress in myelodysplastic syndromes (MDS). FASEB J. 2020, 34, 15400–15416. [Google Scholar] [CrossRef]
  58. Kamimura, K.; Hojo, H.; Abe, M. Characterization of expression of protein kinase C isozymes in human B-cell lymphoma: Relationship between its expression and prognosis. Pathol. Int. 2004, 54, 224–230. [Google Scholar] [CrossRef]
  59. Bakalova, R.; Ohba, H.; Zhelev, Z.; Kubo, T.; Fujii, M.; Ishikawa, M.; Shinohara, Y.; Baba, Y. Atypical protein-kinase Cζ, but neither conventional Ca2+-dependent protein-kinase C isoenzymes nor Ca2+-calmodulin, participates in regulation of telomerase activity in Burkitt’s lymphoma cells. Cancer Chemother. Pharmacol. 2004, 54, 161–172. [Google Scholar] [CrossRef]
  60. Decouvelaere, A.V.; Morschhauser, F.; Buob, D.; Copin, M.C.; Dumontet, C. Heterogeneity of protein kinase C β2 expression in lymphoid malignancies. Histopathology 2007, 50, 561–566. [Google Scholar] [CrossRef]
  61. Chao, C.; Silverberg, M.J.; Xu, L.; Chen, L.H.; Castor, B.; Martínez-Maza, O.; Abrams, D.I.; Zha, H.D.; Haque, R.; Said, J. A comparative study of molecular characteristics of diffuse large B-cell lymphoma from patients with and without human immunodeficiency virus infection. Clin. Cancer Res. 2015, 21, 1429–1437. [Google Scholar] [CrossRef] [Green Version]
  62. Li, S.; Phong, M.; Lahn, M.; Brail, L.; Sutton, S.; Lin, B.K.; Thornton, D.; Liao, B. Retrospective analysis of protein kinase C-β (PKC-β) expression in lymphoid malignancies and its association with survival in diffuse large B-cell lymphomas. Biol. Direct. 2007, 2, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chaiwatanatorn, K.; Stamaratis, G.; Opeskin, K.; Firkin, F.; Nandurkar, H. Protein kinase C-βII expression in diffuse large B-cell lymphoma predicts for inferior outcome of anthracycline-based chemotherapy with and without rituximab. Leuk. Lymphoma 2009, 50, 1666–1675. [Google Scholar] [CrossRef] [PubMed]
  64. Espinosa, I.; Briones, J.; Bordes, R.; Brunet, S.; Martino, R.; Sureda, A.; Prat, J.; Sierra, J. Membrane PKC-β2 protein expression predicts for poor response to chemotherapy and survival in patients with diffuse large B-cell lymphoma. Ann. Hematol. 2006, 85, 597–603. [Google Scholar] [CrossRef] [PubMed]
  65. Hans, C.P.; Weisenburger, D.D.; Greiner, T.C.; Chan, W.C.; Aoun, P.; Cochran, G.T.; Pan, Z.; Smith, L.M.; Lynch, J.C.; Bociek, R.G.; et al. Expression of PKC-β or cyclin D2 predicts for inferior survival in diffuse large B-cell lymphoma. Mod. Pathol. 2005, 18, 1377–1384. [Google Scholar] [CrossRef] [PubMed]
  66. Schaffel, R.; Morais, J.C.; Biasoli, I.; Lima, J.; Scheliga, A.; Romano, S.; Milito, C.; Spector, N. PKC-βII expression has prognostic impact in nodal diffuse large B-cell lymphoma. Mod. Pathol. 2007, 20, 326–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Mishra, S.; Vinayak, M. Role of ellagic acid in regulation of apoptosis by modulating novel and atypical PKC in lymphoma bearing mice. BMC Complement Altern. Med. 2015, 15, 281. [Google Scholar] [CrossRef] [Green Version]
  68. Sumarni, U.; Reidel, U.; Eberle, J. Targeting cutaneous T-cell lymphoma cells by ingenol mebutate (PEP005) correlates with PKCδ activation, ROS induction as well as downregulation of XIAP and c-FLIP. Cells 2021, 10, 987. [Google Scholar] [CrossRef]
  69. Yanase, N.; Hayashida, M.; Kanetaka-Naka, Y.; Hoshika, A.; Mizuguchi, J. PKC-δ mediates interferon-α-induced apoptosis through c-Jun NH₂-terminal kinase activation. BMC Cell Biol. 2012, 13, 7. [Google Scholar] [CrossRef] [Green Version]
  70. Parent, N.; Scherer, M.; Liebisch, G.; Schmitz, G.; Bertrand, R. Protein kinase C-δ isoform mediates lysosome labilization in DNA damage-induced apoptosis. Int. J. Oncol. 2011, 38, 313–324. [Google Scholar]
  71. Leseux, L.; Laurent, G.; Laurent, C.; Rigo, M.; Blanc, A.; Olive, D.; Bezombes, C. PKCζ-mTOR pathway: A new target for rituximab therapy in follicular lymphoma. Blood 2008, 111, 285–291. [Google Scholar] [CrossRef]
  72. Arcos-Montoy, A.D.; Wegman-Ostrosky, T.; Mejía-Pérez, S.; De la Fuente-Granada, M.; Camacho-Arroyo, I.; García-Carrancá, A.; Velasco-Velázquez, M.A.; Manjarrez-Marmolejo, J.; González-Arenas, A. Progesterone receptor together with PKCα expression as prognostic factors for astrocytomas malignancy. Onco Targets Ther. 2021, 14, 3757–3768. [Google Scholar] [CrossRef] [PubMed]
  73. Chandrika, G.; Natesh, K.; Ranade, D.; Chugh, A.; Shastry, P. Suppression of the invasive potential of glioblastoma cells by mTOR inhibitors involves modulation of NFκB and PKC-α signaling. Sci. Rep. 2016, 6, 22455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Amos, S.; Mut, M.; diPierro, C.G.; Carpenter, J.E.; Xiao, A.; Kohutek, Z.A.; Redpath, G.T.; Zhao, Y.; Wang, J.; Shaffrey, M.E.; et al. Protein kinase C-α-mediated regulation of low-density lipoprotein receptor related protein and urokinase increases astrocytoma invasion. Cancer Res. 2007, 67, 10241–10251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Tang, C.; Wang, Y.; Zhang, L.; Wang, J.; Wang, W.; Han, X.; Mu, C.; Gao, D. Identification of novel LncRNA targeting Smad2/PKCα signal pathway to negatively regulate malignant progression of glioblastoma. J. Cell. Physiol. 2020, 235, 3835–3848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Valdés-Rives, S.A.; Arcos-Montoya, D.; de la Fuente-Granada, M.; Zamora-Sánchez, C.J.; Arias-Romero, L.E.; Villamar-Cruz, O.; Camacho-Arroyo, I.; Pérez-Tapia, S.M.; González-Arenas, A. LPA1 receptor promotes progesterone receptor phosphorylation through PKCα in human glioblastoma cells. Cells 2021, 10, 807. [Google Scholar] [CrossRef] [PubMed]
  77. Valdés-Rives, S.A.; de la Fuente-Granada, M.; Velasco-Velázquez, M.A.; González-Flores, O.; González-Arenas, A. LPA1 receptor activation induces PKCα nuclear translocation in glioblastoma cells. Int. J. Biochem. Cell Biol. 2019, 110, 91–102. [Google Scholar] [CrossRef]
  78. Grossman, S.A.; Alavi, J.B.; Supko, J.G.; Carson, K.A.; Priet, R.; Dorr, F.A.; Grundy, J.S.; Holmlund, J.T. Efficacy and toxicity of the antisense oligonucleotide aprinocarsen directed against protein kinase C-α delivered as a 21-day continuous intravenous infusion in patients with recurrent high-grade astrocytomas. Neuro. Oncol. 2005, 7, 32–40. [Google Scholar] [CrossRef] [Green Version]
  79. Wong, R.A.; Luo, X.; Lu, M.; An, Z.; Haas-Kogan, D.A.; Phillips, J.J.; Shokat, K.M.; Weiss, W.A.; Fan, Q.W. Cooperative blockade of PKCα and JAK2 drives apoptosis in glioblastoma. Cancer Res. 2020, 80, 709–718. [Google Scholar] [CrossRef]
  80. Cameron, A.J.; Procyk, K.J.; Leitges, M.; Parker, P.J. PKC alpha protein but not kinase activity is critical for glioma cell proliferation and survival. Int. J. Cancer 2008, 123, 769–779. [Google Scholar] [CrossRef]
  81. Baldwin, R.M.; Barrett, G.M.; Parolin, D.A.; Gillies, J.K.; Paget, J.A.; Lavictoire, S.J.; Gray, D.A.; Lorimer, I.A. Coordination of glioblastoma cell motility by PKCι. Mol. Cancer 2010, 9, 233. [Google Scholar] [CrossRef] [Green Version]
  82. Desai, S.R.; Pillai, P.P.; Patel, R.S.; McCray, A.N.; Win-Piazza, H.Y.; Acevedo-Duncan, M.E. Regulation of Cdk7 activity through a phosphatidylinositol (3)-kinase/PKC-ι-mediated signaling cascade in glioblastoma. Carcinogenesis 2012, 33, 10–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Dey, A.; Islam, S.M.A.; Patel, R.; Acevedo-Duncan, M. The interruption of atypical PKC signaling and temozolomide combination therapy against glioblastoma. Cell. Signal. 2021, 77, 109819. [Google Scholar] [CrossRef] [PubMed]
  84. Kenchappa, R.S.; Liu, Y.; Argenziano, M.G.; Banu, M.A.; Mladek, A.C.; West, R.; Luu, A.; Quiñones-Hinojosa, A.; Hambardzumyan, D.; Justilien, V.; et al. Protein kinase Cι and SRC signaling define reciprocally related subgroups of glioblastoma with distinct therapeutic vulnerabilities. Cell Rep. 2021, 37, 110054. [Google Scholar] [CrossRef] [PubMed]
  85. Baldwin, R.M.; Garratt-Lalonde, M.; Parolin, D.A.; Krzyzanowski, P.M.; Andrade, M.A.; Lorimer, I.A. Protection of glioblastoma cells from cisplatin cytotoxicity via protein kinase Cι-mediated attenuation of p38 MAP kinase signaling. Oncogene 2006, 25, 2909–2919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Phillips, E.; Lang, V.; Bohlen, J.; Bethke, F.; Puccio, L.; Tichy, D.; Herold-Mende, C.; Hielscher, T.; Lichter, P.; Goidts, V. Targeting atypical protein kinase C iota reduces viability in glioblastoma stem-like cells via a notch signaling mechanism. Int. J. Cancer 2016, 139, 1776–1787. [Google Scholar] [CrossRef]
  87. Desai, S.; Pillai, P.; Win-Piazza, H.; Acevedo-Duncan, M. PKC-ι promotes glioblastoma cell survival by phosphorylating and inhibiting BAD through a phosphatidylinositol 3-kinase pathway. Biochim. Biophys. Acta 2011, 1813, 1190–1197. [Google Scholar] [CrossRef] [Green Version]
  88. McCray, A.N.; Desai, S.; Acevedo-Duncan, M. The interruption of PKC-ι signaling and TRAIL combination therapy against glioblastoma cells. Neurochem. Res. 2014, 39, 1691–1701. [Google Scholar] [CrossRef]
  89. Mandil, R.; Ashkenazi, E.; Blass, M.; Kronfeld, I.; Kazimirsky, G.; Rosenthal, G.; Umansky, F.; Lorenzo, P.S.; Blumberg, P.M.; Brodie, C. Protein kinase Cα and protein kinase Cδ play opposite roles in the proliferation and apoptosis of glioma cells. Cancer Res. 2001, 61, 4612–4619. [Google Scholar]
  90. Assad Kahn, S.; Costa, S.L.; Gholamin, S.; Nitta, R.T.; Dubois, L.G.; Fève, M.; Zeniou, M.; Coelho, P.L.; El-Habr, E.; Cadusseau, J.; et al. The anti-hypertensive drug prazosin inhibits glioblastoma growth via the PKCδ-dependent inhibition of the AKT pathway. EMBO Mol. Med. 2016, 8, 511–526. [Google Scholar] [CrossRef]
  91. Misuth, M.; Joniova, J.; Horvath, D.; Dzurova, L.; Nichtova, Z.; Novotova, M.; Miskovsky, P.; Stroffekova, K.; Huntosova, V. The flashlights on a distinct role of protein kinase C δ: Phosphorylation of regulatory and catalytic domain upon oxidative stress in glioma cells. Cell. Signal. 2017, 34, 11–22. [Google Scholar] [CrossRef]
  92. Sharif, T.R.; Sharif, M. Overexpression of protein kinase C epsilon in astroglial brain tumor derived cell lines and primary tumor samples. Int. J. Oncol. 1999, 15, 237–243. [Google Scholar] [CrossRef] [PubMed]
  93. Toton, E.; Romaniuk, A.; Konieczna, N.; Hofmann, J.; Barciszewski, J.; Rybczynska, M. Impact of PKCε downregulation on autophagy in glioblastoma cells. BMC Cancer 2018, 18, 85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Guo, H.; Gu, F.; Li, W.; Zhang, B.; Niu, R.; Fu, L.; Zhang, N.; Ma, Y. Reduction of protein kinase Cζ inhibits migration and invasion of human glioblastoma cells. J. Neurochem. 2009, 109, 203–213. [Google Scholar] [CrossRef] [PubMed]
  95. Donson, A.M.; Banerjee, A.; Gamboni-Robertson, F.; Fleitz, J.M.; Foreman, N.K. Protein kinase C ζ isoform is critical for proliferation in human glioblastoma cell lines. J. Neurooncol. 2000, 47, 109–115. [Google Scholar] [CrossRef]
  96. Uht, R.M.; Amos, S.; Martin, P.M.; Riggan, A.E.; Hussaini, I.M. The protein kinase C-η isoform induces proliferation in glioblastoma cell lines through an ERK/Elk-1 pathway. Oncogene 2007, 26, 2885–2893. [Google Scholar] [CrossRef] [Green Version]
  97. Ali, S.; Al-Sukhun, S.; El-Rayes, B.F.; Sarkar, F.H.; Heilbrun, L.K.; Philip, P.A. Protein kinases C isozymes are differentially expressed in human breast carcinomas. Life Sci. 2009, 84, 766–771. [Google Scholar] [CrossRef]
  98. Pan, Q.; Bao, L.W.; Kleer, C.G.; Sabel, M.S.; Griffith, K.A.; Teknos, T.N.; Merajver, S.D. Protein kinase Cε is a predictive biomarker of aggressive breast cancer and a validated target for RNA interference anticancer therapy. Cancer Res. 2005, 65, 8366–8371. [Google Scholar] [CrossRef] [Green Version]
  99. Khan, K.; Safi, S.; Abbas, A.; Badshah, Y.; Dilshad, E.; Rafiq, M.; Zahra, K.; Shabbir, M. Unravelling structure, localization, and genetic crosstalk of KLF3 in human breast cancer. Biomed. Res. Int. 2020, 2020, 1354381. [Google Scholar] [CrossRef]
  100. Lønne, G.K.; Cornmark, L.; Zahirovic, I.O.; Landberg, G.; Jirström, K.; Larsson, C. PKCα expression is a marker for breast cancer aggressiveness. Mol. Cancer 2010, 9, 76. [Google Scholar] [CrossRef] [Green Version]
  101. Pham, T.N.D.; Perez White, B.E.; Zhao, H.; Mortazavi, F.; Tonetti, D.A. Protein kinase Cα enhances migration of breast cancer cells through FOXC2-mediated repression of p120-catenin. BMC Cancer 2017, 17, 832. [Google Scholar] [CrossRef]
  102. Frankel, L.B.; Lykkesfeldt, A.E.; Hansen, J.B.; Stenvang, J. Protein Kinase C α is a marker for antiestrogen resistance and is involved in the growth of tamoxifen resistant human breast cancer cells. Breast Cancer Res. Treat. 2007, 104, 165–179. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, N.; Li, Z.; Tian, F.; Feng, Y.; Huang, J.; Li, C.; Xie, F. PKCα inhibited apoptosis by decreasing the activity of JNK in MCF-7/ADR cells. Exp. Toxicol. Pathol. 2012, 64, 459–464. [Google Scholar] [CrossRef]
  104. Tonetti, D.A.; Gao, W.; Escarzaga, D.; Walters, K.; Szafran, A.; Coon, J.S. PKCα and ERβ are associated with triple-negative breast cancers in African American and Caucasian patients. Int. J. Breast Cancer 2012, 2012, 740353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Tam, W.L.; Lu, H.; Buikhuisen, J.; Soh, B.S.; Lim, E.; Reinhardt, F.; Wu, Z.J.; Krall, J.A.; Bierie, B.; Guo, W.; et al. Protein kinase C α is a central signaling node and therapeutic target for breast cancer stem cells. Cancer Cell 2013, 24, 347–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Fournier, D.B.; Chisamore, M.; Lurain, J.R.; Rademaker, A.W.; Jordan, V.C.; Tonetti, D.A. Protein kinase C α expression is inversely related to ER status in endometrial carcinoma: Possible role in AP-1-mediated proliferation of ER-negative endometrial cancer. Gynecol. Oncol. 2001, 81, 366–372. [Google Scholar] [CrossRef]
  107. Kim, C.W.; Asai, D.; Kang, J.H.; Kishimura, A.; Mori, T.; Katayama, Y. Reversal of efflux of an anticancer drug in human drug-resistant breast cancer cells by inhibition of protein kinase Cα (PKCα) activity. Tumor Biol. 2016, 37, 1901–1908. [Google Scholar] [CrossRef] [PubMed]
  108. Reedijk, M.; Odorcic, S.; Chang, L.; Zhang, H.; Miller, N.; McCready, D.R.; Lockwood, G.; Egan, S.E. High-level coexpression of JAG1 and NOTCH1 is observed in human breast cancer and is associated with poor overall survival. Cancer Res. 2005, 65, 8530–8537. [Google Scholar] [CrossRef] [Green Version]
  109. Dickson, B.C.; Mulligan, A.M.; Zhang, H.; Lockwood, G.; O’Malley, F.P.; Egan, S.E.; Reedijk, M. High level JAG1 mRNA and protein predict poor outcome in breast cancer. Mod. Pathol. 2007, 20, 85–693. [Google Scholar] [CrossRef] [Green Version]
  110. BeLow, M.; Osipo, C. Notch signaling in breast cancer: A role in drug resistance. Cells 2020, 9, 2204. [Google Scholar] [CrossRef]
  111. Pandya, K.; Wyatt, D.; Gallagher, B.; Shah, D.; Baker, A.; Bloodworth, J.; Zlobin, A.; Pannuti, A.; Green, A.; Ellis, I.O.; et al. PKCα attenuates Jagged-1-mediated notch signaling in ErbB-2-positive breast cancer to reverse trastuzumab resistance. Clin. Cancer Res. 2016, 22, 175–186. [Google Scholar] [CrossRef] [Green Version]
  112. Berardi, D.E.; Ariza Bareño, L.; Amigo, N.; Cañonero, L.; Pelagatti, M.L.N.; Motter, A.N.; Taruselli, M.A.; Díaz Bessone, M.I.; Cirigliano, S.M.; Edelstein, A.; et al. All-trans retinoic acid and protein kinase C α/β1 inhibitor combined treatment targets cancer stem cells and impairs breast tumor progression. Sci. Rep. 2021, 11, 6044. [Google Scholar] [CrossRef]
  113. Bessone, M.I.D.; Berardi, D.E.; Cirigliano, S.M.; Delbart, D.I.; Peters, M.G.; Todaro, L.B.; Urtreger, A.J. Protein Kinase C Alpha (PKCα) overexpression leads to a better response to retinoid acid therapy through Retinoic Acid Receptor Beta (RARβ) activation in mammary cancer cells. J. Cancer Res. Clin. Oncol. 2020, 146, 3241–3253. [Google Scholar] [CrossRef] [PubMed]
  114. Bailey, T.A.; Luan, H.; Tom, E.; Bielecki, T.A.; Mohapatra, B.; Ahmad, G.; George, M.; Kelly, D.L.; Natarajan, A.; Raja, S.M.; et al. A kinase inhibitor screen reveals protein kinase C-dependent endocytic recycling of ErbB2 in breast cancer cells. J. Biol. Chem. 2014, 89, 30443–30458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Allen-Petersen, B.L.; Carter, C.J.; Ohm, A.M.; Reyland, M.E. Protein kinase Cδ is required for ErbB2-driven mammary gland tumorigenesis and negatively correlates with prognosis in human breast cancer. Oncogene 2014, 33, 1306–1315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Chen, Z.; Forman, L.W.; Williams, R.M.; Faller, D.V. Protein kinase C-δ inactivation inhibits the proliferation and survival of cancer stem cells in culture and in vivo. BMC Cancer 2014, 14, 90. [Google Scholar] [CrossRef] [Green Version]
  117. Assender, J.W.; Gee, J.M.; Lewis, I.; Ellis, I.O.; Robertson, J.F.; Nicholson, R.I. Protein kinase C isoform expression as a predictor of disease outcome on endocrine therapy in breast cancer. J. Clin. Pathol. 2007, 60, 1216–1221. [Google Scholar] [CrossRef]
  118. He, Y.; Liu, J.; Durrant, D.; Yang, H.S.; Sweatman, T.; Lothstein, L.; Lee, R.M. N-benzyladriamycin-14-valerate (AD198) induces apoptosis through protein kinase C-delta-induced phosphorylation of phospholipid scramblase 3. Cancer Res. 2005, 65, 10016–10023. [Google Scholar] [CrossRef] [Green Version]
  119. Díaz Bessone, M.I.; Berardi, D.E.; Campodónico, P.B.; Todaro, L.B.; Lothstein, L.; Bal de Kier Joffé, E.D.; Urtreger, A.J. Involvement of PKC delta (PKCδ) in the resistance against different doxorubicin analogs. Breast Cancer Res. Treat. 2011, 126, 577–587. [Google Scholar] [CrossRef]
  120. Yin, J.; Liu, Z.; Li, H.; Sun, J.; Chang, X.; Liu, J.; He, S.; Li, B. Association of PKCζ expression with clinicopathological characteristics of breast cancer. PLoS ONE 2014, 9, e90811. [Google Scholar] [CrossRef] [Green Version]
  121. Smalley, T.; Islam, S.M.A.; Apostolatos, C.; Apostolatos, A.; Acevedo-Duncan, M. Analysis of PKC-ζ protein levels in normal and malignant breast tissue subtypes. Oncol. Lett. 2019, 17, 1537–1546. [Google Scholar]
  122. Belguise, K.; Cherradi, S.; Sarr, A.; Boissière, F.; Boulle, N.; Simony-Lafontaine, J.; Choesmel-Cadamuro, V.; Wang, X.; Chalbos, D. PKCθ-induced phosphorylations control the ability of Fra-1 to stimulate gene expression and cancer cell migration. Cancer Lett. 2017, 385, 97–107. [Google Scholar] [CrossRef] [PubMed]
  123. Dunn, J.; McCuaig, R.D.; Tan, A.H.Y.; Tu, W.J.; Wu, F.; Wagstaff, K.M.; Zafar, A.; Ali, S.; Diwakar, H.; Dahlstrom, J.E.; et al. Selective targeting of protein kinase C (PKC)-θ nuclear translocation reduces mesenchymal gene signatures and reinvigorates dysfunctional CD8+ T cells in immunotherapy-resistant and metastatic cancers. Cancers 2022, 14, 1596. [Google Scholar] [CrossRef] [PubMed]
  124. Byerly, J.; Halstead-Nussloch, G.; Ito, K.; Katsyv, I.; Irie, H.Y. PRKCQ promotes oncogenic growth and anoikis resistance of a subset of triple-negative breast cancer cells. Breast Cancer Res. 2016, 18, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Pal, D.; Outram, S.P.; Basu, A. Upregulation of PKCη by PKCε and PDK1 involves two distinct mechanisms and promotes breast cancer cell survival. Biochim. Biophys. Acta. 2013, 1830, 4040–4045. [Google Scholar] [CrossRef] [Green Version]
  126. Karp, G.; Abu-Ghanem, S.; Novack, V.; Mermershtain, W.; Ariad, S.; Sion-Vardy, N.; Livneh, E. Localization of PKCη in cell membranes as a predictor for breast cancer response to treatment. Onkologie 2012, 35, 260–266. [Google Scholar] [CrossRef]
  127. Zurgil, U.; Ben-Ari, A.; Rotem-Dai, N.; Karp, G.; Krasnitsky, E.; Frost, S.A.; Livneh, E. PKCη is an anti-apoptotic kinase that predicts poor prognosis in breast and lung cancer. Biochem. Soc. Trans. 2014, 42, 1519–1523. [Google Scholar] [CrossRef]
  128. Motomura, H.; Nozaki, Y.; Onaga, C.; Ozaki, A.; Tamori, S.; Shiina, T.A.; Kanai, S.; Ohira, C.; Hara, Y.; Harada, Y.; et al. High expression of c-Met, PKCλ and ALDH1A3 predicts a poor prognosis in late-stage breast cancer. Anticancer Res. 2020, 40, 35–52. [Google Scholar] [CrossRef]
  129. Nozaki, Y.; Motomura, H.; Tamori, S.; Kimura, Y.; Onaga, C.; Kanai, S.; Ishihara, Y.; Ozaki, A.; Hara, Y.; Harada, Y.; et al. High PKCλ expression is required for ALDH1-positive cancer stem cell function and indicates a poor clinical outcome in late-stage breast cancer patients. PLoS ONE 2020, 15, e0235747. [Google Scholar] [CrossRef]
  130. Motomura, H.; Tamori, S.; Yatani, M.A.; Namiki, A.; Onaga, C.; Ozaki, A.; Takasawa, R.; Mano, Y.; Sato, T.; Hara, Y.; et al. GLO 1 and PKCλ regulate ALDH1-positive breast cancer stem cell survival. Anticancer Res. 2021, 41, 5959–5971. [Google Scholar] [CrossRef]
  131. Blanchard, A.A.; Ma, X.; Wang, N.; Hombach-Klonisch, S.; Penner, C.; Ozturk, A.; Klonisch, T.; Pitz, M.; Murphy, L.; Leygue, E.; et al. Claudin 1 is highly upregulated by PKC in MCF7 human breast cancer cells and correlates positively with PKCε in patient biopsies. Transl. Oncol. 2019, 12, 561–575. [Google Scholar] [CrossRef]
  132. Azuma, K.; Ikeda, K.; Suzuki, T.; Aogi, K.; Horie-Inoue, K.; Inoue, S. TRIM47 activates NF-κB signaling via PKC-ε/PKD3 stabilization and contributes to endocrine therapy resistance in breast cancer. Proc. Natl. Acad. Sci USA 2021, 118, e2100784118. [Google Scholar] [CrossRef] [PubMed]
  133. Basu, A. Regulation of autophagy by protein kinase C-ε in breast cancer cells. Int. J. Mol. Sci. 2020, 21, 4247. [Google Scholar] [CrossRef]
  134. Wu, B.; Zhou, H.; Hu, L.; Mu, Y.; Wu, Y. Involvement of PKCα activation in TF/VIIa/PAR2-induced proliferation, migration, and survival of colon cancer cell SW620. Tumor Biol. 2013, 34, 837–846. [Google Scholar] [CrossRef]
  135. Lee, S.K.; Shehzad, A.; Jung, J.C.; Sonn, J.K.; Lee, J.T.; Park, J.W.; Lee, Y.S. Protein kinase Cα protects against multidrug resistance in human colon cancer cells. Mol. Cells 2012, 34, 61–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Hu, L.; Xia, L.; Zhou, H.; Wu, B.; Mu, Y.; Wu, Y.; Yan, J. TF/FVIIa/PAR2 promotes cell proliferation and migration via PKCα and ERK-dependent c-Jun/AP-1 pathway in colon cancer cell line SW620. Tumor Biol. 2013, 34, 2573–2581. [Google Scholar] [CrossRef] [PubMed]
  137. Zhang, L.; Jiang, B.; Zhu, N.; Tao, M.; Jun, Y.; Chen, X.; Wang, Q.; Luo, C. Mitotic checkpoint kinase Mps1/TTK predicts prognosis of colon cancer patients and regulates tumor proliferation and differentiation via PKCα/ERK1/2 and PI3K/Akt pathway. Med. Oncol. 2019, 37, 5. [Google Scholar] [CrossRef]
  138. Fang, J.Y.; Li, Z.H.; Li, Q.; Huang, W.S.; Kang, L.; Wang, J.P. Resveratrol affects protein kinase C activity and promotes apoptosis in human colon carcinoma cells. Asian Pac. J. Cancer Prev. 2012, 13, 6017–6022. [Google Scholar] [CrossRef]
  139. Cheng, Y.; Zhu, Y.; Xu, W.; Xu, J.; Yang, M.; Chen, P.; Zhao, J.; Geng, L.; Gong, S. PKCα in colon cancer cells promotes M1 macrophage polarization via MKK3/6-P38 MAPK pathway. Mol. Carcinog. 2018, 57, 1017–1029. [Google Scholar] [CrossRef]
  140. Gwak, J.; Jung, S.J.; Kang, D.I.; Kim, E.Y.; Kim, D.E.; Chung, Y.H.; Shin, J.G.; Oh, S. Stimulation of protein kinase C-α suppresses colon cancer cell proliferation by down-regulation of β-catenin. J. Cell. Mol. Med. 2009, 13, 2171–2180. [Google Scholar] [CrossRef]
  141. Oster, H.; Leitges, M. Protein kinase C α but not PKCζ suppresses intestinal tumor formation in ApcMin/+ mice. Cancer Res. 2006, 66, 6955–6963. [Google Scholar] [CrossRef] [Green Version]
  142. Suga, K.; Sugimoto, I.; Ito, H.; Hashimoto, E. Down-regulation of protein kinase C-α detected in human colorectal cancer. Biochem. Mol. Biol. Int. 1998, 44, 523–528. [Google Scholar] [CrossRef] [PubMed]
  143. Dupasquier, S.; Blache, P.; Picque Lasorsa, L.; Zhao, H.; Abraham, J.D.; Haigh, J.J.; Ychou, M.; Prévostel, C. Modulating PKCα activity to target Wnt/β-catenin signaling in colon cancer. Cancers 2019, 11, 693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Chen, S.; Wang, Y.; Zhang, Y.; Wan, Y. Low expression of PKCα and high expression of KRAS predict poor prognosis in patients with colorectal cancer. Oncol. Lett. 2016, 12, 1655–1660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Islam, S.M.A.; Dey, A.; Patel, R.; Smalley, T.; Acevedo-Duncan, M. Atypical protein kinase-C inhibitors exhibit a synergistic effect in facilitating DNA damaging effect of 5-fluorouracil in colorectal cancer cells. Biomed. Pharmacother. 2020, 121, 109665. [Google Scholar] [CrossRef]
  146. Zhang, S.; Zhang, Y.; Cheng, Q.; Ma, Z.; Gong, G.; Deng, Z.; Xu, K.; Wang, G.; Wei, Y.; Zou, X. Silencing protein kinase C ζ by microRNA-25-5p activates AMPK signaling and inhibits colorectal cancer cell proliferation. Oncotarget 2017, 8, 65329–65338. [Google Scholar] [CrossRef] [Green Version]
  147. Umemori, Y.; Kuribayashi, K.; Nirasawa, S.; Kondoh, T.; Tanaka, M.; Kobayashi, D.; Watanabe, N. Protein kinase C ζ regulates survivin expression and inhibits apoptosis in colon cancer. Int. J. Oncol. 2014, 45, 1043–1050. [Google Scholar] [CrossRef] [Green Version]
  148. Islam, S.M.A.; Patel, R.; Acevedo-Duncan, M. Protein kinase C-ζ stimulates colorectal cancer cell carcinogenesis via PKC-ζ/Rac1/Pak1/β-Catenin signaling cascade. Biochim. Biophys. Acta Mol. Cell. Res. 2018, 1865, 650–664. [Google Scholar] [CrossRef]
  149. Yeo, M.K.; Kim, J.Y.; Seong, I.O.; Kim, J.M.; Kim, K.H. Phosphorylated protein kinase C (Zeta/Lambda) expression in colorectal adenocarcinoma and its correlation with clinicopathologic characteristics and prognosis. J. Cancer 2017, 8, 3371–3377. [Google Scholar] [CrossRef]
  150. Dowling, C.M.; Phelan, J.; Callender, J.A.; Cathcart, M.C.; Mehigan, B.; McCormick, P.; Dalton, T.; Coffey, J.C.; Newton, A.C.; O’Sullivan, J.; et al. Protein kinase beta II suppresses colorectal cancer by regulating IGF-1 mediated cell survival. Oncotarget 2016, 7, 20919–20933. [Google Scholar] [CrossRef] [Green Version]
  151. Spindler, K.L.; Lindebjerg, J.; Lahn, M.; Kjaer-Frifeldt, S.; Jakobsen, A. Protein kinase C-beta II (PKC-βII) expression in patients with colorectal cancer. Int. J. Colorectal. Dis. 2009, 24, 641–645. [Google Scholar] [CrossRef]
  152. Kahl-Rainer, P.; Sedivy, R.; Marian, B. Protein kinase C tissue localization in human colonic tumors suggests a role for adenoma growth control. Gastroenterology 1996, 110, 1753–1759. [Google Scholar] [CrossRef] [PubMed]
  153. Serova, M.; Astorgues-Xerri, L.; Bieche, I.; Albert, S.; Vidaud, M.; Benhadji, K.A.; Emami, S.; Vidaud, D.; Hammel, P.; Theou-Anton, N.; et al. Epithelial-to-mesenchymal transition and oncogenic Ras expression in resistance to the protein kinase Cβ inhibitor enzastaurin in colon cancer cells. Mol. Cancer Ther. 2010, 9, 1308–1317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Perletti, G.; Marras, E.; Dondi, D.; Osti, D.; Congiu, T.; Ferrarese, R.; de Eguileor, M.; Tashjian, A.H., Jr. p21Waf1/Cip1 and p53 are downstream effectors of protein kinase Cδ in tumor suppression and differentiation in human colon cancer cells. Int. J. Cancer 2005, 113, 42–53. [Google Scholar] [CrossRef] [PubMed]
  155. Mhaidat, N.M.; Bouklihacene, M.; Thorne, R.F. 5-Fluorouracil-induced apoptosis in colorectal cancer cells is caspase-9-dependent and mediated by activation of protein kinase C-δ. Oncol. Lett. 2014, 8, 699–704. [Google Scholar] [CrossRef] [Green Version]
  156. Zhou, B.; Lu, Y.; Zhao, Z.; Shi, T.; Wu, H.; Chen, W.; Zhang, L.; Zhang, X. B7-H4 expression is upregulated by PKCδ activation and contributes to PKCδ-induced cell motility in colorectal cancer. Cancer Cell Int. 2022, 22, 147. [Google Scholar] [CrossRef]
  157. Su, C.M.; Weng, Y.S.; Kuan, L.Y.; Chen, J.H.; Hsu, F.T. Suppression of PKCδ/NF-κB signaling and apoptosis induction through extrinsic/intrinsic pathways are associated magnolol-inhibited tumor progression in colorectal cancer in vitro and in vivo. Int. J. Mol. Sci. 2020, 21, 3527. [Google Scholar] [CrossRef]
  158. Cheng, J.; He, S.; Wang, M.; Zhou, L.; Zhang, Z.; Feng, X.; Yu, Y.; Ma, J.; Dai, C.; Zhang, S.; et al. The caspase-3/PKCδ/Akt/VEGF-A signaling pathway mediates tumor repopulation during radiotherapy. Clin. Cancer Res. 2019, 25, 3732–3743. [Google Scholar] [CrossRef]
  159. Du, G.S.; Qiu, Y.; Wang, W.S.; Peng, K.; Zhang, Z.C.; Li, X.S.; Xiao, W.D.; Yang, H. Knockdown on aPKC-ι inhibits epithelial-mesenchymal transition, migration and invasion of colorectal cancer cells through Rac1-JNK pathway. Exp. Mol. Pathol. 2019, 107, 57–67. [Google Scholar] [CrossRef]
  160. Linares, J.F.; Zhang, X.; Martinez-Ordoñez, A.; Duran, A.; Kinoshita, H.; Kasashima, H.; Nakanishi, N.; Nakanishi, Y.; Carelli, R.; Cappelli, L.; et al. PKCλ/ι inhibition activates an ULK2-mediated interferon response to repress tumorigenesis. Mol. Cell 2021, 81, 4509–4526. [Google Scholar] [CrossRef]
  161. Lin, K.Y.; Fang, C.L.; Uen, Y.H.; Chang, C.C.; Lou, H.Y.; Hsieh, C.R.; Tiong, C.; Pan, S.; Chen, S.H. Overexpression of protein kinase Cα mRNA may be an independent prognostic marker for gastric carcinoma. J. Surg. Oncol. 2008, 97, 538–543. [Google Scholar] [CrossRef]
  162. Lin, S.C.; Chen, W.Y.; Lin, K.Y.; Chen, S.H.; Chang, C.C.; Lin, S.E.; Fang, C.L. Clinicopathological correlation and prognostic significance of protein kinase Cα overexpression in human gastric carcinoma. PLoS ONE 2013, 8, e56675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Wu, D.L.; Sui, F.Y.; Du, C.; Zhang, C.W.; Hui, B.; Xu, S.L.; Lu, H.Z.; Song, G.J. Antisense e expression of PKCα improved sensitivity of SGC7901/VCR cells to doxorubicin. World J. Gastroenterol. 2009, 15, 1259–1263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Han, Y.; Han, Z.Y.; Zhou, X.M.; Shi, R.; Zheng, Y.; Shi, Y.Q.; Miao, J.Y.; Pan, B.R.; Fan, D.M. Expression and function of classical protein kinase C isoenzymes in gastric cancer cell line and its drug-resistant sublines. World J. Gastroenterol. 2002, 8, 441–445. [Google Scholar] [CrossRef]
  165. Tseng, L.L.; Cheng, H.H.; Yeh, T.S.; Huang, S.C.; Syu, Y.Y.; Chuu, C.P.; Yuh, C.H.; Kung, H.J.; Wang, W.C. Targeting the histone demethylase PHF8-mediated PKCα-Src-PTEN axis in HER2-negative gastric cancer. Proc. Natl. Acad. Sci. USA 2020, 117, 24859–24866. [Google Scholar] [CrossRef]
  166. Hashimoto, I.; Sakamaki, K.; Oue, N.; Kimura, Y.; Hiroshima, Y.; Hara, K.; Maezawa, Y.; Kano, K.; Aoyama, T.; Yamada, T.; et al. Clinical significance of PRKCI gene expression in cancerous tissue in patients with gastric cancer. Anticancer Res. 2019, 39, 5715–5720. [Google Scholar] [CrossRef] [PubMed]
  167. Wu, L.; Li, Y.; Xu, X.M.; Zhu, X. Circular RNA circ-PRKCI promotes cell proliferation and invasion by binding to microRNA-545 in gastric cancer. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 9418–9426. [Google Scholar] [PubMed]
  168. Batsaikhan, B.E.; Yoshikawa, K.; Kurita, N.; Iwata, T.; Takasu, C.; Kashihara, H.; Shimada, M. Expression of Stathmin1 in gastric adenocarcinoma. Anticancer Res. 2014, 34, 4217–4421. [Google Scholar]
  169. Takagawa, R.; Akimoto, K.; Ichikawa, Y.; Akiyama, H.; Kojima, Y.; Ishiguro, H.; Inayama, Y.; Aoki, I.; Kunisaki, C.; Endo, I.; et al. High expression of atypical protein kinase C λ/ι in gastric cancer as a prognostic factor for recurrence. Ann. Surg. Oncol. 2010, 17, 81–88. [Google Scholar] [CrossRef]
  170. Johnson, D.E.; Burtness, B.; Leemans, C.R.; Lui, V.W.Y.; Bauman, J.E.; Grandis, J.R. Head and neck squamous cell carcinoma. Nat. Rev. Dis. Primers 2020, 6, 92. [Google Scholar] [CrossRef]
  171. Martínez-Gimeno, C.; Díaz-Meco, M.T.; Domínguez, I.; Moscat, J. Alterations in levels of different protein kinase C isotypes and their influence on behavior of squamous cell carcinoma of the oral cavity: εPKC, a novel prognostic factor for relapse and survival. Head Neck. 1995, 17, 516–525. [Google Scholar] [CrossRef]
  172. Chu, P.Y.; Hsu, N.C.; Lin, S.H.; Hou, M.F.; Yeh, K.T. High nuclear protein kinase CβII expression is a marker of disease recurrence in oral squamous cell carcinoma. Anticancer Res. 2012, 32, 3987–3991. [Google Scholar] [PubMed]
  173. Gao, W.; Guo, H.; Niu, M.; Zheng, X.; Zhang, Y.; Xue, X.; Bo, Y.; Guan, X.; Li, Z.; Guo, Y.; et al. circPARD3 drives malignant progression and chemoresistance of laryngeal squamous cell carcinoma by inhibiting autophagy through the PRKCI-Akt-mTOR pathway. Mol. Cancer 2020, 19, 166. [Google Scholar] [CrossRef] [PubMed]
  174. Parzefall, T.; Schnoell, J.; Monschein, L.; Foki, E.; Liu, D.T.; Frohne, A.; Grasl, S.; Pammer, J.; Lucas, T.; Kadletz, L.; et al. PRKCA overexpression is frequent in young oral tongue squamous cell carcinoma patients and is associated with poor prognosis. Cancers 2021, 13, 2082. [Google Scholar] [CrossRef] [PubMed]
  175. Zhen-jin, Z.; Peng, L.; Fa-yu, L.; Liyan, S.; Chang-fu, S. PKCα take part in CCR7/NF-κB autocrine signaling loop in CCR7-positive squamous cell carcinoma of head and neck. Mol. Cell Biochem. 2011, 357, 181–187. [Google Scholar] [CrossRef]
  176. Cohen, E.E.; Zhu, H.; Lingen, M.W.; Martin, L.E.; Kuo, W.L.; Choi, E.A.; Kocherginsky, M.; Parker, J.S.; Chung, C.H.; Rosner, M.R. A feed-forward loop involving protein kinase Cα and microRNAs regulates tumor cell cycle. Cancer Res. 2009, 69, 65–74. [Google Scholar] [CrossRef] [Green Version]
  177. Baba, J.; Kioi, M.; Akimoto, K.; Nagashima, Y.; Taguri, M.; Inayama, Y.; Aoki, I.; Ohno, S.; Mitsudo, K.; Tohnai, I. Atypical protein Kinase Cλ/ι expression is associated with malignancy of oral squamous cell carcinoma. Anticancer Res. 2018, 38, 6291–6297. [Google Scholar] [CrossRef]
  178. Chu, P.Y.; Hsu, N.C.; Tai, H.C.; Yeh, C.M.; Lin, S.H.; Hou, M.F.; Yeh, K.T. High nuclear protein kinase Cθ expression may correlate with disease recurrence and poor survival in oral squamous cell carcinoma. Hum. Pathol. 2012, 43, 276–281. [Google Scholar] [CrossRef]
  179. Caspa Gokulan, R.; Devaraj, H. Stem cell markers CXCR-4 and CD133 predict aggressive phenotype and their double positivity indicates poor prognosis of oral squamous cell carcinoma. Cancers 2021, 13, 5895. [Google Scholar] [CrossRef]
  180. Tsai, J.H.; Tsai, M.T.; Su, W.W.; Chen, Y.L.; Wu, T.T.; Hsieh, Y.S.; Huang, C.Y.; Yeh, K.T.; Liu, J.Y. Expression of protein kinase Cα in biopsies and surgical specimens of human hepatocellular carcinoma. Chin. J. Physiol. 2005, 48, 139–143. [Google Scholar]
  181. Wu, T.T.; Hsieh, Y.H.; Wu, C.C.; Hsieh, Y.S.; Huang, C.Y.; Liu, J.Y. Overexpression of protein kinase Cα mRNA in human hepatocellular carcinoma: A potential marker of disease prognosis. Clin. Chim. Acta 2007, 382, 54–58. [Google Scholar] [CrossRef]
  182. Wu, T.T.; Hsieh, Y.H.; Hsieh, Y.S.; Liu, J.Y. Reduction of PKCα decreases cell proliferation, migration, and invasion of human malignant hepatocellular carcinoma. J. Cell. Biochem. 2008, 103, 9–20. [Google Scholar] [CrossRef] [PubMed]
  183. Lin, M.; Liu, Y.; Ding, X.; Ke, Q.; Shi, J.; Ma, Z.; Gu, H.; Wang, H.; Zhang, C.; Yang, C.; et al. E2F1 transactivates IQGAP3 and promotes proliferation of hepatocellular carcinoma cells through IQGAP3-mediated PKC-alpha activation. Am. J. Cancer Res. 2019, 9, 285–299. [Google Scholar] [PubMed]
  184. Wang, J.; Shao, M.; Liu, M.; Peng, P.; Li, L.; Wu, W.; Wang, L.; Duan, F.; Zhang, M.; Song, S.; et al. PKCα promotes generation of reactive oxygen species via DUOX2 in hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2015, 463, 839–845. [Google Scholar] [CrossRef]
  185. Wei, C.Y.; Zhu, M.X.; Zhang, P.F.; Huang, X.Y.; Wan, J.K.; Yao, X.Z.; Hu, Z.T.; Chai, X.Q.; Peng, R.; Yang, X.; et al. PKCα/ZFP64/CSF1 axis resets the tumor microenvironment and fuels anti-PD1 resistance in hepatocellular carcinoma. J. Hepatol. 2022, 77, 163–176. [Google Scholar] [CrossRef] [PubMed]
  186. Wu, C.H.; Hsu, F.T.; Chao, T.L.; Lee, Y.H.; Kuo, Y.C. Revealing the suppressive role of protein kinase Cδ and p38 mitogen-activated protein kinase (MAPK)/NF-κB axis associates with lenvatinib-inhibited progression in hepatocellular carcinoma in vitro and in vivo. Biomed. Pharmacother. 2022, 145, 112437. [Google Scholar] [CrossRef]
  187. Takai, S.; Matsushima-Nishiwaki, R.; Tokuda, H.; Yasuda, E.; Toyoda, H.; Kaneoka, Y.; Yamaguchi, A.; Kumada, T.; Kozawa, O. Protein kinase Cδ regulates the phosphorylation of heat shock protein 27 in human hepatocellular carcinoma. Life Sci. 2007, 81, 585–591. [Google Scholar] [CrossRef] [PubMed]
  188. Lee, S.E.; Yang, H.; Jeong, S.I.; Jin, Y.H.; Park, C.S.; Park, Y.S. Induction of heme oxygenase-1 inhibits cell death in crotonaldehyde-stimulated HepG2 cells via the PKC-δ-p38-Nrf2 pathway. PLoS ONE 2012, 7, e41676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Mandal, J.P.; Shiue, C.N.; Chen, Y.C.; Lee, M.C.; Yang, H.H.; Chang, H.H.; Hu, C.T.; Liao, P.C.; Hui, L.C.; You, R.I.; et al. PKCδ mediates mitochondrial ROS generation and oxidation of HSP60 to relieve RKIP inhibition on MAPK pathway for HCC progression. Free Radic. Biol. Med. 2021, 163, 69–87. [Google Scholar] [CrossRef]
  190. Cao, M.; Gao, J.; Zhou, H.; Huang, J.; You, A.; Guo, Z.; Fang, F.; Zhang, W.; Song, T.; Zhang, T. HIF-2α regulates CDCP1 to promote PKCδ-mediated migration in hepatocellular carcinoma. Tumor Biol. 2016, 37, 1651–1662. [Google Scholar] [CrossRef]
  191. Kudo, Y.; Sugimoto, M.; Arias, E.; Kasashima, H.; Cordes, T.; Linares, J.F.; Duran, A.; Nakanishi, Y.; Nakanishi, N.; L’Hermitte, A.; et al. PKCλ/ι loss induces autophagy, oxidative phosphorylation, and NRF2 to promote liver cancer progression. Cancer Cell 2020, 38, 247–262. [Google Scholar] [CrossRef]
  192. Moscat, J.; Diaz-Meco, M.T. The interplay between PRKCI/PKCλ/ι, SQSTM1/p62, and autophagy orchestrates the oxidative metabolic response that drives liver cancer. Autophagy 2020, 16, 1915–1917. [Google Scholar] [CrossRef] [PubMed]
  193. Lu, H.C.; Chou, F.P.; Yeh, K.T.; Chang, Y.S.; Hsu, N.C.; Chang, J.G. Expression of protein kinase C family in human hepatocellular carcinoma. Pathol. Oncol. Res. 2010, 16, 385–391. [Google Scholar] [CrossRef] [PubMed]
  194. Huang, W.; Mehta, D.; Sif, S.; Kent, L.N.; Jacob, S.T.; Ghoshal, K.; Mehta, K.D. Dietary fat/cholesterol-sensitive PKCβ-RB signaling: Potential role in NASH/HCC axis. Oncotarget 2017, 8, 73757–73765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Guo, K.; Li, Y.; Kang, X.; Sun, L.; Cui, J.; Gao, D.; Liu, Y. Role of PKCβ in hepatocellular carcinoma cells migration and invasion in vitro: A potential therapeutic target. Clin. Exp. Metastasis 2009, 26, 189–195. [Google Scholar] [CrossRef]
  196. Lu, H.C.; Chou, F.P.; Yeh, K.T.; Chang, Y.S.; Hsu, N.C.; Chang, J.G. Analysing the expression of protein kinase Cη in human hepatocellular carcinoma. Pathology 2009, 41, 626–629. [Google Scholar] [CrossRef]
  197. Singhal, S.S.; Wickramarachchi, D.; Singhal, J.; Yadav, S.; Awasthi, Y.C.; Awasthi, S. Determinants of differential doxorubicin sensitivity between SCLC and NSCLC. FEBS Lett. 2006, 580, 2258–2264. [Google Scholar] [CrossRef] [Green Version]
  198. Lang, W.; Wang, H.; Ding, L.; Xiao, L. Cooperation between PKC-α and PKC-ε in the regulation of JNK activation in human lung cancer cells. Cell. Signal. 2004, 16, 457–467. [Google Scholar] [CrossRef]
  199. Lahn, M.; Su, C.; Li, S.; Chedid, M.; Hanna, K.R.; Graff, J.R.; Sandusky, G.E.; Ma, D.; Niyikiza, C.; Sundell, K.L.; et al. Expression levels of protein kinase C-α in non-small-cell lung cancer. Clin. Lung Cancer 2004, 6, 184–189. [Google Scholar] [CrossRef]
  200. Tzeng, H.T.; Li, T.H.; Tang, Y.A.; Tsai, C.H.; Frank Lu, P.J.; Lai, W.W.; Chiang, C.W.; Wang, Y.C. Phosphorylation of Rab37 by protein kinase Cα inhibits the exocytosis function and metastasis suppression activity of Rab37. Oncotarget 2017, 8, 108556–108570. [Google Scholar] [CrossRef]
  201. Salama, M.F.; Liu, M.; Clarke, C.J.; Espaillat, M.P.; Haley, J.D.; Jin, T.; Wang, D.; Obeid, L.M.; Hannun, Y.A. PKCα is required for Akt-mTORC1 activation in non-small cell lung carcinoma (NSCLC) with EGFR mutation. Oncogene 2019, 38, 7311–7328. [Google Scholar] [CrossRef]
  202. Gao, X.; Xu, F.; Zhang, H.T.; Chen, M.; Huang, W.; Zhang, Q.; Zeng, Q.; Liu, L. PKCα-GSK3β-NF-κB signaling pathway and the possible involvement of TRIM21 in TRAIL-induced apoptosis. Biochem. Cell Biol. 2016, 94, 256–264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Cheng, X.D.; Gu, J.F.; Yuan, J.R.; Feng, L.; Jia, X.B. Suppression of A549 cell proliferation and metastasis by calycosin via inhibition of the PKC-α/ERK1/2 pathway: An in vitro investigation. Mol. Med. Rep. 2015, 12, 7992–8002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Abera, M.B.; Kazanietz, M.G. Protein kinase Cα mediates erlotinib resistance in lung cancer cells. Mol. Pharmacol. 2015, 87, 832–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Kang, J.H. Protein kinase C (PKC) isozymes and cancer. New J. Sci. 2014, 2014, 231418. [Google Scholar]
  206. Halvorsen, A.R.; Haugen, M.H.; Öjlert, Å.K.; Lund-Iversen, M.; Jørgensen, L.; Solberg, S.; Mælandsmo, G.M.; Brustugun, O.T.; Helland, Å. Protein kinase C isozymes associated with relapse free survival in non-small cell lung cancer patients. Front. Oncol. 2020, 10, 590755. [Google Scholar] [CrossRef]
  207. Hill, K.S.; Erdogan, E.; Khoor, A.; Walsh, M.P.; Leitges, M.; Murray, N.R.; Fields, A.P. Protein kinase Cα suppresses Kras-mediated lung tumor formation through activation of a p38 MAPK-TGFβ signaling axis. Oncogene 2014, 33, 2134–2144. [Google Scholar] [CrossRef] [Green Version]
  208. Tsai, J.Y.; Rédei, D.; Hohmann, J.; Wu, C.C. 12-Deoxyphorbol esters induce growth arrest and apoptosis in human lung cancer A549 cells via activation of PKC-δ/PKD/ERK signaling pathway. Int. J. Mol. Sci. 2020, 21, 7579. [Google Scholar] [CrossRef]
  209. Iitaka, D.; Moodley, S.; Shimizu, H.; Bai, X.H.; Liu, M. PKCδ-iPLA2-PGE2-PPARγ signaling cascade mediates TNF-α induced Claudin 1 expression in human lung carcinoma cells. Cell. Signal. 2015, 27, 568–577. [Google Scholar] [CrossRef]
  210. Zhang, H.; Okamoto, M.; Panzhinskiy, E.; Zawada, W.M.; Das, M. PKCδ/midkine pathway drives hypoxia-induced proliferation and differentiation of human lung epithelial cells. Am. J. Physiol. Cell Physiol. 2014, 306, C648–C658. [Google Scholar] [CrossRef] [Green Version]
  211. Yueh, P.F.; Lee, Y.H.; Chiang, I.T.; Chen, W.T.; Lan, K.L.; Chen, C.H.; Hsu, F.T. Suppression of EGFR/PKC-δ/NF-κB signaling associated with imipramine-inhibited progression of non-small cell lung cancer. Front. Oncol. 2021, 11, 735183. [Google Scholar] [CrossRef] [PubMed]
  212. Baek, J.H.; Yun, H.S.; Kwon, G.T.; Lee, J.; Kim, J.Y.; Jo, Y.; Cho, J.M.; Lee, C.W.; Song, J.Y.; Ahn, J.; et al. PLOD3 suppression exerts an anti-tumor effect on human lung cancer cells by modulating the PKC-delta signaling pathway. Cell Death Dis. 2019, 10, 156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Lee, P.C.; Fang, Y.F.; Yamaguchi, H.; Wang, W.J.; Chen, T.C.; Hong, X.L.; Ke, B.; Xia, W.; Wei, Y.; Zha, Z.; et al. Targeting PKCδ as a therapeutic strategy against heterogeneous mechanisms of EGFR inhibitor resistance in EGFR-mutant lung cancer. Cancer Cell 2018, 34, 954–969. [Google Scholar] [CrossRef] [Green Version]
  214. Bae, K.M.; Wang, H.; Jiang, G.; Chen, M.G.; Lu, L.; Xiao, L. Protein kinase Cε is overexpressed in primary human non-small cell lung cancers and functionally required for proliferation of non-small cell lung cancer cells in a p21/Cip1-dependent manner. Cancer Res. 2007, 67, 6053–6063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Garg, R.; Cooke, M.; Benavides, F.; Abba, M.C.; Cicchini, M.; Feldser, D.M.; Kazanietz, M.G. PKC ε is required for KRAS-driven lung tumorigenesis. Cancer Res. 2020, 80, 5166–5173. [Google Scholar] [CrossRef] [PubMed]
  216. Caino, M.C.; Lopez-Haber, C.; Kim, J.; Mochly-Rosen, D.; Kazanietz, M.G. Proteins kinase Cɛ is required for non-small cell lung carcinoma growth and regulates the expression of apoptotic genes. Oncogene 2012, 31, 2593–2600. [Google Scholar] [CrossRef] [Green Version]
  217. Pardo, O.E.; Wellbrock, C.; Khanzada, U.K.; Aubert, M.; Arozarena, I.; Davidson, S.; Bowen, F.; Parker, P.J.; Filonenko, V.V.; Gout, I.T.; et al. FGF-2 protects small cell lung cancer cells from apoptosis through a complex involving PKCε, B-Raf and S6K2. EMBO J. 2006, 25, 3078–3088. [Google Scholar] [CrossRef]
  218. Liu, L.; Lei, B.; Wang, L.; Chang, C.; Yang, H.; Liu, J.; Huang, G.; Xie, W. Protein kinase C-iota-mediated glycolysis promotes non-small-cell lung cancer progression. Onco Targets Ther. 2019, 12, 5835–5848. [Google Scholar] [CrossRef] [Green Version]
  219. Krasnitsky, E.; Baumfeld, Y.; Freedman, J.; Sion-Vardy, N.; Ariad, S.; Novack, V.; Livneh, E. PKCη is a novel prognostic marker in non-small cell lung cancer. Anticancer Res. 2012, 32, 1507–1513. [Google Scholar]
  220. Lemjabbar-Alaoui, H.; Sidhu, S.S.; Mengistab, A.; Gallup, M.; Basbaum, C. TACE/ADAM-17 phosphorylation by PKC-epsilon mediates premalignant changes in tobacco smoke-exposed lung cells. PLoS ONE 2011, 6, e17489. [Google Scholar] [CrossRef] [Green Version]
  221. Jin, Z.; Xin, M.; Deng, X. Survival function of protein kinase Cι as a novel nitrosamine 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone-activated Bad kinase. J. Biol. Chem. 2005, 280, 16045–16052. [Google Scholar] [CrossRef] [Green Version]
  222. Zhao, L.J.; Xu, H.; Qu, J.W.; Zhao, W.Z.; Zhao, Y.B.; Wang, J.H. Modulation of drug resistance in ovarian cancer cells by inhibition of protein kinase C-alpha (PKC-α) with small interference RNA (siRNA) agents. Asian Pac. J. Cancer Prev. 2012, 13, 3631–3636. [Google Scholar] [CrossRef] [PubMed]
  223. Wang, N.N.; Zhao, L.J.; Wu, L.N.; He, M.F.; Qu, J.W.; Zhao, Y.B.; Zhao, W.Z.; Li, J.S.; Wang, J.H. Mechanistic analysis of taxol-induced multidrug resistance in an ovarian cancer cell line. Asian Pac. J. Cancer Prev. 2013, 14, 4983–4988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Qi, H.; Sun, B.; Zhao, X.; Du, J.; Gu, Q.; Liu, Y.; Cheng, R.; Dong, X. Wnt5a promotes vasculogenic mimicry and epithelial-mesenchymal transition via protein kinase Cα in epithelial ovarian cancer. Oncol. Rep. 2014, 32, 771–779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Mahanivong, C.; Chen, H.M.; Yee, S.W.; Pan, Z.K.; Dong, Z.; Huang, S. Protein kinase Cα-CARMA3 signaling axis links Ras to NF-κB for lysophosphatidic acid-induced urokinase plasminogen activator expression in ovarian cancer cells. Oncogene 2008, 27, 1273–1280. [Google Scholar] [CrossRef] [Green Version]
  226. Lili, X.; Xiaoyu, T. Expression of PKCα, PKCε, and P-gp in epithelial ovarian carcinoma and the clinical significance. Eur. J. Gynaecol. Oncol. 2015, 36, 181–185. [Google Scholar]
  227. Weichert, W.; Gekeler, V.; Denkert, C.; Dietel, M.; Hauptmann, S. Protein kinase C isoform expression in ovarian carcinoma correlates with indicators of poor prognosis. Int. J. Oncol. 2003, 23, 633–639. [Google Scholar] [CrossRef]
  228. Zhang, L.; Huang, J.; Yang, N.; Liang, S.; Barchetti, A.; Giannakakis, A.; Cadungog, M.G.; O’Brien-Jenkins, A.; Massobrio, M.; Roby, K.F.; et al. Integrative genomic analysis of protein kinase C (PKC) family identifies PKCι as a biomarker and potential oncogene in ovarian carcinoma. Cancer Res. 2006, 66, 4627–4635. [Google Scholar] [CrossRef] [Green Version]
  229. Eder, A.M.; Sui, X.; Rosen, D.G.; Nolden, L.K.; Cheng, K.W.; Lahad, J.P.; Kango-Singh, M.; Lu, K.H.; Warneke, C.L.; Atkinson, E.N.; et al. Atypical PKCι contributes to poor prognosis through loss of apical-basal polarity and cyclin E overexpression in ovarian cancer. Proc. Natl. Acad. Sci USA 2005, 102, 12519–12524. [Google Scholar] [CrossRef] [Green Version]
  230. Rehmani, H.; Li, Y.; Li, T.; Padia, R.; Calbay, O.; Jin, L.; Chen, H.; Huang, S. Addiction to protein kinase Cι due to PRKCI gene amplification can be exploited for an aptamer-based targeted therapy in ovarian cancer. Signal. Transduct. Target Ther. 2020, 5, 140. [Google Scholar] [CrossRef]
  231. Sarkar, S.; Bristow, C.A.; Dey, P.; Rai, K.; Perets, R.; Ramirez-Cardenas, A.; Malasi, S.; Huang-Hobbs, E.; Haemmerle, M.; Wu, S.Y.; et al. PRKCI promotes immune suppression in ovarian cancer. Genes Dev. 2017, 31, 1109–1121. [Google Scholar] [CrossRef] [Green Version]
  232. Wang, Y.; Justilien, V.; Brennan, K.I.; Jamieson, L.; Murray, N.R.; Fields, A.P. PKCι regulates nuclear YAP1 localization and ovarian cancer tumorigenesis. Oncogene 2017, 36, 534–545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Nazarenko, I.; Jenny, M.; Keil, J.; Gieseler, C.; Weisshaupt, K.; Sehouli, J.; Legewie, S.; Herbst, L.; Weichert, W.; Darb-Esfahani, S.; et al. Atypical protein kinase C ζ exhibits a proapoptotic function in ovarian cancer. Mol. Cancer Res. 2010, 8, 919–934. [Google Scholar] [CrossRef]
  234. Smalley, T.; Metcalf, R.; Patel, R.; Islam, S.M.A.; Bommareddy, R.R.; Acevedo-Duncan, M. The atypical protein kinase C small molecule inhibitor ζ-Stat, and its effects on invasion through decreases in PKC-ζ protein expression. Front. Oncol. 2020, 10, 209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Baffi, T.R.; Van, A.N.; Zhao, W.; Mills, G.B.; Newton, A.C. Protein kinase C quality control by phosphatase PHLPP1 unveils loss-of-function mechanism in cancer. Mol. Cell. 2019, 74, 378–392.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Lin, R.; Bao, X.; Wang, H.; Zhu, S.; Liu, Z.; Chen, Q.; Ai, K.; Shi, B. TRPM2 promotes pancreatic cancer by PKC/MAPK pathway. Cell Death Dis. 2021, 12, 585. [Google Scholar] [CrossRef] [PubMed]
  237. Kim, S.Y.; Park, S.; Yoo, S.; Rho, J.K.; Jun, E.S.; Chang, S.; Kim, K.K.; Kim, S.C.; Kim, I. Downregulation of X-linked inhibitor of apoptosis protein by ′7-Benzylidenenaltrexone maleate′ sensitizes pancreatic cancer cells to TRAIL-induced apoptosis. Oncotarget 2017, 8, 61057–61071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Chow, J.Y.; Dong, H.; Quach, K.T.; Van Nguyen, P.N.; Chen, K.; Carethers, J.M. TGF-β mediates PTEN suppression and cell motility through calcium-dependent PKC-α activation in pancreatic cancer cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2008, 294, G899–G905. [Google Scholar] [CrossRef] [Green Version]
  239. Ma, J.; Xue, H.; He, L.H.; Wang, L.Y.; Wang, X.J.; Li, X.; Zhang, L. The role and mechanism of autophagy in pancreatic cancer: An update review. Cancer Manag. Res. 2021, 13, 8231–8240. [Google Scholar] [CrossRef]
  240. Jia, S.; Xu, X.; Zhou, S.; Chen, Y.; Ding, G.; Cao, L. Fisetin induces autophagy in pancreatic cancer cells via endoplasmic reticulum stress- and mitochondrial stress-dependent pathways. Cell Death Dis. 2019, 10, 142. [Google Scholar] [CrossRef] [Green Version]
  241. Kyuno, D.; Kojima, T.; Ito, T.; Yamaguchi, H.; Tsujiwaki, M.; Takasawa, A.; Murata, M.; Tanak, A.S.; Hirata, K.; Sawada, N. Protein kinase Cα inhibitor enhances the sensitivity of human pancreatic cancer HPAC cells to Clostridium perfringens enterotoxin via claudin-4. Cell Tissue Res. 2011, 346, 369–381. [Google Scholar] [CrossRef]
  242. Taniuchi, K.; Yokotani, K.; Saibara, T. BART inhibits pancreatic cancer cell invasion by PKCα inactivation through binding to ANX7. PLoS ONE 2012, 7, e35674. [Google Scholar] [CrossRef] [PubMed]
  243. Xie, X.; Wu, M.Y.; Shou, L.M.; Chen, L.P.; Gong, F.R.; Chen, K.; Li, D.M.; Duan, W.M.; Xie, Y.F.; Mao, Y.X.; et al. Tamoxifen enhances the anticancer effect of cantharidin and norcantharidin in pancreatic cancer cell lines through inhibition of the protein kinase C signaling pathway. Oncol. Lett. 2015, 9, 837–844. [Google Scholar] [CrossRef] [PubMed]
  244. Ganapathy, S.; Peng, B.; Shen, L.; Yu, T.; Lafontant, J.; Li, P.; Xiong, R.; Makriyannis, A.; Chen, C. Suppression of PKC causes oncogenic stress for triggering apoptosis in cancer cells. Oncotarget 2017, 8, 30992–31002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Kato, S.; Akimoto, K.; Nagashima, Y.; Ishiguro, H.; Kubota, K.; Kobayashi, N.; Hosono, K.; Watanabe, S.; Sekino, Y.; Sato, T.; et al. aPKCλ/ι is a beneficial prognostic marker for pancreatic neoplasms. Pancreatology 2013, 13, 360–368. [Google Scholar] [CrossRef]
  246. Abdelatty, A.; Fang, D.; Wei, G.; Wu, F.; Zhang, C.; Xu, H.; Yao, C.; Wang, Y.; Xia, H. PKCι is a promising prognosis biomarker and therapeutic target for pancreatic cancer. Pathobiology 2022, 4, 1–12. [Google Scholar] [CrossRef]
  247. Scotti, M.L.; Bamlet, W.R.; Smyrk, T.C.; Fields, A.P.; Murray, N.R. Protein kinase Cι is required for pancreatic cancer cell transformed growth and tumorigenesis. Cancer Res. 2010, 70, 2064–2074. [Google Scholar] [CrossRef] [Green Version]
  248. Wang, P.; Zhang, H.; Yang, J.; Li, Z.; Wang, Y.; Leng, X.; Ganapathy, S.; Isakson, P.; Chen, C.; Zhu, T. Mu-KRAS attenuates Hippo signaling pathway through PKCι to sustain the growth of pancreatic cancer. J. Cell. Physiol. 2020, 235, 408–420. [Google Scholar] [CrossRef]
  249. Butler, A.M.; Scotti Buzhardt, M.L.; Erdogan, E.; Li, S.; Inman, K.S.; Fields, A.P.; Murray, N.R. A small molecule inhibitor of atypical protein kinase C signaling inhibits pancreatic cancer cell transformed growth and invasion. Oncotarget 2015, 6, 15297–15310. [Google Scholar] [CrossRef] [Green Version]
  250. Yang, J.; Wang, J.; Zhang, H.; Li, C.; Chen, C.; Zhu, T. Transcription factor Sp1 is upregulated by PKCι to drive the expression of YAP1 during pancreatic carcinogenesis. Carcinogenesis 2021, 42, 344–356. [Google Scholar] [CrossRef]
  251. Laudanna, C.; Sorio, C.; Tecchio, C.; Butcher, E.C.; Bonora, A.; Bassi, C.; Scarpa, A. Motility analysis of pancreatic adenocarcinoma cells reveals a role for the atypical ζ isoform of protein kinase C in cancer cell movement. Lab. Invest. 2003, 83, 1155–1163. [Google Scholar] [CrossRef] [Green Version]
  252. Ryota, H.; Ishida, M.; Ebisu, Y.; Yanagimoto, H.; Yamamoto, T.; Kosaka, H.; Hirooka, S.; Yamaki, S.; Kotsuka, M.; Matsui, Y.; et al. Clinicopathological characteristics of pancreatic ductal adenocarcinoma with invasive micropapillary carcinoma component with emphasis on the usefulness of PKCζ immunostaining for detection of reverse polarity. Oncol. Lett. 2021, 22, 525. [Google Scholar] [CrossRef] [PubMed]
  253. Mauro, L.V.; Grossoni, V.C.; Urtreger, A.J.; Yang, C.; Colombo, L.L.; Morandi, A.; Pallotta, M.G.; Kazanietz, M.G.; Bal de Kier Joffé, E.D.; Puricelli, L.L. PKC delta (PKCδ) promotes tumoral progression of human ductal pancreatic cancer. Pancreas 2010, 39, e31–e41. [Google Scholar] [CrossRef] [PubMed]
  254. Wang, J.; Jin, W.; Zhou, X.; Li, J.; Xu, C.; Ma, Z.; Wang, J.; Qin, L.; Zhou, B.; Ding, W.; et al. Identification, structure-activity relationships of marine-derived indolocarbazoles, and a dual PKCθ/δ inhibitor with potent antipancreatic cancer efficacy. J. Med. Chem. 2020, 63, 12978–12991. [Google Scholar] [CrossRef] [PubMed]
  255. Huang, H.L.; Wu, H.Y.; Chu, P.C.; Lai, I.L.; Huang, P.H.; Kulp, S.K.; Pan, S.L.; Teng, C.M.; Chen, C.S. Role of integrin-linked kinase in regulating the protein stability of the MUC1-C oncoprotein in pancreatic cancer cells. Oncogenesis 2017, 6, e359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Shi, G.; Zhu, L.; Sun, Y.; Bettencourt, R.; Damsz, B.; Hruban, R.H.; Konieczny, S.F. Loss of the acinar-restricted transcription factor Mist1 accelerates Kras-induced pancreatic intraepithelial neoplasia. Gastroenterology 2009, 136, 1368–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Johnson, C.L.; Peat, J.M.; Volante, S.N.; Wang, R.; McLean, C.A.; Pin, C.L. Activation of protein kinase Cδ leads to increased pancreatic acinar cell dedifferentiation in the absence of MIST1. J. Pathol. 2012, 228, 351–365. [Google Scholar] [CrossRef]
  258. Cheng, J.; Tian, L.; Ma, J.; Gong, Y.; Zhang, Z.; Chen, Z.; Xu, B.; Xiong, H.; Li, C.; Huang, Q. Dying tumor cells stimulate proliferation of living tumor cells via caspase-dependent protein kinase Cδ activation in pancreatic ductal adenocarcinoma. Mol. Oncol. 2015, 9, 105–114. [Google Scholar] [CrossRef]
  259. Singh, B.N.; Kumar, D.; Shankar, S.; Srivastava, R.K. Rottlerin induces autophagy which leads to apoptotic cell death through inhibition of PI3K/Akt/mTOR pathway in human pancreatic cancer stem cells. Biochem. Pharmacol. 2012, 84, 1154–1163. [Google Scholar] [CrossRef]
  260. Sorescu, G.P.; Forman, L.W.; Faller, D.V. Effect of inhibition of protein kinase C delta (PKCδ) on pancreatic cancer cells. J. Clin. Oncol. 2012, 30, e14591. [Google Scholar] [CrossRef]
  261. Takahashi, T.; Uehara, H.; Ogawa, H.; Umemoto, H.; Bando, Y.; Izumi, K. Inhibition of EP2/EP4 signaling abrogates IGF-1R-mediated cancer cell growth: Involvement of protein kinase C-θ activation. Oncotarget 2015, 6, 4829–4844. [Google Scholar] [CrossRef] [Green Version]
  262. Banales, J.M.; Marin, J.J.G.; Lamarca, A.; Rodrigues, P.M.; Khan, S.A.; Roberts, L.R.; Cardinale, V.; Carpino, G.; Andersen, J.B.; Braconi, C.; et al. Cholangiocarcinoma 2020: The next horizon in mechanisms and management. Nat. Rev. Gastroenterol. Hepatol. 2020, 17, 557–588. [Google Scholar] [CrossRef] [PubMed]
  263. Li, Q.; Wang, J.M.; Liu, C.; Xiao, B.L.; Lu, J.X.; Zou, S.Q. Correlation of aPKC-iota and E-cadherin expression with invasion and prognosis of cholangiocarcinoma. Hepatobiliary Pancreat. Dis. Int. 2008, 7, 70–75. [Google Scholar] [PubMed]
  264. Qian, Y.; Yao, W.; Yang, T.; Yang, Y.; Liu, Y.; Shen, Q.; Zhang, J.; Qi, W.; Wang, J. aPKC-ι/P-Sp1/Snail signaling induces epithelial-mesenchymal transition and immunosuppression in cholangiocarcinoma. Hepatology 2017, 66, 1165–1182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Yang, Y.; Liu, Y.; He, J.C.; Wang, J.M.; Schemmer, P.; Ma, C.Q.; Qian, Y.W.; Yao, W.; Zhang, J.; Qi, W.P.; et al. 14-3-3ζ and aPKC-ι synergistically facilitate epithelial-mesenchymal transition of cholangiocarcinoma via GSK-3β/Snail signaling pathway. Oncotarget 2016, 7, 55191–55210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Okumura, K.; Gogna, S.; Gachabayov, M.; Felsenreich, D.M.; McGuirk, M.; Rojas, A.; Quintero, L.; Seshadri, R.; Gu, K.; Dong, X.D. Gallbladder cancer: Historical treatment and new management options. World J. Gastrointest. Oncol. 2021, 13, 1317–1335. [Google Scholar] [CrossRef] [PubMed]
  267. Tian, L.; Lu, Y.; Yang, T.; Deng, Z.; Xu, L.; Yao, W.; Ma, C.; Li, X.; Zhang, J.; Liu, Y.; et al. aPKCι promotes gallbladder cancer tumorigenesis and gemcitabine resistance by competing with Nrf2 for binding to Keap1. Redox. Biol. 2019, 22, 101149. [Google Scholar] [CrossRef]
  268. Tian, L.; Deng, Z.; Xu, L.; Yang, T.; Yao, W.; Ji, L.; Lu, Y.; Zhang, J.; Liu, Y.; Wang, J. Downregulation of ASPP2 promotes gallbladder cancer metastasis and macrophage recruitment via aPKC-ι/GLI1 pathway. Cell Death Dis. 2018, 9, 1115. [Google Scholar] [CrossRef] [Green Version]
  269. Zhang, G.F.; Wu, J.C.; Wang, H.Y.; Jiang, W.D.; Qiu, L. Overexpression of microRNA-205-5p exerts suppressive effects on stem cell drug resistance in gallbladder cancer by down-regulating PRKCE. Biosci. Rep. 2020, 40, BSR20194509. [Google Scholar] [CrossRef]
  270. Wang, H.; Zhan, M.; Xu, S.W.; Chen, W.; Long, M.M.; Shi, Y.H.; Liu, Q.; Mohan, M.; Wang, J. miR-218-5p restores sensitivity to gemcitabine through PRKCE/MDR1 axis in gallbladder cancer. Cell Death Dis. 2017, 8, e2770. [Google Scholar] [CrossRef] [Green Version]
  271. Koren, R.; Ben Meir, D.; Langzam, L.; Dekel, Y.; Konichezky, M.; Baniel, J.; Livne, P.M.; Gal, R.; Sampson, S.R. Expression of protein kinase C isoenzymes in benign hyperplasia and carcinoma of prostate. Oncol. Rep. 2004, 11, 321–326. [Google Scholar] [CrossRef]
  272. Cornford, P.; Evans, J.; Dodson, A.; Parsons, K.; Woolfenden, A.; Neoptolemos, J.; Foster, C.S. Protein kinase C isoenzyme patterns characteristically modulated in early prostate cancer. Am. J. Pathol. 1999, 154, 137–144. [Google Scholar] [CrossRef] [Green Version]
  273. Villar, J.; Arenas, M.I.; MacCarthy, C.M.; Blánquez, M.J.; Tirado, O.M.; Notario, V. PCPH/ENTPD5 expression enhances the invasiveness of human prostate cancer cells by a protein kinase Cδ-dependent mechanism. Cancer Res. 2007, 67, 10859–10868. [Google Scholar] [CrossRef] [PubMed]
  274. Castilla, C.; Chinchón, D.; Medina, R.; Torrubia, F.J.; Japón, M.A.; Sáez, C. PTPL1 and PKCδ contribute to proapoptotic signalling in prostate cancer cells. Cell Death Dis. 2013, 4, e576. [Google Scholar] [CrossRef] [PubMed]
  275. Kim, J.; Choi, Y.L.; Vallentin, A.; Hunrichs, B.S.; Hellerstein, M.K.; Peehl, D.M.; Mochly-Rosen, D. Centrosomal PKCβII and pericentrin are critical for human prostate cancer growth and angiogenesis. Cancer Res. 2008, 68, 6831–6839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Paone, A.; Starace, D.; Galli, R.; Padula, F.; De Cesaris, P.; Filippini, A.; Ziparo, E.; Riccioli, A. Toll-like receptor 3 triggers apoptosis of human prostate cancer cells through a PKC-α-dependent mechanism. Carcinogenesis 2008, 29, 1334–1342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Zhu, T.; Tsuji, T.; Chen, C. Roles of PKC isoforms in the induction of apoptosis elicited by aberrant Ras. Oncogene 2010, 29, 1050–1061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Villar, J.; Quadri, H.S.; Song, I.; Tomita, Y.; Tirado, O.M.; Notario, V. PCPH/ENTPD5 expression confers to prostate cancer cells resistance against cisplatin-induced apoptosis through protein kinase Cα-mediated Bcl-2 stabilization. Cancer Res. 2009, 69, 102–110. [Google Scholar] [CrossRef] [Green Version]
  279. Truman, J.P.; Rotenberg, S.A.; Kang, J.H.; Lerman, G.; Fuks, Z.; Kolesnick, R.; Marquez, V.E.; Haimovitz-Friedman, A. PKCα activation downregulates ATM and radio-sensitizes androgen-sensitive human prostate cancer cells in vitro and in vivo. Cancer Biol. Ther. 2009, 8, 54–63. [Google Scholar] [CrossRef] [Green Version]
  280. Gurbuz, N.; Park, M.A.; Dent, P.; Abdel Mageed, A.B.; Sikka, S.C.; Baykal, A. Cystine dimethyl ester induces apoptosis through regulation of PKC-δ and PKC-ε in prostate cancer cells. Anticancer Agents Med. Chem. 2015, 15, 217–227. [Google Scholar] [CrossRef]
  281. Von Burstin, V.A.; Xiao, L.; Kazanietz, M.G. Bryostatin 1 inhibits phorbol ester-induced apoptosis in prostate cancer cells by differentially modulating protein kinase C (PKC)δ translocation and preventing PKCδ-mediated release of tumor necrosis factor-α. Mol. Pharmacol. 2010, 78, 325–332. [Google Scholar] [CrossRef] [Green Version]
  282. Wang, H.; Xiao, L.; Kazanietz, M.G. p23/Tmp21 associates with protein kinase Cδ (PKCδ) and modulates its apoptotic function. J. Biol. Chem. 2011, 286, 15821–15831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Yoon, J.S.; Lee, H.J.; Sim, D.Y.; Im, E.; Park, J.E.; Park, W.Y.; Koo, J.I.; Shim, B.S.; Kim, S.H. Moracin D induces apoptosis in prostate cancer cells via activation of PPARγ/PKCδ and inhibition of PKCα. Phytother. Res. 2021, 35, 6944–6953. [Google Scholar] [CrossRef] [PubMed]
  284. Lu, P.H.; Yu, C.C.; Chiang, P.C.; Chen, Y.C.; Ho, Y.F.; Kung, F.L.; Guh, J.H. Paclitaxel induces apoptosis through activation of nuclear protein kinase C-δ and subsequent activation of Golgi associated Cdk1 in human hormone refractory prostate cancer. J. Urol. 2011, 186, 2434–2441. [Google Scholar] [CrossRef] [PubMed]
  285. Benavides, F.; Blando, J.; Perez, C.J.; Garg, R.; Conti, C.J.; DiGiovanni, J.; Kazanietz, M.G. Transgenic overexpression of PKCε in the mouse prostate induces preneoplastic lesions. Cell Cycle. 2011, 10, 268–277. [Google Scholar] [CrossRef] [Green Version]
  286. Garg, R.; Blando, J.M.; Perez, C.J.; Abba, M.C.; Benavides, F.; Kazanietz, M.G. Protein kinase Cε cooperates with PTEN loss for prostate tumorigenesis through the CXCL13-CXCR5 pathway. Cell Rep. 2017, 19, 375–388. [Google Scholar] [CrossRef] [PubMed]
  287. Aziz, M.H.; Hafeez, B.B.; Sand, J.M.; Pierce, D.B.; Aziz, S.W.; Dreckschmidt, N.E.; Verma, A.K. Protein kinase Cε mediates Stat3Ser727 phosphorylation, Stat3-regulated gene expression, and cell invasion in various human cancer cell lines through integration with MAPK cascade (RAF-1, MEK1/2, and ERK1/2). Oncogene 2010, 29, 3100–3109. [Google Scholar] [CrossRef] [Green Version]
  288. Hafeez, B.B.; Zhong, W.; Weichert, J.; Dreckschmidt, N.E.; Jamal, M.S.; Verma, A.K. Genetic ablation of PKCε inhibits prostate cancer development and metastasis in transgenic mouse model of prostate adenocarcinoma. Cancer Res. 2011, 71, 2318–2327. [Google Scholar] [CrossRef] [Green Version]
  289. Yao, S.; Bee, A.; Brewer, D.; Dodson, A.; Beesley, C.; Ke, Y.; Ambroisine, L.; Fisher, G.; Møller, H.; Dickinson, T.; et al. PRKC-ζ expression promotes the aggressive phenotype of human prostate cancer cells and is a novel target for therapeutic intervention. Genes Cancer 2010, 1, 444–464. [Google Scholar] [CrossRef] [Green Version]
  290. Apostolatos, A.H.; Apostolatos, C.A.; Ratnayake, W.S.; Neuger, A.; Sansil, S.; Bourgeois, M.; Acevedo-Duncan, M. Preclinical testing of 5-amino-1-((1R,2S,3S,4R)-2,3-dihydroxy-4-methylcyclopentyl)-1H-imidazole-4-carboxamide: A potent protein kinase C-ι inhibitor as a potential prostate carcinoma therapeutic. Anticancer Drugs 2019, 30, 65–671. [Google Scholar] [CrossRef]
  291. Apostolatos, A.H.; Ratnayake, W.S.; Win-Piazza, H.; Apostolatos, C.A.; Smalley, T.; Kang, L.; Salup, R.; Hill, R.; Acevedo-Duncan, M. Inhibition of atypical protein kinase C-ι effectively reduces the malignancy of prostate cancer cells by downregulating the NF-κB signaling cascade. Int. J. Oncol. 2018, 53, 1836–1846. [Google Scholar] [CrossRef] [Green Version]
  292. Hamshaw, I.; Ajdarirad, M.; Mueller, A. The role of PKC and PKD in CXCL12 directed prostate cancer migration. Biochem. Biophys. Res. Commun. 2019, 519, 86–92. [Google Scholar] [CrossRef] [PubMed]
  293. Ratnayake, W.S.; Apostolatos, C.A.; Breedy, S.; Dennison, C.L.; Hill, R.; Acevedo-Duncan, M. Atypical PKCs activate Vimentin to facilitate prostate cancer cell motility and invasion. Cell Adh. Migr. 2021, 15, 37–57. [Google Scholar] [CrossRef] [PubMed]
  294. Akamatsu, S.; Inoue, T.; Ogawa, O.; Gleave, M.E. Clinical and molecular features of treatment-related neuroendocrine prostate cancer. Int. J. Urol. 2018, 25, 345–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Reina-Campos, M.; Linares, J.F.; Duran, A.; Cordes, T.; L’Hermitte, A.; Badur, M.G.; Bhangoo, M.S.; Thorson, P.K.; Richards, A.; Rooslid, T.; et al. Increased serine and one-carbon pathway metabolism by PKCλ/ι deficiency promotes. Cancer Cell 2019, 35, 385–400. [Google Scholar] [CrossRef] [Green Version]
  296. Hashemi, M.; Shahkar, G.; Simforoosh, N.; Basiri, A.; Ziaee, S.A.; Narouie, B.; Taheri, M. Association of polymorphisms in PRKCI gene and risk of prostate cancer in a sample of Iranian Population. Cell. Mol. Biol. 2015, 61, 16–21. [Google Scholar]
  297. Li, Q.; Gu, C.; Zhu, Y.; Wang, M.; Yang, Y.; Wang, J.; Jin, L.; Zhu, M.L.; Shi, T.Y.; He, J.; et al. Two novel PRKCI polymorphisms and prostate cancer risk in an Eastern Chinese Han population. Mol. Carcinog. 2015, 54, 632–641. [Google Scholar] [CrossRef]
  298. Cairns, P. Renal Cell Carcinoma. Cancer Biomark. 2010, 9, 461–473. [Google Scholar] [CrossRef]
  299. Hsieh, J.J.; Le, V.; Cao, D.; Cheng, E.H.; Creighton, C.J. Genomic classifications of renal cell carcinoma: A critical step towards the future application of personalized kidney cancer care with pan-omics precision. J. Pathol. 2018, 244, 525–537. [Google Scholar] [CrossRef] [Green Version]
  300. Brenner, W.; Färber, G.; Herget, T.; Wiesner, C.; Hengstler, J.G.; Thüroff, J.W. Protein kinase Cη is associated with progression of renal cell carcinoma (RCC). Anticancer Res. 2003, 23, 4001–4006. [Google Scholar]
  301. Pu, Y.S.; Huang, C.Y.; Chen, J.Y.; Kang, W.Y.; Lin, Y.C.; Shiu, Y.S.; Chuang, S.J.; Yu, H.J.; Lai, M.K.; Tsai, Y.C.; et al. Down-regulation of PKCζ in renal cell carcinoma and its clinicopathological implications. J. Biomed. Sci. 2012, 19, 39. [Google Scholar] [CrossRef] [Green Version]
  302. Von Brandenstein, M.; Pandarakalam, J.J.; Kroon, L.; Loeser, H.; Herden, J.; Braun, G.; Wendland, K.; Dienes, H.P.; Engelmann, U.; Fries, J.W. MicroRNA 15a, inversely correlated to PKCα, is a potential marker to differentiate between benign and malignant renal tumors in biopsy and urine samples. Am. J. Pathol. 2012, 180, 1787–1797. [Google Scholar] [CrossRef] [PubMed]
  303. Razorenova, O.V.; Finger, E.C.; Colavitti, R.; Chernikova, S.B.; Boiko, A.D.; Chan, C.K.; Krieg, A.; Bedogni, B.; LaGory, E.; Weissman, I.L.; et al. VHL loss in renal cell carcinoma leads to up-regulation of CUB domain-containing protein 1 to stimulate PKCδ-driven migration. Proc. Natl. Acad. Sci. USA 2011, 108, 1931–1936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Brenner, W.; Greber, I.; Gudejko-Thiel, J.; Beitz, S.; Schneider, E.; Walenta, S.; Peters, K.; Unger, R.; Thüroff, J.W. Migration of renal carcinoma cells is dependent on protein kinase Cδ via β1 integrin and focal adhesion kinase. Int. J. Oncol. 2008, 32, 1125–1131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Brenner, W.; Benzing, F.; Gudejko-Thiel, J.; Fischer, R.; Färber, G.; Hengstler, J.G.; Seliger, B.; Thüroff, J.W. Regulation of β1 integrin expression by PKCε in renal cancer cells. Int. J. Oncol. 2004, 25, 1157–1163. [Google Scholar] [PubMed]
  306. Huang, B.; Cao, K.; Li, X.; Guo, S.; Mao, X.; Wang, Z.; Zhuang, J.; Pan, J.; Mo, C.; Chen, J.; et al. The expression and role of protein kinase C (PKC)ε in clear cell renal cell carcinoma. J. Exp. Clin. Cancer Res. 2011, 30, 88. [Google Scholar] [CrossRef] [Green Version]
  307. Huang, B.; Fu, S.J.; Fan, W.Z.; Wang, Z.H.; Chen, Z.B.; Guo, S.J.; Chen, J.X.; Qiu, S.P. PKCε inhibits isolation and stemness of side population cells via the suppression of ABCB1 transporter and PI3K/Akt, MAPK/ERK signaling in renal cell carcinoma cell line 769P. Cancer Lett. 2016, 376, 148–154. [Google Scholar] [CrossRef]
  308. Owari, T.; Sasaki, T.; Fujii, K.; Fujiwara-Tani, R.; Kishi, S.; Mori, S.; Mori, T.; Goto, K.; Kawahara, I.; Nakai, Y.; et al. Role of nuclear claudin-4 in renal cell carcinoma. Int. J. Mol. Sci. 2020, 21, 8340. [Google Scholar] [CrossRef]
  309. Engers, R.; Mrzyk, S.; Springer, E.; Fabbro, D.; Weissgerber, G.; Gernharz, C.D.; Gabbert, H.E. Protein kinase C in human renal cell carcinomas: Role in invasion and differential isoenzyme expression. Br. J. Cancer 2000, 82, 1063–1069. [Google Scholar] [CrossRef] [Green Version]
  310. Ciążyńska, M.; Kamińska-Winciorek, G.; Lange, D.; Lewandowski, B.; Reich, A.; Sławińska, M.; Pabianek, M.; Szczepaniak, K.; Hankiewicz, A.; Ułańska, M.; et al. The incidence and clinical analysis of non-melanoma skin cancer. Sci. Rep. 2021, 11, 4337. [Google Scholar] [CrossRef]
  311. Krasagakis, K.; Fimmel, S.; Genten, D.; Eberle, J.; Quas, P.; Ziegler, W.; Haller, H.; Orfanos, C.E. Lack of protein kinase C (PKC)-β and low PKC-α, -δ, -ε, and -ζ isozyme levels in proliferating human melanoma cells. Int. J. Oncol. 2002, 20, 865–871. [Google Scholar]
  312. Selzer, E.; Okamoto, I.; Lucas, T.; Kodym, R.; Pehamberger, H.; Jansen, B. Protein kinase C isoforms in normal and transformed cells of the melanocytic lineage. Melanoma Res. 2002, 12, 201–209. [Google Scholar] [CrossRef] [PubMed]
  313. Gilhooly, E.M.; Morse-Gaudio, M.; Bianchi, L.; Reinhart, L.; Rose, D.P.; Connolly, J.M.; Reed, J.A.; Albino, A.P. Loss of expression of protein kinase C β is a common phenomenon in human malignant melanoma: A result of transformation or differentiation? Melanoma Res. 2001, 11, 355–369. [Google Scholar] [CrossRef] [PubMed]
  314. Krasagakis, K.; Tsentelierou, E.; Chlouverakis, G.; Stathopoulos, E.N. Topography of Protein Kinase C βII in Benign and Malignant Melanocytic Lesions. Int. J. Surg. Pathol. 2017, 25, 497–501. [Google Scholar] [CrossRef] [PubMed]
  315. Voris, J.P.; Sitailo, L.A.; Rahn, H.R.; Defnet, A.; Gerds, A.T.; Sprague, R.; Yadav, V.; Caroline Le Poole, I.; Denning, M.F. Functional alterations in protein kinase C beta II expression in melanoma. Pigment. Cell Melanoma Res. 2010, 23, 216–224. [Google Scholar] [CrossRef] [PubMed]
  316. Mahapatra, L.; Andruska, N.; Mao, C.; Gruber, S.B.; Johnson, T.M.; Fullen, D.R.; Raskin, L.; Shapiro, D.J. Protein kinase C-α is upregulated by IMP1 in melanoma and is linked to poor survival. Melanoma Res. 2019, 29, 539–543. [Google Scholar] [CrossRef]
  317. Halder, K.; Banerjee, S.; Ghosh, S.; Bose, A.; Das, S.; Chowdhury, B.P.; Majumdar, S. Mycobacterium indicus pranii (Mw) inhibits invasion by reducing matrix metalloproteinase (MMP-9) via AKT/ERK-1/2 and PKCα signaling: A potential candidate in melanoma cancer therapy. Cancer Biol. Ther. 2017, 18, 850–862. [Google Scholar] [CrossRef] [Green Version]
  318. Putnam, A.J.; Schulz, V.V.; Freiter, E.M.; Bill, H.M.; Miranti, C.K. Src, PKCα, and PKCδ are required for αvβ3 integrin-mediated metastatic melanoma invasion. Cell Commun. Signal. 2009, 7, 10. [Google Scholar] [CrossRef] [Green Version]
  319. Halder, K.; Banerjee, S.; Bose, A.; Majumder, S.; Majumdar, S. Overexpressed PKCδ downregulates the expression of PKCα in B16F10 melanoma: Induction of apoptosis by PKCδ via ceramide generation. PLoS ONE 2014, 9, e91656. [Google Scholar]
  320. Vartanian, A.; Stepanova, E.; Grigorieva, I.; Solomko, E.; Baryshnikov, A.; Lichinitser, M. VEGFR1 and PKCα signaling control melanoma vasculogenic mimicry in a VEGFR2 kinase-independent manner. Melanoma Res. 2011, 21, 91–98. [Google Scholar] [CrossRef]
  321. Mhaidat, N.M.; Thorne, R.F.; Zhang, X.D.; Hersey, P. Regulation of docetaxel-induced apoptosis of human melanoma cells by different isoforms of protein kinase C. Mol. Cancer Res. 2007, 5, 1073–1081. [Google Scholar] [CrossRef] [Green Version]
  322. Heijkants, R.C.; Nieveen, M.; Hart, K.C.; Teunisse, A.F.A.S.; Jochemsen, A.G. Targeting MDMX and PKCδ to improve current uveal melanoma therapeutic strategies. Oncogenesis 2018, 7, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Piperno-Neumann, S.; Larkin, J.; Carvajal, R.D.; Luke, J.J.; Schwartz, G.K.; Hodi, F.S.; Sablin, M.P.; Shoushtari, A.N.; Szpakowski, S.; Chowdhury, N.R.; et al. Genomic profiling of metastatic uveal melanoma and clinical results of a phase I study of the protein kinase C inhibitor AEB071. Mol. Cancer Ther. 2020, 19, 1031–1039. [Google Scholar] [CrossRef] [PubMed]
  324. Ratnayake, W.S.; Apostolatos, C.A.; Apostolatos, A.H.; Schutte, R.J.; Huynh, M.A.; Ostrov, D.A.; Acevedo-Duncan, M. Oncogenic PKC-ι activates vimentin during epithelial-mesenchymal transition in melanoma; a study based on PKC-ι and PKC-ζ specific inhibitors. Cell Adh. Migr. 2018, 12, 447–463. [Google Scholar] [PubMed] [Green Version]
  325. Ratnayake, W.S.; Apostolatos, A.H.; Ostrov, D.A.; Acevedo-Duncan, M. Two novel atypical PKC inhibitors; ACPD and DNDA effectively mitigate cell proliferation and epithelial to mesenchymal transition of metastatic melanoma while inducing apoptosis. Int. J. Oncol. 2017, 51, 1370–1382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Varsano, T.; Lau, E.; Feng, Y.; Garrido, M.; Milan, L.; Heynen-Genel, S.; Hassig, C.A.; Ronai, Z.A. Inhibition of melanoma growth by small molecules that promote the mitochondrial localization of ATF2. Clin. Cancer Res. 2013, 19, 2710–2722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Lau, E.; Feng, Y.; Claps, G.; Fukuda, M.N.; Perlina, A.; Donn, D.; Jilaveanu, L.; Kluger, H.; Freeze, H.H.; Ronai, Z.A. The transcription factor ATF2 promotes melanoma metastasis by suppressing protein fucosylation. Sci. Signal. 2015, 8, ra124. [Google Scholar] [CrossRef] [PubMed]
  328. Pollock, P.M.; Cohen-Solal, K.; Sood, R.; Namkoong, J.; Martino, J.J.; Koganti, A.; Zhu, H.; Robbins, C.; Makalowska, I.; Shin, S.S.; et al. Melanoma mouse model implicates metabotropic glutamate signaling in melanocytic neoplasia. Nat. Genet. 2003, 34, 108–112. [Google Scholar] [CrossRef]
  329. Marín, Y.E.; Namkoong, J.; Cohen-Solal, K.; Shin, S.S.; Martino, J.J.; Oka, M.; Chen, S. Stimulation of oncogenic metabotropic glutamate receptor 1 in melanoma cells activates ERK1/2 via PKCε. Cell. Signal. 2006, 18, 1279–1286. [Google Scholar] [CrossRef]
  330. Zhang, D.; Fu, M.; Li, L.; Ye, H.; Song, Z.; Piao, Y. PKC-δ attenuates the cancer stem cell among squamous cell carcinoma cells through down-regulating p63. Pathol. Res. Pract. 2017, 213, 1119–1124. [Google Scholar] [CrossRef]
  331. Yadav, V.; Yanez, N.C.; Fenton, S.E.; Denning, M.F. Loss of protein kinase C δ gene expression in human squamous cell carcinomas: A laser capture microdissection study. Am. J. Pathol. 2010, 176, 1091–1096. [Google Scholar] [CrossRef] [Green Version]
  332. Singh, A.; Singh, A.; Sand, J.M.; Heninger, E.; Hafeez, B.B.; Verma, A.K. Protein kinase C ε, which is linked to ultraviolet radiation-induced development of squamous cell carcinomas, stimulates rapid turnover of adult hair follicle stem cells. J. Skin Cancer 2013, 2013, 452425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Aziz, M.H.; Manoharan, H.T.; Verma, A.K. Protein kinase C epsilon, which sensitizes skin to sun’s UV radiation-induced cutaneous damage and development of squamous cell carcinomas, associates with Stat3. Cancer Res. 2007, 67, 1385–1394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  334. Sand, J.M.; Bin Hafeez, B.; Aziz, M.H.; Siebers, E.M.; Dreckschmidt, N.E.; Verma, A.K. Ultraviolet radiation and 12-O-tetradecanoylphorbol-13-acetate-induced interaction of mouse epidermal protein kinase Cε with Stat3 involve integration with ERK1/2. Mol. Carcinog. 2012, 51, 291–302. [Google Scholar] [CrossRef] [Green Version]
  335. Chow, R.Y.; Levee, T.M.; Kaur, G.; Cedeno, D.P.; Doan, L.T.; Atwood, S.X. MTOR promotes basal cell carcinoma growth through atypical PKC. Exp. Dermatol. 2021, 30, 358–366. [Google Scholar] [CrossRef] [PubMed]
  336. Mirza, A.N.; Fry, M.A.; Urman, N.M.; Atwood, S.X.; Roffey, J.; Ott, G.R.; Chen, B.; Lee, A.; Brown, A.S.; Aasi, S.Z.; et al. Combined inhibition of atypical PKC and histone deacetylase 1 is cooperative in basal cell carcinoma treatment. JCI Insight 2017, 2, e97071. [Google Scholar] [CrossRef]
  337. Huang, K.; Cui, M.; Ye, F.; Li, Y.; Zhang, D. Global profiling of the signaling network of papillary thyroid carcinoma. Life Sci. 2016, 147, 9–14. [Google Scholar] [CrossRef]
  338. Knauf, J.A.; Ward, L.S.; Nikiforov, Y.E.; Nikiforova, M.; Puxeddu, E.; Medvedovic, M.; Liron, T.; Mochly-Rosen, D.; Fagin, J.A. Isozyme-specific abnormalities of PKC in thyroid cancer: Evidence for post-transcriptional changes in PKC epsilon. J. Clin. Endocrinol. Metab. 2002, 87, 2150–2159. [Google Scholar] [CrossRef]
  339. Afrasiabi, E.; Ahlgren, J.; Bergelin, N.; Törnquist, K. Phorbol 12-myristate 13-acetate inhibits FRO anaplastic human thyroid cancer cell proliferation by inducing cell cycle arrest in G1/S phase: Evidence for an effect mediated by PKCδ. Mol. Cell. Endocrinol. 2008, 292, 26–35. [Google Scholar] [CrossRef] [Green Version]
  340. Molè, D.; Gentilin, E.; Gagliano, T.; Tagliati, F.; Bondanelli, M.; Pelizzo, M.R.; Rossi, M.; Filieri, C.; Pansini, G.; degli Uberti, E.C.; et al. Protein kinase C: A putative new target for the control of human medullary thyroid carcinoma cell proliferation in vitro. Endocrinology 2012, 153, 2088–2098. [Google Scholar] [CrossRef] [Green Version]
  341. Zhu, Y.; Dong, Q.; Tan, B.J.; Lim, W.G.; Zhou, S.; Duan, W. The PKCα-D294G mutant found in pituitary and thyroid tumors fails to transduce extracellular signals. Cancer Res. 2005, 65, 4520–4524. [Google Scholar] [CrossRef] [Green Version]
  342. Prévostel, C.; Martin, A.; Alvaro, V.; Jaffiol, C.; Joubert, D. Protein kinase C α and tumorigenesis of the endocrine gland. Horm. Res. 1997, 47, 140–144. [Google Scholar] [PubMed]
  343. Assert, R.; Kötter, R.; Schiemann, U.; Goretzki, P.; Pfeiffer, A.F. Effects of the putatively oncogenic protein kinase Cα D294G mutation on enzymatic activity and cell growth and its occurrence in human thyroid neoplasias. Horm. Metab. Res. 2002, 34, 311–317. [Google Scholar] [CrossRef] [PubMed]
  344. Motegi, A.; Sakurai, S.; Nakayama, H.; Sano, T.; Oyama, T.; Nakajima, T. PKC theta, a novel immunohistochemical marker for gastrointestinal stromal tumors (GIST), especially useful for identifying KIT-negative tumors. Pathol. Int. 2005, 55, 106–112. [Google Scholar] [CrossRef] [PubMed]
  345. Kang, G.H.; Srivastava, A.; Kim, Y.E.; Park, H.J.; Park, C.K.; Sohn, T.S.; Kim, S.; Kang, D.Y.; Kim, K.M. DOG1 and PKC-θ are useful in the diagnosis of KIT-negative gastrointestinal stromal tumors. Mod. Pathol. 2011, 24, 866–875. [Google Scholar] [CrossRef] [Green Version]
  346. Wang, C.; Jin, M.S.; Zou, Y.B.; Gao, J.N.; Li, X.B.; Peng, F.; Wang, H.Y.; Wu, Z.D.; Wang, Y.P.; Duan, X.M. Diagnostic significance of DOG-1 and PKC-θ expression and c-Kit/PDGFRA mutations in gastrointestinal stromal tumours. Scand. J. Gastroenterol. 2013, 48, 1055–1065. [Google Scholar] [CrossRef]
  347. Kang, J.H.; Asai, D.; Toita, R.; Kitazaki, H.; Katayama, Y. Plasma protein kinase C (PKC)α as a biomarker for the diagnosis of cancers. Carcinogenesis 2009, 30, 1927–1931. [Google Scholar] [CrossRef]
  348. Kang, J.H.; Mori, T.; Kitazaki, H.; Niidome, T.; Takayama, K.; Nakanishi, Y.; Katayama, Y. Serum protein kinase Cα as a diagnostic biomarker of cancers. Cancer Biomark. 2013, 13, 99–103. [Google Scholar] [CrossRef]
  349. Yamada, K.; Oikawa, T.; Kizawa, R.; Motohashi, S.; Yoshida, S.; Kumamoto, T.; Saeki, C.; Nakagawa, C.; Shimoyama, Y.; Aoki, K.; et al. Unconventional secretion of PKCδ exerts tumorigenic function via stimulation of ERK1/2 signaling in liver Cancer. Cancer Res. 2021, 81, 414–425. [Google Scholar] [CrossRef]
  350. Kang, J.H.; Mori, T.; Kitazaki, H.; Niidome, T.; Takayama, K.; Nakanishi, Y.; Katayama, Y. Kinase activity of protein kinase Cα in serum as a diagnostic biomarker of human lung cancer. Anticancer Res. 2013, 33, 485–488. [Google Scholar]
  351. El-Sisi, M.G.; Radwan, S.M.; Saeed, A.M.; El-Mesallamy, H.O. Serum levels of FAK and some of its effectors in adult AML: Correlation with prognostic factors and survival. Mol. Cell. Biochem. 2021, 476, 1949–1963. [Google Scholar] [CrossRef]
  352. Safi, S.; Badshah, Y.; Shabbir, M.; Zahra, K.; Khan, K.; Dilshad, E.; Afsar, T.; Almajwal, A.; Alruwaili, N.W.; Al-Disi, D.; et al. Predicting 3D structure, cross talks, and prognostic significance of KLF9 in cervical cancer. Front. Oncol. 2022, 11, 797007. [Google Scholar] [CrossRef] [PubMed]
  353. Körner, C.; Keklikoglou, I.; Bender, C.; Wörner, A.; Münstermann, E.; Wiemann, S. MicroRNA-31 sensitizes human breast cells to apoptosis by direct targeting of protein kinase C ε (PKCε). J. Biol. Chem. 2013, 288, 8750–8761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Von Brandenstein, M.; Schlosser, M.; Herden, J.; Heidenreich, A.; Störkel, S.; Fries, J.W.U. MicroRNAs as urinary biomarker for oncocytoma. Dis. Markers 2018, 2018, 6979073. [Google Scholar] [CrossRef] [Green Version]
  355. Kawano, T.; Tachibana, Y.; Inokuchi, J.; Kang, J.H.; Murata, M.; Eto, M. Identification of activated protein kinase Cα (PKCα) in the urine of orthotopic bladder cancer xenograft model as a potential biomarker for the diagnosis of bladder cancer. Int. J. Mol. Sci. 2021, 22, 9276. [Google Scholar] [CrossRef]
  356. Köditz, B.; Brandenstein, M.V.; Huerta-Arana, M.; Fries, J.W.U. Novel noninvasive marker of regression of clear cell renal cell carcinoma (ccRCC). Turk. J. Urol. 2022, 48, 49–57. [Google Scholar] [CrossRef] [PubMed]
  357. Davidson, L.A.; Aymond, C.M.; Jiang, Y.H.; Turner, N.D.; Lupton, J.R.; Chapkin, R.S. Non-invasive detection of fecal protein kinase C βII and ζ messenger RNA: Putative biomarkers for colon cancer. Carcinogenesis 1998, 19, 253–257. [Google Scholar] [CrossRef]
Figure 1. The PKC family consists of at least 11 isozymes that are classified into three subfamilies (cPKC, nPKC, and aPKC). The activation of PKC isozymes is positively associated with poor survival rate, anticancer drug resistance, or increased recurrence in patients with cancer. Furthermore, higher levels of PKC isozymes are found in tissues or body fluids of patients with cancer compared to those in healthy individuals. These data suggest that PKC isozymes represent useful therapeutic targets and potential diagnostic and prognostic biomarkers for cancer.
Figure 1. The PKC family consists of at least 11 isozymes that are classified into three subfamilies (cPKC, nPKC, and aPKC). The activation of PKC isozymes is positively associated with poor survival rate, anticancer drug resistance, or increased recurrence in patients with cancer. Furthermore, higher levels of PKC isozymes are found in tissues or body fluids of patients with cancer compared to those in healthy individuals. These data suggest that PKC isozymes represent useful therapeutic targets and potential diagnostic and prognostic biomarkers for cancer.
Cancers 14 05425 g001
Table 1. PKC isozymes as diagnostic and prognostic biomarkers and therapeutic targets in various cancer types.
Table 1. PKC isozymes as diagnostic and prognostic biomarkers and therapeutic targets in various cancer types.
Cancer TypesPKC IsozymesActivityEffect of Change in PKC Activation on the CancerRefs.
Bladder cancerPKCαUpregulationPoor prognosis
Increased anticancer drug resistance
[14,15]
[16,18]
Blood and bone marrow cancer
Multiple myelomaPKCβUpregulationPotential therapeutic target[24]
Leukemia: lymphocytic leukemiaPKCαUpregulationEnhanced chemoresistance[27,28]
PKCβUpregulationPotential therapeutic target[30]
Leukemia: myeloid leukemiaPKCαUpregulationPoor survival
Promoted anticancer drug resistance
[40]
[41,51]
PKCβUpregulationEnhanced anticancer drug resistance[52]
PKCδUpregulationIncreased anticancer drug-mediated apoptosis[45,47]
PKCεUpregulationPoor survival and increased anticancer drug resistance[43,44]
Myelodysplastic syndromesPKCαUpregulation (1)Induced erythropoiesis[56]
LymphomaPKCβIIUpregulationPoor prognostic marker and chemotherapeutic target[63,66]
PKCδUpregulationIncreased anticancer drug-mediated apoptosis[68,69]
Brain cancer (glioblastoma)PKCαUpregulationPotential therapeutic target
Potential prognostic marker
[73,79]
[72]
PKCδUpregulationAntiproliferative and proapoptotic[90,91]
PKCεUpregulationPotential therapeutic target[93]
PKCιUpregulationPotential therapeutic target[83,84]
Breast cancerPKCαUpregulationPoor survival and prognosis
Maintenance of migratory and invasive behavior
Decreased ER levels and increased antiestrogen resistance
Enhanced anti-ErbB-1 sensitivity in ErbB-2-positive breast cancer
[100]
[101]
[102,103]
[111]
PKCδUpregulationEnhanced mammary tumorigenesis[114,115]
PKCθUpregulationIncreased migratory and invasive behavior[122,123]
PKCεUpregulationDecreased disease-free survival[131]
PKCηUpregulationEnhanced breast cancer malignancy
Poor survival following anticancer treatment
[125]
[126]
PKCζUpregulationIncreased invasive behavior
Poor prognosis, disease-free survival, and survival rate
[121]
[120]
PKCλUpregulationPoor prognosis[130]
Colorectal (colon) cancerPKCα (2)Downregulation
Upregulation
Potential therapeutic target
Enhanced anticancer drug resistance
[144]
[135]
PKCδUpregulation (1)Increased cancer progression and poor prognosis[156,158]
PKCζUpregulationPotential therapeutic target[145,146]
PKCιUpregulationPotential therapeutic target[159]
Gastric (stomach) cancerPKCαUpregulationPoor prognosis and increased anticancer drug resistance[162,164]
PKCιUpregulationEnhanced recurrence of cancer and poor survival[166,169]
Head and neck squamous cell carcinomaPKCαUpregulationPoor prognosis and survival[174,175]
PKCβIIUpregulation (1)Poor survival and rapid recurrence[172]
PKCθUpregulation (1)Poor survival and rapid recurrence[179]
PKCιUpregulationIncreased malignancy and poor survival[177]
Liver cancer (hepatocellular carcinoma)PKCαUpregulationPoor prognosis and survival
Immune escape and anti-PD1 tolerance
[181,184]
[185]
PKCβUpregulationPotential tumor suppressor[194]
PKCδUpregulationPotential prognostic marker
Poor disease-free survival
[186,189]
[190]
PKCλ/ιUpregulationPotential tumor suppressor[191]
PKCηDownregulationPoor long-term survival[196]
Lung cancerPKCαUpregulationPotential therapeutic target and poor survival[200]
PKCδUpregulationIncreased cell survival
Increased anticancer drug resistance and potential therapeutic target
[208]
[213]
PKCεUpregulationPotential therapeutic target
Elevated survival and anticancer drug resistance
[215]
[215,217]
PKCηUpregulationPoor prognosis and survival[219]
PKCιUpregulationPoor prognosis[218]
Ovarian cancerPKCαUpregulationPoor prognosis and survival
Increased anticancer drug resistance
[226]
[222,223]
PKCιUpregulationPoor prognosis and survival
Potential therapeutic target
[227,228]
[230]
PKCζUpregulationPoor prognosis[233,234]
Pancreatic, bile duct, and gallbladder cancer
Pancreatic cancerPKCαUpregulationPotential therapeutic target
Potential prognostic marker
[241,244]
[236]
PKCδUpregulationEnhanced cancer progression and poor survival
Potential therapeutic target
[254,255]
[259,260]
PKCθUpregulationPoor survival and therapeutic target[254,261]
PKCιUpregulationPotential prognostic marker and therapeutic target[246,249]
PKCζUpregulationEnhanced worse prognosis[251,252]
Bile duct cancerPKCιUpregulationPotential prognostic marker and therapeutic target[263,264]
Gallbladder cancerPKCιUpregulationPoor prognosis
Enhanced cell growth, migration, and anticancer drug resistance
[267]
[268,269]
PKCεUpregulationEnhanced anticancer drug resistance, proliferation, and colony formation rate[269,270]
Prostate cancerPKCαUpregulationPromoted cell growth and anticancer drug resistance[275,276]
PKCδUpregulationEnhanced anticancer drug-induced cell apoptosis[280,284]
PKCεUpregulationPotential therapeutic target[288]
PKCζUpregulationWorse survival and poor overall survival
Potential preventive and therapeutic target
[289]
[290,292]
PKCιUpregulationPotential preventive and therapeutic target[290,291]
Renal cell carcinomaPKCδUpregulationPromoted cancer cell migration[303]
PKCεUpregulationPotential therapeutic target[307]
Skin cancer
MelanomaPKCαUpregulationPoor prognosis and survival
Potential therapeutic target for pancreatic cancer stem cells
[316]
[317]
PKCδUpregulationEnhanced proapoptotic response[319,321]
PKCζ and ιUpregulationPotential therapeutic target[324]
PKCεUpregulationPotential therapeutic target[326,327]
Non-melanomaPKCδUpregulationProtective role in squamous cell carcinomas[330,331]
PKCεUpregulationEnhanced development of squamous cell carcinomas[332,334]
Thyroid carcinomaPKCαMutationLoss of function[341,343]
(1) Increased nuclear translocation or expression. (2) Note that there are two different reports, the upregulation or downregulation of PKCα in colorectal cancer.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kawano, T.; Inokuchi, J.; Eto, M.; Murata, M.; Kang, J.-H. Protein Kinase C (PKC) Isozymes as Diagnostic and Prognostic Biomarkers and Therapeutic Targets for Cancer. Cancers 2022, 14, 5425. https://doi.org/10.3390/cancers14215425

AMA Style

Kawano T, Inokuchi J, Eto M, Murata M, Kang J-H. Protein Kinase C (PKC) Isozymes as Diagnostic and Prognostic Biomarkers and Therapeutic Targets for Cancer. Cancers. 2022; 14(21):5425. https://doi.org/10.3390/cancers14215425

Chicago/Turabian Style

Kawano, Takahito, Junichi Inokuchi, Masatoshi Eto, Masaharu Murata, and Jeong-Hun Kang. 2022. "Protein Kinase C (PKC) Isozymes as Diagnostic and Prognostic Biomarkers and Therapeutic Targets for Cancer" Cancers 14, no. 21: 5425. https://doi.org/10.3390/cancers14215425

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