Next Article in Journal
ER Stress-Perturbed Intracellular Protein O-GlcNAcylation Aggravates Podocyte Injury in Diabetes Nephropathy
Next Article in Special Issue
Classification of MLH1 Missense VUS Using Protein Structure-Based Deep Learning-Ramachandran Plot-Molecular Dynamics Simulations Method
Previous Article in Journal
Immune Profiling of Patients with Systemic Sclerosis through Targeted Proteomic Analysis
Previous Article in Special Issue
An Update on Protein Kinases as Therapeutic Targets—Part II: Peptides as Allosteric Protein Kinase C Modulators Targeting Protein–Protein Interactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Update on Protein Kinases as Therapeutic Targets—Part I: Protein Kinase C Activation and Its Role in Cancer and Cardiovascular Diseases

1
The Azrieli Faculty of Medicine in the Galilee, Bar-Ilan University, Henrietta Szold St. 8, Safed 1311502, Israel
2
Department of Medicine, School of Medicine, Stanford University, 300 Pasteur Drive, Stanford, CA 94305, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(24), 17600; https://doi.org/10.3390/ijms242417600
Submission received: 1 November 2023 / Revised: 10 December 2023 / Accepted: 12 December 2023 / Published: 18 December 2023

Abstract

:
Protein kinases are one of the most significant drug targets in the human proteome, historically harnessed for the treatment of cancer, cardiovascular disease, and a growing number of other conditions, including autoimmune and inflammatory processes. Since the approval of the first kinase inhibitors in the late 1990s and early 2000s, the field has grown exponentially, comprising 98 approved therapeutics to date, 37 of which were approved between 2016 and 2021. While many of these small-molecule protein kinase inhibitors that interact orthosterically with the protein kinase ATP binding pocket have been massively successful for oncological indications, their poor selectively for protein kinase isozymes have limited them due to toxicities in their application to other disease spaces. Thus, recent attention has turned to the use of alternative allosteric binding mechanisms and improved drug platforms such as modified peptides to design protein kinase modulators with enhanced selectivity and other pharmacological properties. Herein we review the role of different protein kinase C (PKC) isoforms in cancer and cardiovascular disease, with particular attention to PKC-family inhibitors. We discuss translational examples and carefully consider the advantages and limitations of each compound (Part I). We also discuss the recent advances in the field of protein kinase modulators, leverage molecular docking to model inhibitor–kinase interactions, and propose mechanisms of action that will aid in the design of next-generation protein kinase modulators (Part II).

1. Introduction

Protein kinases are a large and diverse family of more than 500 proteins encoded by ~2% of the human genome, classified into eight superfamilies (AGC, aPK, CAMK, CMGC, STE, TK, TKL, and CK1) according to homology within their catalytic domains. Kinases catalyze phosphorylation, the chemical addition of a phosphoryl group (PO3) to the target substrate by the hydrolysis of a γ-phosphate group from adenosine triphosphate (ATP), a reversible and the most widespread type of post-translational modification (PTM) used in signal transduction. In eukaryotic organisms, protein phosphorylation primarily occurs on serine, threonine, and tyrosine residues. Although sugars and lipids can also be phosphorylated, here we are focusing on protein phosphorylation. Approximately one-third, and up to two-thirds, of the proteins in a cell may be phosphorylated at one time or another, affecting a very large set of cellular pathways by turning activities “on” or “off” [1]. Phosphorylation plays major roles in numerous cellular functions, including transcription, translation, metabolism, proliferation, division, cell-cycle progression, biosynthesis, movement, and survival. These processes are critical to cellular homeostasis, and dysregulated kinase activity has been linked to a variety of pathological conditions, such as neurodegeneration [2], inflammation [3], autoimmunity [3], cancer [4,5], and cardiovascular diseases (CVDs) [6]. Furthermore, protein phosphorylation is significantly underlined by the Nobel Prize in Physiology or Medicine (1992) awarded to Edmond H. Fischer and Edwin G. Krebs “for their discoveries concerning reversible protein phosphorylation as a biological regulatory mechanism”. Their major contribution was the discovery that the conversion of the inactive enzyme phosphorylase b to active phosphorylase a is caused by phosphorylation, which is regulated by the protein kinase called phosphorylase kinase [7,8]. Given this importance, it is not surprising that protein kinases are the second most therapeutically targeted group of proteins, after the G-protein-coupled receptors (GPCRs) [9,10,11,12], and the pharmaceutical industry has dedicated approximately one-third of new drug development programs over the last decade to the development of protein kinase modulators [13,14].
Imatinib (i.e., STI571, or Gleevec), which received Food and Drug Administration (FDA) approval in 2001, is considered the first protein kinase inhibitor that was clinically approved [15]. Although fasudil (an inhibitor of Rho-dependent kinases) was approved in 1995, and rapamycin (i.e., sirolimus, an inhibitor of the protein kinase TORC1) was approved in 1999, both were approved without knowledge of the identity of their target proteins. Imatinib is an oral treatment for chronic myelogenous leukemia (CML), an uncommon type of cancer of the bone marrow, which arises from a pluripotent hematopoietic stem cell that has undergone a reciprocal translocation between the breakpoint cluster region protein (Bcr) gene on chromosome 22 and the Abelson tyrosine-protein kinase 1 (Abl1) proto-oncogene on chromosome 9. The fusion Bcr–Abl protein results in a shortened chromosome 22 (i.e., the Philadelphia chromosome) that lacks the autoinhibitory domains of the Abl1 kinase and, therefore, results in the constitutive activation of Abl1 proliferative function. Imatinib was developed in the late 1990s by the biochemist Nicholas Lyndon, and its use to treat CML was driven by Brian Druker. The drug is a potent small-molecule kinase inhibitor (SMKIs) that functions as a competitive inhibitor of the ATP binding site that interacts with the Abl1 domain via six hydrogen bond interactions, leading to a stabilized imatinib–Bcr–Abl complex; this prevents ATP from reaching its binding site, resulting in inhibition of the Bcr–Abl tyrosine kinase [16,17]. Imatinib has revolutionized drug therapy for CML, and the drug was featured on the front cover of Time Magazine (28 May 2001, Vol. 157 No. 21) and termed a “magic bullet”. Lyndon, Druker, and their colleagues were granted the Lasker–DeBakey Clinical Medical Research Award in 2009 for “converting fatal cancer into a manageable condition”. Imatinib also became one of the world’s most commercially successful drugs with peak sales of USD 4.6 billion in 2012. The spectacular efficacy of Imatinib was observed as soon as it entered clinical trials in 1998. A large-scale phase III trial, the international randomized study of interferon and STI571 (IRIS), compared a single daily dose of imatinib to interferon alpha (IFNα)/cytarabine treatment in newly diagnosed CML patients and showed the superior efficacy of imatinib over IFNα/cytarabine (76% versus 15%, respectively) [18], a five-year survival rate of 85% [19], and enormous commercial success. This changed the perception of protein kinases as drug targets, which had previously received skepticism from many pharmaceutical companies.
Since imatinib, a plethora of targeted therapeutic agents have been developed and tested in humans, and many of these trials, predominantly those targeting mitogen-activated protein kinase (MAPK) signaling, have demonstrated significant improvement versus the standard of care and resulted in FDA approvals [20,21]. Currently, there are 98 approved kinase inhibitors worldwide, 71 of which are SMKIs that have been approved by the FDA, targeting 21 kinase families constituting approximately 20% of the kinome. Interestingly, the number of SMKIs approved by the FDA has more than doubled since 2016, with 37 new approvals, making SMKIs approximately 15% of all novel drug approvals in the last 5 years (2016–2021). In addition, 16 more SMKIs have been granted approval by other regulatory agencies [22].
Most approved SMKIs are indicated for oncologic conditions. The lack of SMKI selectivity is the main reason why there has been less success in producing kinase-inhibitor drugs for other conditions [23]. Nevertheless, SMKIs have been approved for several other diseases. For example, Janus kinase (JAK) inhibitors (e.g., ruxolitinib, tofacitinib, baricitinib, upadacitinib, and fedratinib) have been approved for the treatment of autoimmune diseases, inflammatory disorders, and dermatoses [24]. Rho-associated protein kinase (ROCKs) inhibitors (e.g., netarsudil and ripasudil) have also been approved for various conditions related to primary open-angle glaucoma (POAG) and other etiologies of increased intraocular pressure [25,26]. The number of SMKIs that are being approved for the treatment of diseases in nononcologic therapeutic fields continues to rise, and, currently, one-third of the SMKIs in clinical development aim to address disorders beyond oncology [13,24,27,28].
The pharmacological modulation of kinase activity is almost exclusively achieved through SMKIs that target the conserved kinase domain. Attaining the selectivity required for SMKIs to become successful drugs remains one of the largest hurdles in an early-stage kinase-inhibitor project. Certain disease processes (e.g., oncology) that are a result of multiple aberrant pathway activation may still benefit from the use of multikinase inhibitors that target several pathways at the same time (e.g., lenvatinib and sorafenib multikinase inhibitors approved for the treatment of progressive radioiodine-refractory differentiated thyroid cancer). Yet, such compounds still need to be selective inhibitors of a well-defined set of kinase enzymes, and the disadvantage of this approach may be the increase in adverse effects that comes with inhibiting multiple signal-transduction pathways, which can also be critical to physiological functions. Therefore, the lack of selectivity of multikinase inhibitors causes, in many cases, toxicities that result in the discontinuation of promising drug candidates [29,30,31]. Currently, the majority of the FDA-approved kinase inhibitors are SMKIs targeting the kinase ATP-binding site (63 SMKIs) [32]. However, in many cases, these SMKIs demonstrate low specificity towards the target kinase, resulting in problems with toxicity and a variety of side effects. Therefore, the pharmaceutical industry continues to invest in emerging trends focused on identifying alternative approaches to specifically target protein kinases more selectively.
One rapidly emerging technology leverages proteolysis-targeting chimera (PROTAC) molecules, which harness the ubiquitin–proteasome system to degrade a target protein and was first proposed by the pioneers Crews and coworkers in 2001 [33] (for recent reviews [34,35,36]). Since its first report, the technology has expanded from academia to industry. To date, PROTACs are being validated by several programs in preclinical and early clinical development, and the first oral PROTACs (ARV-110 and ARV-471) have shown encouraging results in clinical trials for prostate and breast cancer treatment [37]. Another exciting approach involves the use of antibodies, which can provide superior potency, selectivity, and pharmacokinetic properties compared to small molecules [38]. Indeed, several antibodies have been approved, such as cetuximab in colorectal and head and neck cancer, and trastuzumab in breast cancer [39,40,41]. However, both PROTACs and antibodies also demonstrated some inherent challenges. PROTACs are more difficult and more expensive to develop than SMKIs, they do not readily degrade many types of proteins, their toxicology is uncertain, their adverse effects are largely unknown, and their oral bioavailability is low [37,42,43]. Antibodies demonstrate several challenges as well, including production cost, tissue accessibility, and immunogenicity [41,44,45,46]. Since the majority of SMKIs target the ATP site that is structurally similar, many of them suffer from a lack of selectivity for structurally related kinase families, resulting in off-target toxicity that can cause dangerous side effects and is a major cause of clinical trial failure [14,47,48].
Protein kinase C (PKC) is a prototypical class of serine/threonine kinases within the AGC-kinase superfamily named after protein kinase A, cyclic GMP-regulated kinases, and PKC. The AGC-kinase superfamily consists of 63 evolutionarily related serine/threonine protein kinases that are divided into 14 families and 21 subfamilies. PKC was originally identified by Nishizuka and colleagues over 40 years ago as a phospholipid and calcium-dependent protein kinase that phosphorylated histone and protamine. Originally, the enzyme was called protein kinase M (PKM) since only magnesium was required for activity. In subsequent studies, it was discovered that PKM was the proteolytic product of an enzyme whose activity was stimulated by Ca2+. This enzyme was named PKC because it was activated by second-messenger Ca2+ [49,50]. PKCs play a vital role in signal transduction by phosphorylation of target proteins to control numerous cellular functions. The PKC family of mammals consists of nine different genes and includes a variety of members due to alternative splicing [51,52]. Based on their structural components and activation mechanisms, PKCs are subdivided into three groups: classical PKC, novel PKC, and atypical PKC. The complexity in PKC signaling arises from the fact that PKC is a superfamily of structurally related kinases with diverse biological functions and, in some cases, opposing roles in cell proliferation, differentiation, apoptosis, and angiogenesis [53,54]. Precise control of PKC signal amplitude is necessary for the healthy condition of a cell and an organism, and the alteration of normal PKC activity is linked to numerous human diseases, including metabolic disorders [55,56], neurodegenerative disease [2], autoimmune disease [57,58,59,60], cancer [61], and CVDs [62,63,64]. Therefore, the development of selective PKC modulators is predicted to represent a breakthrough in drug discovery and has become the subject of intense studies by academic laboratories and pharmaceutical companies [65,66,67,68]. Unraveling the enormous complexity of the mechanisms by which PKC isozymes impact human diseases is key for assessing their potential as therapeutic targets.
A selection of reviews on the protein kinase family and on protein kinase modulators have been published recently. Several reviews from a basic research perspective featured critical advances in understanding kinase signal-transduction pathways [69,70,71,72]. Other reviews focused on the advantages of targeting protein kinases in treating various human diseases [14,73,74,75,76], such as immune disorders [77,78], cancer [71,79], CVDs [6], and other associated conditions [24,80]. Finally, several reviews focused on the FDA and other regulatory approvals of kinase inhibitors [76,81] and targets in clinical trials [13]. To the best of our knowledge, there is not an up-to-date review published on kinase allosteric modulators nor the role of peptides in therapeutic applications targeting PKC. In these two reviews, we first provide an updated summary of the current literature about PKC’s role in human diseases, emphasizing its critical role in cancer and CVDs (Part I). Next, we aim to provide insight into alternative therapeutic approaches developed to target PKC and highlight the opportunities and challenges in applying these advances to cancer and CVDs. Furthermore, using available structures, the AlphaFold protein-structure database [82,83], and computational methods, we leverage molecular docking to model the interaction of previously developed modulators and explore their mechanisms of action. We discuss examples of successful applications and insights into the strengths and limitations of each modulator, thus providing novel tools and guidance to new and established researchers in the field (Part II).

2. The Protein Kinase C (PKC) Family

PKC is a ubiquitous enzyme found in almost all human cell types including the endothelium, vascular smooth muscle, and fibroblasts. The PKC family comprises a group of highly related protein kinases that phosphorylate serine and threonine residues on a large number of proteins and therefore regulate diverse cellular responses. PKC mediates significant and sometimes opposing effects in different tissues and is widely implicated across physiological and pathological processes [84]. The members that populate the mammalian PKC family evolved from the single PKC in saccharomyces cerevisiae, PKC1 [85,86]. PKC isozymes have been classified into three groups, ‘conventional’ or ‘classical’ PKCs (cPKCs) that are composed of PKCα, two splice variants of PKCβ (PKCβI and PKCβII), and PKCγ, which all require phosphatidylserine (PS), diacylglycerol (DAG), and Ca2+ to be activated; ‘novel’ PKCs (nPKCs), a group that includes PKCδ, PKCε, PKCη, and PKCθ, which require only PS and DAG to be activated; and ‘atypical’ PKCs (aPKCs), including PKCζ and PKCι/λ, which require only PS.
All family members share a common architecture of an amino-terminal regulatory moiety (approximately 35 kDa) linked by a flexible hinge segment to a carboxyl-terminal kinase domain (approximately 45 kDa). PKC comprises a single polypeptide structure consisting of a regulatory domain and a catalytic (kinase) domain separated by a hinge region (V3) with four functional conserved regions (C1–C4), which are interspersed by five variable regions (V1–V5). The regulatory domain comprises C1 and C2 regions that interact with DAG, PS, and Ca2+, thereby acting as the membrane-targeting module. The catalytic domain comprises the C3 and C4 domains that constitute the ATP- and substrate-binding lobes of the kinase core. Conventional isozymes contain tandem C1 domains (C1A and C1B) that bind to DAG or phorbol esters found in cell membranes, as well as a C2 domain that binds membranes in the presence of the second-messenger Ca2+. Novel PKC isoenzymes similarly possess two tandem C1 domains that bind to DAG or phorbol esters. Interestingly, the C1B domain of novel PKCs has a 100-fold higher affinity for DAG compared with the C1B domain of classical PKCs. Novel PKCs also contain a variant of the C2 domain (novel C2 domain), which lacks key residues that interact with Ca2+. To compensate for the lack of C2 domain contribution to membrane recruitment, novel PKC activation only requires DAG [87,88]. Atypical PKC isozymes contain a single atypical C1 domain that retains the ability to bind anionic phospholipids, as well as a Phox and Bem1 (PB1) domain that mediates protein–protein interactions (PPIs). While the PKC isozyme catalytic domains are very similar, they differ from each other in the C2 region of the regulatory domain as well as in the variable regions. The regulatory domain of all PKC isoenzymes contains a short autoinhibitory pseudo-substrate sequence whose occupation of the kinase substrate-binding cavity maintains these kinases in an inactive state [54,89]. While both catalytic and regulatory domains represent target opportunities for the development of novel therapeutics to modulate PKC activity, each domain also presents its own challenges for achieving selective compounds (Figure 1).
PKC activation relies on three molecular mechanisms: (i) phosphorylation, (ii) cofactor binding, and (iii) intracellular translocation. Like many other kinases, PKC is regulated by phosphorylation, which can alter enzyme activity, regulate protein stability, affect protein interactions or localization, and influence other PTMs. For PKC to become activated, the protein must first undergo phosphorylation on several critical residues, which occurs shortly after biosynthesis via a process called PKC priming. Newly synthesized isozymes are in an open and degradation-sensitive confirmation until they undergo a sequence of stabilizing phosphorylations on several conserved residues, resulting in a stable, autoinhibited enzyme that is primed to respond to second messengers. Conventional and novel PKCs are constitutively phosphorylated at three conserved domains: the activation loop, the turn motif, and the hydrophobic motif. Atypical PKCs are phosphorylated at the activation loop and turn motif but contain a phosphomimetic glutamic acid at the hydrophobic motif. Additionally, several PKC isozymes require ERK activation to exert important downstream effects.
Mass spectrometric studies of the conventional PKCβII isozyme identified three constitutively phosphorylated sites located in the PKC kinase domain (Thr500 at the activation loop, Ser641 at the turn motif, and Ser660 at the hydrophobic motif); interestingly, these phosphorylation sites are highly conserved among most AGC kinases, including protein kinase B (PKB, or Akt), in addition to PKC [90,91]. In order for PKC to undergo priming phosphorylation, conventional and novel PKCs bind to the chaperone heat-shock protein 90 (Hsp90) and its cochaperone cell-division cycle 37 (Cdc37), which promotes the maturation of these kinases by acting as an adapter and loading them onto the Hsp90 complex; the complex then binds open PKC via a conserved PXXP motif in the C-terminal tail that harbors a conserved phosphorylation site termed the hydrophobic motif [92]. Next, phosphoinositide-dependent kinase-1 (PDK-1, a key regulator of PKCs) phosphorylates amino acids in the activation loop (e.g., Thr500 in PKCβII). Phosphorylation by PDK-1 is likely the first phosphorylation event in the processing of PKC by phosphorylation. This phosphorylation makes a conformational change that triggers two rapid subsequent phosphorylation events at the carboxy terminus. First occurs phosphorylation of the turn motif that, in many cases, is regulated by the mammalian target of rapamycin complex 2 (mTORC2) [93,94], followed by phosphorylation at the hydrophobic motif (e.g., Thr641 and Ser660 in PKCβII) [95]. The phosphorylation of the turn motif stabilizes the structure of mature PKC by anchoring the carboxyl-terminal tail on the upper lobe of the kinase to stabilize the active conformation of the kinase. Finally, PKC autophosphorylates by an intramolecular reaction at the hydrophobic motif, which is regulated by the interaction of Hsp90 with the PXXP clamp. The result is a stable PKC that is released into the cytosol in a catalytically competent closed conformation (i.e., autoinhibited state) in which the pseudo-substrate region masks the substrate-binding cavity.
Each phosphorylation event induces conformational changes in the PKC molecule that result in altered thermal stability, resistance to phosphatases, and catalytic competency. Phosphorylation is vital to render PKC in a catalytically competent conformation and to protect PKC from degradation. Mechanisms that prevent the phosphorylation of PKC, such as the loss of PDK-1, inhibition of mTORC2, or impairment of PKC’s intrinsic catalytic activity, prevent the maturation of PKC, resulting in PKCs that are highly susceptible to degradation. Therefore, these steps play a critical role in controlling cellular levels of PKC [93,94,96,97]. Since PKC phosphorylation is constitutive, its activity is regulated by the cellular levels of PKC, which are directly regulated by its phosphorylation. Understanding how to modulate these levels has important therapeutic implications, as altered PKC levels correlate with various human diseases [98,99,100,101]. In addition to the described phosphorylation sequence, PKC has an abundance of other PTMs, such as additional phosphorylation events, acetylation, and ubiquitination [100,102,103]. Whether any of these or other yet-to-be-identified sites are critical for PKC “priming” and maturing remains to be established (Figure 2).

2.1. Regulation by Lipid Second Messengers

The fully phosphorylated “mature” PKC is localized to the cytosol in an autoinhibited and stable conformation that is poised to respond to second messengers. Lipid second messengers are signaling molecules produced in response to extracellular stimuli and coordinate signaling by recruiting and activating kinases. Although after phosphorylation PKC is catalytically competent, an autoinhibitory pseudo-substrate that binds the substrate-binding cavity maintains PKC inactive until the appropriate second messengers bind. The activity of mature conventional and novel PKC isozymes is regulated by binding to second messengers, such as DAG and Ca2+. Yet, the differential binding of second messengers between conventional PKCs (DAG and Ca2+) and novel PKCs (solely DAG, as they are not responsive to Ca2+) leads to substantial differences in their spatiotemporal signaling dynamics.
Two prominent lipid second-messenger pathways are those mediated by DAG and phosphatidylinositol (PI), a minor phospholipid with a characteristic fatty acid profile. Initially, extracellular signals cause the hydrolysis of phosphatidylinositol-4,5-bisphosphate (PIP2), DAG is generated, and Ca2+ is released from intracellular stores. Next, the binding of Ca2+ to the C2 domain recruits conventional PKC isozymes primarily to the plasma membrane. Ca2+ binding induces a conformational change in which the C2 domain is displaced from the kinase domain, exposing a PIP2-binding basic face that is masked in the autoinhibited conformation, resulting in the localization of PKC to the plasma membrane. The C1B domain of inactive PKC also binds to its membrane-embedded ligand, DAG. The coordinated engagement of both the C1 and C2 domains on the membrane provides the energy, prompting a second conformational change that expels the autoinhibitory pseudo-substrate from the substrate-binding cavity and reveals an active open confirmation, allowing substrate phosphorylation. Novel PKC isozymes are activated by DAG alone, allowing them to be activated by phospholipase C-catalyzed hydrolysis of lipids other than PIP2. Since novel isozymes do not have a functional C2 domain but have a C1A domain that has a 100-fold higher affinity for DAG, they are translocated mainly to DAG-rich endomembranes such as the Golgi. Atypical PKCs are regulated by neither DAG nor calcium, but they are regulated by the PPIs of their PB1 domain to the PB1 domains of protein scaffolds, such as p62 and Par6. After binding to these proteins, atypical PKC isozymes are harbored near the plasma membrane in proximity to their substrates. Furthermore, in many cases, this interaction moves the pseudo-substrate domain away from the substrate-binding cavity, thus relieving autoinhibitory constraints [104].
Conformational changes in all PKC isoforms allow the proteins to interact with specific anchoring and/or scaffold proteins, which determine intracellular translocation and functional activity. Tsien et al. pioneered technologies for the development of fluorescence resonance energy transfer (FRET) reporters to measure the spatiotemporal dynamics of PKC signaling in live cells [105], which revealed that Ca2+ oscillations, with or without DAG oscillations, drive oscillations of conventional PKC activity [106,107]. Live-cell imaging studies demonstrated that there are different “signatures” of PKC activity at defined sublocations in cells [108]. For example, rapid rises in intracellular Ca2+ drive rapid activation of conventional PKC isozymes at the plasma membrane [109], while relatively high levels of DAG (e.g., in the Golgi membranes) mediates a “basal” interaction of novel PKC isozymes at the Golgi [110].
Interestingly, the PKC enzyme superfamily exhibits an especially high degree of complexity and variability, as specific conventional and novel PKCs can also be activated independently of second messengers. Certain PKCs can be activated by the accumulation of reactive oxygen species (ROS), which are elevated in various human disease processes, including neurodegeneration, cancer, and CVDs [111,112]. For example, PKCδ is modified at multiple tyrosine residues by Src-dependent phosphorylation in the regulatory moiety in response to cell stimulation with H2O2, inducing DAG-independent activation and effects on localization [113,114,115]. H2O2-induced Tyr311 phosphorylation has also been proposed to activate PKCδ by inducing caspase-3 cleavage between its regulatory and catalytic domains, resulting in a nuclear-localized uninhibited catalytic domain [116]. In addition, the PKCδ nuclear localization sequence (NLS) binds importin-α, which facilitates PKCδ transport through the nuclear pore complex [115]. Thus, in addition to reversible activation by binding second messengers, alternative mechanisms of activation fine-tune the localization, function, and activity of individual PKC isozymes, depending on their microenvironments.

2.2. Regulation by Scaffold Interactions

A critical mechanism to control protein kinase specificity relies on scaffold and/or anchored proteins. As multiple PKC isozymes are expressed in the same cell and respond to the same stimuli (e.g., the same second messenger), specificity in their localization and activity is mediated also by protein scaffolds. Scaffolds are a group of proteins that coordinate and allow specificity and fidelity by compartmentalizing kinases and their downstream substrates. These proteins provide ways to mediate the action of promiscuous enzymes by controlling their access to particular substrates (e.g., an anion channel in the cell membrane vs. a promoter of gene expression in the nucleus [117,118]).
The best-characterized class of kinase scaffolds is the A-kinase anchoring proteins (AKAPs) that compartmentalize protein kinase A (PKA). AKAPs were first identified as PKA protein scaffolds [119] and were demonstrated to play a critical role in signal transduction, placing binding effectors into specific subcellular locations, and optimizing signaling events on colocalized enzymes, receptors, and mRNA molecules to maintain cell homeostasis [120,121]. Protein scaffolds also have an integral role in PKC signaling, as they play a critical role in determining PKC isozyme localization. Furthermore, these protein scaffolds are vital in establishing substrate specificity [122,123,124]. Initial PPI studies by Mochly-Rosen and colleagues identified several PKC anchoring proteins, collectively termed receptors for activated C kinases (RACKs), which were shown to target a different PKC isozyme [125] (for reviews, see [126,127,128]). Over the years, additional protein scaffolds were identified, including 14-3-3 [129], annexins [130], HSPs [131], importins [132], and AKAPs that universally bind PKA [121]. While AKAPs have since been shown to interact with both PKA and PKC, PKC binding occurs through regions that are distinct from that of PKA anchoring [133,134].
RACKs bind to activated PKC isozymes in a selective and saturable manner, anchoring the PKC isozymes to their respective subcellular sites. For example, RACK1 (i.e., guanine nucleotide-binding protein subunit beta-2-like 1, or GNB2L1) is selective for PKCβII [135], and RACK2 (previously identified as β‘COP’ [136,137]) is selective for PKCε [138]. In general, RACKs are present in a particular location, their binding to PKC is dependent on a second messenger, PKC binding to RACKs is saturable and specific, and RACKs are neither a PKC substrate nor an inhibitor [139]. Both RACK1, which is selective to PKCβII, and RACK2, which is selective to PKCε, are members of the tryptophan-aspartate (WD) repeat family, containing seven repeats of the WD40 motif. The Mochly-Rosen laboratory not only identified and characterized several RACKs but also engineered peptides derived from scaffold proteins that resemble domains in specific PKC isozymes and proposed that these sequences participate in intramolecular interactions that clamp PKC in an autoinhibited state [140]. They hypothesized that peptides derived from these sequences disrupt the PPIs of specific PKC isozymes with their cognate RACKs and, therefore, modulate the cellular function of PKC isozymes [141]. Such peptides have been particularly valuable pharmacological tools to regulate the activity of PKC isozymes in cancer [142] and CVDs [143,144]. We further discuss these peptides, feathering their rational design and bioactivity vide infra. Another interesting example of the role of protein scaffolds in coordinating PKC activity was contributed by Zuker and colleagues, who demonstrated that the disruption of the interaction between the eye-specific PKC in Drosophila melanogaster and the PDZ scaffold InaD impairs light signaling [145]. Currently, it is well established that scaffold PPIs are an integral and vital part of PKC regulation, poising specific isozymes near key protein and lipid regulators as well as substrates. Therefore, disruption of these critical PPIs presents a unique avenue for specifically tuning PKC output without the inhibition of catalytic activity to reduce adverse effects (Figure 2).

3. PKC in Cancer

Individual cells constitute the fundamental units of life with a specific physiological purpose in different organs and organ systems. Under physiological conditions, cells would divide in a coordinated and controlled process, maintaining a homeostatic environment in the body. However, in pathological conditions, some cell types can become unresponsive to the cell-division regulatory signals and start to proliferate aberrantly; this clinical situation is most frequently termed cancer, although there are specific terms to describe the exact type of cellular growth dysregulation (hyperplasia, metaplasia, dysplasia, neoplasia, etc.). Cancer is a dynamic process that evolves in response to both genetic and epigenetic alterations, a highly challenging disease due to heterogeneity and robust compensatory mechanisms that modify cellular signaling pathways, giving rise to multiple cellular populations that confer resistance either de novo or in response to treatment. Hundreds of protein kinases play overlapping roles in cancer behavior, including cell transformation, tumor initiation, tumor progression, survival, proliferation, and recurrence. Kinases are associated with cancer via dysregulated gene expression and/or amplification, gene mutation, chromosomal translocation, epigenetic modification, and aberrant protein phosphorylation. Protein kinase dysfunction is implicated in multiple different components of carcinogenesis. Furthermore, protein kinases are among the early favored targets for cancer therapy. The development of protein kinase inhibitors (PKIs) has been a significant breakthrough in targeted cancer therapy, driving strategies against specific cancer-signaling pathways. Since the early success of the first SMKI, imatinib, which led to a paradigm shift in cancer therapy, the treatment of numerous additional cancers has proven clinically successful with kinase modulators. Many comprehensive reviews provide an overview of kinase-targeted drug discovery and development in relation to oncology, discussing the challenges, mechanisms of inhibition, outcomes, adverse effects, and future potential for kinase-targeted cancer therapies [13,14,31,74,146,147]. Currently, there are 58 agents targeting different protein kinases approved by the FDA for the treatment of cancer; 49 of these modulators are indicated for solid tumors including breast, lung, and colon tumors; 5 are indicated for nonsolid tumors, such as leukemias; and 4 (e.g., acalabrutinib, ibrutinib, imatinib, and midostaurin) are indicated for both solid and nonsolid tumors. Despite these encouraging results, problems with drug resistance and toxicity present major challenges for both clinical and experimental oncology [76,148].
The recognition of PKC as the long sought-after receptor for the tumor-promoting phorbol esters established the potential role of PKC in carcinogenesis and as a highly valuable target in oncology [149,150,151]. Phorbol esters (i.e., tetracyclic diterpenoids) are natural compounds derived from the plants of the family Euphorbiaceae, which mimic the action of the lipid second-messenger DAG by binding to the PKC C1 domain in a competitive manner. Early studies starting in the 1940s established that phorbol esters have potent activity as skin tumor promoters in mice [152]. However, unlike DAG, they are not readily metabolized and thus lead to constitutive activation of PKC, resulting in chronic loss, or downregulation, of PKC. The general assumption is that, upon phorbol ester activation, all classic and novel PKCs will be activated and phosphorylate specific substrates [153]. Moreover, there is an abundance of cancer-associated mutations in PKC that are loss-of-function (LOF) mutations and are found in a multitude of cancers. There are over 1000 cancer-associated somatic mutations in PKC isozymes identified to date. There are about 20–25% PKC mutations in several cancers, including melanoma, colorectal cancer, and lung squamous cell carcinoma [95,154]. The majority of these mutations are LOF related to processing phosphorylations, second-messenger binding, or catalysis. Numerous other mutations are in conserved motifs required for catalytic activity [155]. However, there is still a poor understanding of PKC isozyme-specific substrates and the distinct effector role in cancer. Among these unresolved issues is the role of PKC isozymes as tumor promoters versus tumor suppressors [156].
The expression and functions of PKC isozymes in cancerous cells largely depend on the type of cancer from which they originate. For example, PKCδ acts as an antiapoptotic regulator in chronic lymphocytic leukemia (CLL) [157] and as a proapoptotic factor in acute myeloid leukemia (AML) [158], and PKCα shows proliferative functions in several types of cancer [159] but has antiproliferative functions in colon cancer cells [160]. The PKC isoforms signal through multiple pathways and controls the expression of numerous genes relevant for cell-cycle progression, tumorigenesis, and metastatic dissemination, demonstrating variable expression profiles during cancer progression depending on cell types. The first reported cancer-associated mutation in PKC was one in PKCα found in human pituitary tumors [161,162]. While today we better understand the involvement of individual PKCs in various cancer types and in the context of specific oncogenic alterations, PKC contributions to malignancy remain a challenge, and the relevance of individual PKC isozymes in the progression of human cancer, in many cases, is still a matter of controversy.
Skin cancer is the most common cancer in the United States (US), when including melanomas and nonmelanomas, and about one in five Americans will develop skin cancer in their lifetime. Nonmelanomas are more common and include basal cell carcinoma (BCC) and squamous cell carcinoma (SCC), while melanoma is a more serious yet less common type of skin cancer. PKC isozymes exhibit specificities in their signals relative to the development of skin cancer. PKCα is highly abundant in the skin, and its activation is typically associated with increased tumor cell proliferation, invasiveness, and decreased differentiation, which was suggested by studies done with PKCα inhibitors [163,164]. PKCε induces SCC by an inhibition of apoptosis and enhancement of preneoplastic cell proliferation [151,156,165].
Breast cancer is the second most common cancer in women in the US, and about one in three of all new female cancer diagnoses each year are breast cancer. Breast cancer is also the second leading cause of cancer death in women. PKC isozyme expression and localization are tightly controlled during the processes of mammary gland differentiation and involution. Overexpression of several PKC isozymes has been reported in malignant breast tissue and breast cancer cell lines. PKCβ, PKCδ, and PKCη are found mainly in early-stage breast cancers and decrease with increasing stage, while PKCα and PKCζ levels typically increase as the cancer progresses [156,166].
Lung cancer is the third most common cancer in the US. The two key types of lung cancer are nonsmall cell lung cancer (NSCLC), which is the most common lung cancer accounting for about 85% of lung cancer cases, and small cell lung cancer (SCLC), which is the most aggressive form of lung cancer. Lung cancer is by far the leading cause of cancer death, making up almost 25% of all cancer deaths in both men and women. PKCα, PKCε, PKCη, PKCι/λ, and PKCζ are upregulated in lung cancer. Furthermore, overexpression of these PKC isozymes correlates with a worse prognosis in NSCLC patients, mainly due to tyrosine kinase-inhibitor resistance [156,167].
Prostate cancer is the second most common cancer in men and the fourth most common cancer overall. Prostate cancer is the second leading cause of cancer death in American men, behind only lung cancer, and about one man in eight will be diagnosed with prostate cancer during his lifetime. Various PKC isozymes play key roles during the progression of prostate cancer. PKCι/λ was overexpressed in invasive prostate cancer tissue, and its inhibition significantly promoted apoptosis and reduced the proliferation of prostate cancer cells [168], although there are other conflicting reports related to prostate cancer growth and metastasis as well [151,156].
Colorectal cancer (CRC) is the third most common cancer worldwide and accounts for eight percent of all cancer deaths both in men and women in the US. There are many types of CRC, the most common being adenocarcinoma, which develops in the lining of the large intestine (colon) or the distal end of the colon (rectum) and makes up 95% of all colorectal cancer cases. While, in general, the assumption is that increased PKC activation and expression promotes carcinogen-induced tumorigenesis, analysis of PKC gene-expression levels in CRC tissues revealed a downregulation of PKCβ. In addition, activation and overexpression of PKCβII dramatically reduced PKB phosphorylation and consequently reduced cell survival, suggesting that PKCβII has a tumor suppressor role in colon cancer [169].
Kidney cancer is less common but still among the top ten most common cancers in the US. The kidneys are made of a lot of small tubules, called the renal tubules, and most kidney cancers develop from these tubules. There are many types of kidney cancer, including renal cell carcinomas (RCC), transitional cell carcinomas, Wilms tumors, and renal sarcomas. Yet, RCC is the most common type of kidney cancer in adults and certainly the most well studied. Increased PKC activity has been noticed during the development and progression of RCC, and several PKC inhibitors have been found to decrease the invasiveness of aggressive human RCC cell lines, suggesting a proinvasive role for PKC in this context [170,171]. For example, increased PKCζ expression was associated with larger tumor size and worse survival, although PKCζ downregulation produced significant chemoresistance in RCC cell lines [172,173].
Bladder cancer is the second most common genitourinary malignancy in the US, the fourth most common cancer, and three times more common in men than in women [174]. Bladder cancer is a complex disease associated with high morbidity and mortality. Urothelial carcinoma (i.e., transitional cell carcinoma or TCC), is by far the most common type of bladder cancer, accounting for 95% of bladder cancer cases [175,176]. PKCα, PKCβ, PKCδ, PKCε, PKCη, and PKCζ have been studied in bladder cancer cells and tissues. Activated PKCα was increased in bladder cancer cells, and several studies demonstrated a correlation between PKCα activation and bladder cancer cell proliferation, survival, invasion, migration, and drug resistance [177,178,179].
Non-Hodgkin’s lymphoma (NHL) is a group of blood cancers originating from white blood cells (lymphocytes) that generally develop in the lymph nodes and lymphatic tissue. NHL accounts for about four percent of all cancers in the US [180,181]. PKCβ promotes B-cell receptor signaling, enhancing B-cell proliferation and survival, which is critical for B-cell-derived lymphomas [182]. PKCβ also modulates angiogenesis via vascular endothelial growth factor (VEGF), which enhances tumor progression [183] and is associated with worse prognoses in NHL [184]. PKCβ was especially overexpressed in treatment-refractory diffuse large B-cell lymphoma (DLBCL), an aggressive, and the most common, type of NHL [185,186].
Thyroid cancer is more common in women than men, and its incidence continues to rise worldwide [187,188]. A single-point mutation in PKCα (Gly for Asp substitution at position 294) was present in 50% of follicular thyroid neoplasms [189] and was associated with altered subcellular distribution and greater growth potential in rat fibroblasts [190]. Furthermore, a rearrangement in the gene-encoding PKCε was reported in a thyroid follicular carcinoma cell line, suggesting that these signaling proteins may play a role in thyroid tumorigenesis [155,191,192].
Endometrial cancer develops from the endometrium lining of the uterus, and the less common gynecological cancer uterine sarcoma develops from the myometrium muscle wall of the uterus [193]. PKCα plays a fundamental role in endometrial carcinogenesis by regulating endometrial cancer cell proliferation, anchor-independent growth, and xenograft tumorigenesis in animal models [194]. PKCδ is also a critical regulator of apoptosis and cell survival in endometrial cancer cells. There is a decrease in PKCδ expression in endometrial tumors, and endometrial cancer cell lines derived from poorly differentiated tumors exhibited reduced PKCδ levels relative to well-differentiated lines. In this context, decreased PKCδ expression may compromise the ability of cells to undergo apoptosis, perhaps conferring resistance to chemotherapy [195,196].
Pancreatic cancer accounts for about three percent of all cancers in the US and about seven percent of all cancer deaths, and it is the tenth most common cancer worldwide. There are two types of tumors that grow in the pancreas: exocrine and neuroendocrine tumors. More than 90% of all pancreatic tumors are exocrine tumors, and the most common kind of pancreatic cancer is called pancreatic ductal adenocarcinoma (PDAC) [197,198,199]. Several PKC isozymes have been shown to play a role in pancreatic cancer [200,201]. PKCι/λ was highly expressed in human pancreatic cancers, and high PKCι/λ expression predicted poor survival [202]. In addition, overexpression of PKCδ in cells induced a more malignant phenotype in vivo due to enhanced cell proliferation and survival [203].
Protein kinases were among the early and favored targets for precision cancer treatment. Since the PKC family has been well-characterized in numerous cancers, PKC became a particularly sought-after target for anticancer agents. Clinical data reveal reduced or elevated protein levels of PKC isozymes in tumor tissue compared with cognate normal tissue depending on the tissue of origin. For example, decreased levels of PKCβ, PKCδ, PKCε, and PKCη were reported in colon cancer [95], while increased levels were reported for PKCγ in colon cancer [204] and PKCι/λ in pancreatic cancer [202]. An additional level of complexity is that certain PKC isozymes are upregulated or downregulated in different cancers. For example, PKCα is upregulated in bladder, endometrial, and breast cancers, while it is downregulated in colorectal tumors and renal cell carcinomas. Yet, over three decades of clinical trials for various cancers using PKC inhibitors not only failed but, in some cases, worsened patient outcomes [205]. A meta-analysis of five clinical trials for NSCLC revealed worse patient outcomes when PKC inhibitors (enzastaurin, an ATP competitive inhibitor, or aprinocarsen, a PKCα antisense oligonucleotide) were combined with chemotherapy versus chemotherapy alone [205], highlighting the complexity of protein kinase signal transduction and the need for the development of more selective tools for basic research as well as therapeutic application.

4. PKC in Cardiovascular Diseases

Cardiovascular diseases are a general term for conditions affecting the heart or blood vessels, and they are the leading cause of death globally, taking an estimated 17.9 million lives each year, representing 32% of all global deaths. Almost half of all adults in the US have at least one form of heart disease, and about 659,000 people in the US die from heart disease each year, which is one in every four deaths per year. On average, one person dies from CVD every 36 s. In the US, CVDs claim more lives each year than all forms of cancer and chronic lower respiratory disease combined. Between 2017 and 2018, total direct and indirect costs of CVD were USD 378 billion, and they accounted for 12% of total US health expenditures during this period, which is more than any other major diagnostic group [206,207].
CVDs include all the diseases of the heart and circulation, yet there are many different types of CVDs, and the underlying mechanisms vary depending on the disease. CVDs have a multifactorial origin, and thus a mechanistic understanding of these diseases has been difficult to achieve because of clinical and biological heterogeneity. Furthermore, there is a strong component of race, gender, and environmental factors in the etiology of CVDs. CVDs include different pathologies, such as heart failure, ischemic heart disease, ischemia–reperfusion injury, arrhythmia, cardiomyopathies, and diseases of blood vessels (vascular diseases) including hypertension, atherosclerosis, and stroke. The pathogenesis of many of these CVDs involves protein kinases signaling.
In the heart, kinases mediate signal-transduction pathways that are vital for cardiac function, cation transport, myocardial metabolism, gene expression, cellular growth, and cell apoptosis. However, the relative importance of individual kinases is not clear, and many highly expressed cardiomyocyte kinases remain unstudied in this context. Protein kinase regulation and function are usually studied in proliferating cells in relation to cancer, for which they are attractive therapeutic targets, as previously discussed. Therefore, it is not surprising that most kinase inhibitors that have been approved or are in development are for the treatment of cancer. While the nature of kinase involvement in cancer is an intrinsic property mainly due to aberrant kinase activity that is a result of mutations, in CVDs, kinase-mediated pathogenesis is more frequently due to enhanced stimulation by activated neurohormonal systems. Nevertheless, the success of kinase inhibitors in cancer treatment has strongly supported their application to the treatment of CVDs as well. Furthermore, kinases that promote cancer are often required for cardiac function, and at least some of these kinase inhibitors have cardiotoxic effects in a significant percentage of patients, suggesting activity in this organ [208,209,210].
PKC isozymes are ubiquitously expressed in all tissues at all times of development, and they have emerged as important regulators of the cardiovascular system. The activation of PKC isozymes triggers various intracellular events influencing multiple physiological processes in the heart, including heart rate, contraction, and relaxation. Therefore, PKC isozymes have emerged as important regulators of cardiac function [211,212]. Many PKC isozymes, including PKCα, PKCβ, PKCγ, PKCδ, PKCε, PKCθ, PKCι/λ, and PKCζ, have been found in the hearts of different animal species [213,214,215]; yet, with the vast amount of studies on animal models, there exist only a few reports available that examined the expression of PKCs in humans. In the human heart, six PKC isoforms are expressed, including PKCα, PKCβ, PKCδ, PKCε, PKCι/λ, and PKCζ, while PKCγ and PKCθ are not present [216,217,218]. Even in other species, the relative content of each PKC isozyme in the heart tissue has been a controversial issue, as evidenced by work on rats [219,220], rabbits [221,222], guinea pigs [223], hamsters [224], and dogs [225].
To add to the complexity of these data, PKC expression levels in the heart change over time. PKCα, PKCβ, PKCε, and PKCζ expression is high in fetal and neonatal hearts but decreased in adult rat hearts [226]. The predominant PKC isozymes in the adult cardiovascular system are PKCα, PKCβ, PKCδ, and PKCε, which participate in various CVDs, including atherosclerosis, hypertension, atrial fibrillation, and cardiac hypertrophy. Interestingly, the activation of PKC isozymes in the cardiovascular system was demonstrated in some cases to have opposing effects. While activation of PKCε protects mitochondrial function prior to ischemic events or during reperfusion by activating aldehyde dehydrogenase 2 (ALDH2), removing toxic aldehyde-lipid peroxidation products [227], inhibiting L-type calcium channel activation, and preventing ventricular arrhythmias associated with ischemia–reperfusion injury [228], PKCδ activation mediates damage mainly by activating mitochondrial pyruvate dehydrogenase kinase, which inhibits pyruvate dehydrogenase and ATP regeneration, resulting in necrosis [229,230,231].
Based on the important role that PKC isozymes have in the cardiovascular system, several PKC inhibitors have been tested in animal models of CVDs. Ruboxistaurin (i.e., LY333531, or arxxant) is an orally active macrocyclic bisindolylmaleimide compound that specifically targets PKCβ, which acts as a competitive inhibitor by interacting with the ATP binding site. Ruboxistaurin has shown promising preclinical results in animal models of diabetic retinopathy and macular edema [232,233], and it is being studied as a systemic treatment for diabetic retinopathy, which is a microvascular complication of diabetes. In various myocardial infarction animal models, ruboxistaurin showed a cardioprotective effect. In a rodent model of myocardial infarction, ruboxistaurin protected against cardiac microvascular ischemia–reperfusion injury and reversed cardiac microvascular barrier dysfunction [234,235]. In a porcine myocardial infarction model, ruboxistaurin partially reversed cardiac remodeling by improving contractility [236].
Studies also suggest that PKCδ activity may promote reperfusion injury after ischemia. PKCδ inhibition during reperfusion can block these effects and protect the heart from further damage. Rottlerin is a polyphenol natural product isolated from the Asian tree Mallotus philippensis, which has been reported to selectively inhibit PKCδ translocation and reduce infarct size in a rat model of myocardial infarction [237]. However, there were problems with rottlerin related to a lack of specificity, causing significant off-target effects [238]. An attractive alternative approach is the inhibition of PKCδ with specific peptide allosteric inhibitors that can protect the heart from injury, which we will discuss later.
Atherosclerosis is the thickening or hardening of arteries caused by the buildup of plaque in the inner lining of an artery. When atherosclerosis affects the arteries that carry blood to the heart muscle, the terms coronary artery disease (CAD) or ischemic heart disease (IHD) are used to describe a group of CVDs where the heart arteries cannot deliver enough oxygen-rich blood to the heart. CAD encompasses the most common forms of CVD and leads to conditions called stable angina, unstable angina, myocardial infarction, and sudden cardiac death. About 18 million adults have CAD, and, in 2019, over 350,000 people in the US alone died from complications of CAD. Atherosclerosis is also a chronic inflammatory reaction, and it is the primary cause of myocardial infarction and stroke. Numerous studies have been conducted to investigate the significance of PKC in atherosclerosis. Current research suggests that a major role for PKC in atherosclerosis and restenosis relates to its involvement in programmed cell death (apoptosis), yet the exact role that apoptosis plays in atherosclerotic plaque formation is a matter of considerable debate [239,240].
Heart failure (HF) is a clinical syndrome characterized by an impaired ability of the ventricle (most often the left ventricle) to fill and/or eject blood. Heart failure is an ultimate state that can be caused by various insults or stressors to the heart and is characterized by cardiomyocyte hypertrophy, a reduced number of cardiomyocytes, and cardiac fibrosis. The American Heart Association (AHA) statistical update in 2015 reported that one in nine deaths have heart failure mentioned on the death certificate, and, in 2012, the total cost for heart failure was estimated to be about USD 31 billion. A projection shows that by 2030, the total cost of heart failure will increase to almost USD 70 billion [241]. The PKC family has been identified to play a role in the heart’s response to mechanical, electrical, or neurohumoral stimuli, in the case of disease leading to a maladaptive cardiac system characterized by fibrosis, apoptosis, and cardiac remodeling [144,242]. Levels of PKCα and PKCβ isozymes increase during heart-failure progression [215,243], and their inhibition has shown dramatic protective effects [143,244]. Although the role of PKCδ in the pathogenesis of heart failure is not entirely clear, many studies have confirmed that PKCδ is associated with the occurrence of heart failure, and it is activated in the early stages of pressure overload and heart failure [245]. Reduced cardiac expression of PKCε was found in rabbit left ventricular hypertrophy [246] and was also consistent in human left ventricular heart failure [216].
Hypertension (HTN) is a multifactorial disorder that involves pathological changes in the neuronal, renal, and vascular control mechanisms of blood pressure and may lead to various additional cardiovascular complications [247]; 68 million US citizens, or one in every three adults, have hypertension, which contributes to nearly 1000 deaths a day. In 2009, nearly 350,000 American deaths included hypertension as a primary or contributing cause. Annual costs directly attributable to hypertension are projected to more than double over the next two decades, bringing the projected annual total to USD 200 billion by 2030. An increased expression of PKC in vascular smooth muscle can cause excessive vasoconstriction, narrowing of blood vessels, and trophic vascular changes resulting in an increase in vascular resistance and hypertension [248,249,250]. PKCα overexpression has been implicated in the pathogenesis of hypertension [251,252,253], and an increased level of PKCβ was also demonstrated in a rat model of hypertension-induced cardiac dysfunction by a high-salt diet [254], which was consistent with the analysis of human tissue from heart-failure patients [67,215,216,255]. PKCδ is also involved in promoting hypertension via endothelin-1 signaling [231], and inhibition of PKCε in hypertension-induced heart failure led to reduced pathological remodeling and improved myocardial function [256].
Stroke is another complication of CVD. There are three main types of stroke, including ischemic stroke (85% of strokes), in which a blood clot prevents oxygenated blood from reaching an area of the brain; hemorrhagic stroke, in which a blood vessel ruptures; and transient ischemic attack, in which the blood flow to a part of the brain is inadequate for a brief period of time and normal blood flow resumes after a short amount of time with a resolution of symptoms. Stroke is the number five cause of death and a leading cause of disability in the US; about 800,000 people have a stroke each year, and around one person has a stroke every 40 s [257]. A largely conflicting body of work suggests that PKC is activated and plays a pathogenic role during stroke. In contrast, many studies also report a rapid loss of PKC activity and expression after ischemia, suggesting that PKC is degraded. These conflicting data can be explained by different experimental settings, including animal models, types of primary brain tissue, and the duration and intensities of ischemic–reperfusion injury. Furthermore, there are also likely opposing roles of individual PKC isozymes, as discussed previously [258,259,260,261,262].
Arrhythmias include a broad spectrum of disorders relating to heart rate and rhythm abnormalities. Approximately one in eighteen people, or five percent of the US population, have an arrhythmia, with atrial fibrillation being the most common, affecting about 20% of the general population at some time in their lives [263]. As PKCs play a critical role in the regulation of ion channels, the initiation and propagation of the cardiac action potential and thus cardiac excitability can be dramatically affected by PKC activity, yet conflicting data exist for the role of each specific PKC isozyme [228].
Cardiac arrest is the sudden loss of blood flow throughout the body resulting from the heart not being able to pump blood efficiently. There are more than 356,000 out-of-hospital cardiac arrests annually in the US, nearly 90% of them fatal. One estimate placed the total cost of cardiac arrest in the US at USD 33 billion each year [264,265]. Several PKC isozymes were demonstrated to be involved in various mechanisms leading to neuronal cell death after cardiac arrest, such as neurotransmitter release, regulation of ion channels, and modulation of synaptic function. Therefore, PKC isozymes have been suggested as valuable targets for the modulation of this system. In addition, PKCs have been implicated in both increased damage after cerebral ischemia as well as neuroprotection after ischemic preconditioning [217,266].
Myocardial infarction (MI) occurs when blood stops flowing properly to a part of the heart, and the heart muscle is subsequently injured due to a lack of oxygen supply. Often, myocardial infarction is a late-stage complication of coronary artery disease. Myocardial infarction is caused by persistent ischemia and hypoxia due to the inadequate function of coronary arteries, accompanied by a tremendous amount of subsequent myocardial necrosis. The principal treatment for myocardial infarction has focused on revascularization and reperfusion of blocked arteries [267], although a considerable portion of the cardiac damage leading to morbidity and mortality occurs with inflammation and the release of toxic species following reperfusion of the cardiac muscle. Every year, about 805,000 people in the US have a myocardial infarction. The estimated direct and indirect costs of myocardial infarctions from 2016 to 2017 were about USD 12 billion [268]. Increased expression of PKCα was observed following myocardial infarction [216], and inhibition of PKCβ in a postmyocardial infarction heart failure rat model improved cardiac function and was associated with reduced pathological myocardial remodeling [255]. While both PKCδ and PKCε are activated in the ischemic human heart, the mechanism and exact role of each PKC in the survival of cardiac cells remain unknown and controversial. These two isozymes have both parallel and opposing effects on the heart [269,270], indicating potential danger in the use of nonselective isozyme inhibitors and activators.

5. PKC in Other Human Diseases

The PKC family has long been considered to play a critical role in the pathogenesis of numerous human diseases due to the diversity of function and broad specificity of different PKC isozymes [101]. PKC isozymes regulate the phosphorylation of proteins that are associated with metabolic disorders, a cluster of various medical conditions that interfere with the body’s metabolism that are usually caused by genetic defects. Furthermore, PKC activation was demonstrated to be involved in signaling pathways dysregulated in metabolic disorders, expression of growth factors, and intracellular levels of ROS that cause cell damage. Therefore, individual PKC isoforms were explored as potential targets in several animal models of metabolic diseases, and some preliminary preclinical studies have identified individual PKC isoforms as potential targets for treating metabolic diseases [55,56,240,271]. For example, PKCβ is a major regulator of cholesterol and fatty acid metabolism in the liver, which phosphorylates and controls the nuclear translocation of the Farnesoid X receptor (FXR), a key regulator for cholesterol removal [272,273]. PKCβ expression levels increased in mice fed high-fat and high-cholesterol diets, and PKCβ-deficient mice are protected against diet-induced obesity and insulin resistance [274].
Several lines of evidence suggest abnormalities of PKC in various renal diseases as well. The two most studied kidney diseases are diabetic nephropathy, the leading cause of end-stage renal disease, and renal cancer [171]. PKC isozymes mediate renal development during the structuring of the embryological metanephros [275] and mediate inflammation in the mature kidney [276]. Studies on the molecular mechanisms of PKC activation in renal disease indicate that activation of PKC signaling, enhanced polyol signaling, increased oxidative stress, and overproduction of advanced glycation end products lead to hyperglycemia and promote diabetic vascular complications [277,278]. Aberrations in PKC expression and localization have also been implicated in the development of kidney cancer. PKCα, which has tumor-suppressor properties, was decreased in renal cell carcinoma (RCC or hypernephroma), the most common kind of kidney cancer in adults [279].
PKC isozymes have also been implicated in pulmonary disease and were specifically found to mediate lung functions, including permeability, contraction, migration, hypertrophy, proliferation, apoptosis, and secretion [280,281,282]. In many cases, the role of these isozymes in specific lung-disease pathogenesis is complicated and has not been fully defined yet. PKC signaling pathways contribute to the key cellular responses central to asthma [283,284,285], chronic obstructive pulmonary disease (COPD) [286], idiopathic pulmonary fibrosis (IPF) [287], acute respiratory distress syndrome (ARDS) [288,289], and lung cancer [167,290], to mention a few. Many markers of nuclear factor kappa B (NF-κB) activity were elevated in asthma and COPD, and the atypical PKCζ regulates NF-κB by phosphorylating its p65 subunit, which is essential for the transcriptional activity of NF-κB. Abdel-Halim et al. developed a specific PKCζ allosteric inhibitor, MA130, which inhibited the expression of only a subset of NF-κB-dependent genes, suggesting that targeting PKCζ may be more tolerable for the treatment of inflammatory lung disorder compared to global inhibition of NF-κB with various systemic toxicities [286].
PKC isozymes are highly enriched in the brain and play a vital role in regulating both the pre- and postsynaptic aspects of neurotransmission. Interestingly, PKC was discovered and initially characterized in the bovine brain [49,291]. PKC inhibition impairs brain function on multiple levels [292,293,294], as does abnormal PKC activation [295,296,297,298,299]. PKC isozymes influence short-term neuronal signaling (e.g., neurotransmitter release and ion flux), medium-term (e.g., receptor regulation), and long-term (e.g., cell proliferation, synaptic remodeling, and gene expression) mechanisms [300]. PKC is important for learning and memory, and PKC isozymes are classified as one of the cognitive kinases, as they were demonstrated to regulate synaptic transmission and affect substrates involved in information processing and storage (e.g., myristoylated alanine-rich C-kinase substrate (MARCKS), growth-associated protein 43 (GAP-43) and the N-methyl-D-aspartate (NMDA) receptor) [301,302].
As PKCs are ubiquitously expressed in the central nervous system (CNS) and mediate phosphorylation of transporters, ion channels, and GPCRs, isozymes are emerging as actionable targets in neurodegenerative diseases [2,299], including Parkinson’s disease [303,304,305,306,307], dementia [308,309], chronic pain [310,311,312,313], and Alzheimer’s disease [314,315,316,317]. For instance, β-amyloid (Aβ) peptides generated from sequential cleavage of amyloid precursor protein (APP) by β-site APP-cleaving enzyme 1 (BACE1) play a key role in the pathogenesis of Alzheimer’s disease. PKCδ overexpression increases lipopolysaccharide (LPS)-induced expression of nitric oxide synthase and cytokines, resulting in increased BACE1 expression in the brain. PKCδ knockdown or inhibition reduced BACE1 expression and corresponding Aβ levels, while PKCδ overexpression increased BACE1 expression and Aβ generation [318].
PKC isozymes are also involved in the transmission and modulation of pain from the spinal cord to the brain and specific cortical regions [311,312,319,320]. PKCε and PKCα appear to be involved in peripheral nociception [321], while PKCγ is important for central nociception [310]. For example, PKCε modulates nociception by activation of the transient receptor potential vanilloid receptor 1 (TRPV1) ion channel, a nonspecific cation channel activated by capsaicin [322,323,324]. Overall, PKC may serve as a master switch by which TRPV1 integrates heat and tissue damage to produce hyperalgesia [325].
PKC abnormalities have been implicated in the pathophysiology of psychiatric illnesses as well [326]. Evidence accumulated from multiple laboratories demonstrated that stress-induced deficits arise from excessive PKC signaling, which diminishes prefrontal neuronal firing [327]. Classic PKC isozymes are the most abundantly expressed PKCs in the CNS, and novel PKC isozymes also facilitate the release of the neurogenic growth factor neuregulin [328,329]. Mounting evidence points to a key role for PKC signaling in the pathology of bipolar disorder, a chronic and life-threatening disorder [330,331,332,333,334]. The mainstay treatment for bipolar disorder, lithium, exerts major effects on the PKC signaling cascade, for example by decreasing membrane-associated PKC, including isozyme-specific decreases in the classic PKCα and novel PKCε [335].
The PKC family also plays a critical role in skin keratinocyte function, including proliferation, differentiation, and neoplastic transformation, as well as in various dermatological diseases, including several forms of skin cancer [336]. PKC mediates squamous cell carcinoma (SCC) development and progression, characterized by reduced PKCδ expression and rat sarcoma virus (RAS) pathway activation [337]. Further, PKCε overexpression in mice led to the development of papilloma-free highly metastatic SCC [338]. The classic PKCα is also highly abundant in skin and has been shown to regulate the cell cycle and keratin cytoskeleton formation [339]. Downregulation of the atypical PKCζ contributed to tumorigenesis by releasing constraints on the activity of PKB [340]. In addition to skin cancer, PKC isozymes play roles in other inflammatory dermatological diseases, such as psoriasis, a chronic disease that affects the skin and joints [341]. In a small manufacturer-sponsored study, researchers evaluated sotrastaurin (AEB071), an orally administered potent and selective pan-PKC inhibitor, to 32 patients with psoriasis, resulting in a dose-dependent improvement in psoriasis during the two-week treatment period [342].
Several PKC isozymes are also important global mediators of inflammation and immunity. A major therapeutic approach to address inflammation has been to target the activity of kinases such as PKC that regulate the production of various cytokines and other inflammatory mediators. In the immune system, PKCs are involved in regulating signal-transduction pathways that are important for both innate and adaptive immunity, ultimately controlling the expression of key immune genes. Animal studies of PKC reported immune perturbations when each different isozyme was genetically deleted, with the exception of PKCγ, whose expression is restricted to the CNS and PKCι/λ, whose deletion results in embryonic lethality [343]. Interestingly, these defects are distinguishable between the different PKC isozymes, indicating that each isozyme has some unique role in the immune system [57,58,59,344]. In another model, PKC downregulation tempered the LPS-induced inflammatory response [345]. Several PKC inhibitors have shown promising results for the treatment of inflammatory diseases in preclinical in vivo studies: RO 32-0432, a PKC inhibitor that decreased phorbol ester-induced paw swelling in rats [346,347]; SKF 108753, a PKC inhibitor that inhibits the development of hind-paw swelling in an adjuvant-induced arthritis rat model [348,349]; and GÖ 6850, a PKC inhibitor that inhibited phorbol myristate acetate-induced edema, neutrophil influx, and vascular permeability in mice at levels comparable to indomethacin [350,351]. The efficacy of these PKC inhibitors suggests that PKC is a valid potential target in the development of anti-inflammatory agents [78].
PKC also plays a key role in pathogen entry and immune-mediated pathogen clearance. PKC isozymes regulate modifications to the cell membrane and cytoskeleton relevant to host–microbe interactions, including caveolae virus internalization by the host cell and membrane microdomains involved in endocytosis [352]. These mechanisms have been described for the entry of filovirus [353], human enterovirus [354], echovirus [355], human immunodeficiency virus (HIV) [356,357,358,359], rhabdovirus [360], alphavirus [361], poxvirus [362], herpesvirus [363,364], influenza virus [365,366], dengue virus [367], hepatitis B virus (HBV) [132], respiratory syncytial virus (RSV) [368,369], and adenovirus [362,370]. The activity of different PKC isozymes is context sensitive, and these kinases can be positive or negative regulators of macrophage function critical for host defense against infection. Depending on that context, PKC function can differentially affect pathogen virulence for bacteria (e.g., Escherichia coli [371], Staphylococcus aureus [372], Listeria monocytogenes [373], and Chlamydia trachomatis [374]), fungi (e.g., Candida albicans [375,376,377]), and parasites (e.g., Leishmania [378,379]). A specific example can be found in visceral leishmaniasis (i.e., Kala-azar), a fatal disease endemic to many parts of the tropical world in which intracellular Ca2+ is elevated. The causative Leishmania donovani contains lipophosphoglycan (LPG), resulting in classical PKC activity downregulation along with upregulation of Ca2+ independent atypical PKCζ expression [380] upon macrophage infection, which enables the parasites to survive within the host macrophages [381]. Table 1 includes examples of PKC isozymes in human diseases.

6. Targeting PKC

The vital role of dysregulated protein kinase activity in the etiology of human disease has fostered major drug-discovery programs over the past 40 years. Yet, without fully understanding the role of each isozyme and the signal transduction that is specific to each disease and organism, it is a challenge to identify appropriate targets for the development of new drugs, which is likely responsible for the failure of many lead compounds [14,385]. Given their many cellular roles, PKC isozymes are undoubtedly sought-after drug targets, but the similarity in their catalytic domains poses significant challenges in designing specific inhibitors. In addition, the complexity of their interactions; the many secondary messenger systems involved; and their cellular-, tissue-, and organism-specific variability make the development of selective modulators extremely difficult [382,386].
There are up to approximately 10,000 different proteins in a typical human cell, with an average length of about 400–450 amino acids. Taking into account the average amino acid frequencies (Ser, 8.5%; Thr 5.7%; Tyr, 3.0%) [387], a total of 17.2% of the amino acids are potentially available for phosphorylation. Therefore, there are about 700,000 potential phosphorylation targets for any given kinase, depending on the localization, activity, conformation, and availability of cofactors. Thus, there are clearly intricate mechanisms that determine the specificity between kinase and substrate, including local and distal interactions and the formation of complexes with scaffold/adaptor proteins [388].
PKC inhibitors are classified according to their sites of interaction within the PKC protein structure. Categories include (i) inhibitors of the catalytic domain, which are directed to either the substrate- or ATP-binding site and mainly consist of ATP-competitive SMKIs; (ii) modulators of the regulatory domain, which mainly target the phospholipid or phorbol ester binding site of the C1 domain, mimicking the binding of DAG; and (iii) inhibitors that disrupt PPIs (iiia) at a specific subcellular location or (iiib) with a specific substrate or scaffold protein (e.g., PKC and its corresponding RACK). These inhibitors that target PPIs provide an exciting approach to inhibiting PKC isozymes in a selective manner. This can be accomplished by targeting unique regions in each PKC isozyme, usually with peptides derived from the interacting domains of the protein and the binding protein (e.g., anchoring/scaffold protein or substrate), therefore modulating the phosphorylation of individual proteins. While various inhibitory compounds are available, only a few demonstrate specificity for either PKC alone or individual PKC isozymes. For the chemical structure and selectivity of major PKC modulators evaluated in the clinic, see Table 2.
Inhibitors of the catalytic domain. Since the catalytic domain demonstrates the greatest similarity in sequence and structure between PKC isozymes, developing isozyme-selective inhibitors for this site is extremely challenging [389]. Several small-molecule candidates with PKC activity have been identified by high-throughput screening followed by medicinal chemistry optimization or by in silico screening. These molecules usually act by competing with the ATP binding site to block the phosphotransferase activity of their targets. Bisindolylmaleimides are a family of structurally related compounds and are the most studied PKC SMKIs that act as ATP-competitive inhibitors. Bisindolylmaleimides bind to the ATP-binding pocket, stabilizing PKC in the activated conformation, and limiting substrate phosphorylation. Staurosporine, which belongs to the bisindolylmaleimide class, is a prototypical ATP competitor originally isolated in 1977 from the bacterium Streptomyces staurosporeus. The compound acts as a broad-spectrum protein kinase inhibitor, including PKC, as well as several other serine/threonine kinases. Based on the native compound, the synthetic bisindolylmaleimide enzastaurin was developed. Enzastaurin is a more selective PKC inhibitor compared to staurosporine, initially believed to selectively inhibit the PKCβ isozyme [390]. However, later studies demonstrated that enzastaurin inhibits other PKC isozymes as well as several additional protein kinases [391].
Modulators of the regulatory domain. The second group of PKC regulators are compounds that target the C1 domain at the regulatory region, mainly by mimicking the binding of the C1 domain to DAG. To better characterize the DAG binding site of the C1 domain [392], a medicinal chemistry study designed potent PKC activators, including DAG-lactones [393], which demonstrate increased isozyme affinity specificity (e.g., compound 57 for PKCα [394] and AJH-836 for novel PKCs [395]) compared to natural DAGs [396]. The best-characterized compound that targets the C1 domain is Bryostatin-1, a partial agonist resulting in PKC autophosphorylation and translocation to the cell membrane [397,398]. Bryostatin-1 is a macrocyclic lactone derived from a marine invertebrate that binds to the regulatory domain of PKC; interestingly, short-term exposure to bryostatin-1 promoted PKC activation, whereas prolonged exposure promoted PKC downregulation [399,400]. Bryostatin-1 has been investigated for its anticancer activity in numerous hematological and solid-tumor cell lines and was found to inhibit proliferation, induce differentiation, and promote apoptosis. Furthermore, phase I and II clinical trials in a wide range of tumors showed promising activity [400]. Nevertheless, as the C1 binding site is highly conserved among PKC isozymes, it is challenging to develop isozyme-selective compounds. Since different isozymes have diverse functions, modulation of PKC signaling by a compound with pan-PKC activity can have vast cellular effects, including adverse effects.
To achieve superior isozyme specificity, two main approaches have been developed, based on the use of antisense oligonucleotides or peptides. Antisense oligonucleotides demonstrated positive results in preclinical models; for example, early studies with antisense inhibition of PKCα demonstrated that it enhanced the antitumor effects of cisplatin in subcutaneously injected H460 cells [401], and additional studies showed that it significantly impaired nonsmall cell lung cancer tumor growth [402]. While this work and other studies provide a rationale for using specific PKC antisense oligonucleotides to reduce PKC in clinical trials, another PKCα antisense oligonucleotide failed in clinical trials for cancer [205]. An alternative approach to the development of ATP-competitive inhibitors is the use of peptides that target allosteric PPI sites (will be discussed in Part II).

7. PKC Inhibitors in Clinical Trials

ATP competitive SMKIs have been broadly developed and applied in clinical trials (Table 3). For example, midostaurin (4′-N-benzoylstaurosporine, also PKC412 or CGP 41251) is a staurosporine analog isolated from Streptomyces staurosporeus. The compound is an ATP-competitive inhibitor with low inhibitory activity for PKC. Midostaurin inhibits multiple protein kinases, reduces the growth of cancer cells in vitro, and reverses P-glycoprotein-mediated multidrug resistance in cancer cell lines [403]. FMS-like tyrosine kinase 3 (FLT3) mutations are associated with high leukemic burden and poor prognosis in patients with AML. Midostaurin was approved by the FDA in April 2017 for the treatment of newly diagnosed adult AML patients with FLT3-positive mutations, where it was more cost-effective than the previous standard of care [404]. While midostaurin was originally developed as a PKC inhibitor, its success in early trials was mainly due to the inhibition of other tyrosine kinases. Further clinical trials of midostaurin are in progress [405] (ClinicalTrials.gov ID NCT05488613). Enzastaurin is a cyclic bisindolylmaleimide and an ATP competitive inhibitor that targets PKCβ. This compound was tested in several phase I and phase II clinical trials alone and in combination with other drugs, mainly for oncologic indications (e.g., glioma [406], multiple myeloma [407], solid-tumor brain metastasis [408], ovarian carcinoma [409], lymphoma [410,411], breast cancer [412], NSCLC [413,414], glioblastoma [415], prostate cancer [416], ovarian cancer [417], metastatic colorectal cancer [418], and pancreatic cancer [419]). In these trials, enzastaurin alone or in combination failed to show any significant clinical benefit [420]. Other ATP competitive PKC inhibitors include sotrastaurin, a potent and selective pan-PKC inhibitor that exhibits immunosuppressive functions (e.g., inhibition of T-cell activation [421]) but was plagued by side effects and a lack of efficacy for the indications tested. Moreover, another nonselective PKC inhibitor called flosequinan (manoplax) previously used for congestive heart failure was withdrawn from the US market (October 1993) due to increased hospitalization and mortality [422].
While several C1 domain-binding PKC inhibitors (e.g., DAG competitive PKC inhibitors) have been reported, there are no published findings to describe the use of these compounds in clinical trials. An antisense oligonucleotide-targeting PKCα called aprinocarsen (ISIS 3521/LY900003) was evaluated in several clinical trials for recurrent high-grade astrocytomas [423], NSCLC [424,425], ovarian carcinoma [426], prostate cancer [427], metastatic colorectal cancer [428], and NHL [429]. However, aprinocarsen treatment alone or in combination with anticancer drugs failed to improve survival or confer other clinical benefit [420].
Table 3. Summary of clinically studied PKC modulators.
Table 3. Summary of clinically studied PKC modulators.
IndicationCompoundProposed MechanismOutcomeRegulatory StatusRefs.
OncologyBryostatinNonselective PKC activatorNo benefitNot approved[430,431,432,433,434,435,436,437,438,439,440,441,442,443]
AprinocarsenPKCα inhibitorNo benefitNot approved[424,427,428,429,444,445]
Enzastaurin (LY317615)PKCβ inhibitorNo benefitNot approved[407,409,410,446,447,448,449,450]
TamoxifenNonselective PKC inhibitor at high doses, ER inhibitorUsed in the management of many breast and gynecologic cancers; failed trials for other malignanciesApproved[451,452,453]
MidostaurinNonselective PKC inhibitor, FLT3 inhibitorUsed in treatment of FLT3-mutated AMLApproved[454]
7-Hydroxystaurosporine (UCN-01)Nonselective PKC inhibitor, CHK1 inhibitorNo benefitNot approved[455,456,457]
PMANonspecific PKC activatorNo benefitNot approved
SafingolPKCβI, PKCδ, and PKCε inhibitor; PI3K inhibitorNo benefitNot approved[458,459]
12-O-tetradecanoylphorbol-13-acetateNonselective PKC activatorNo benefit, severe side effectsNot approved[460]
Diabetes mellitusRuboxistaurin(LY333531)PKCβ inhibitorImproved diabetic retinopathy but not nephropathy in early studies, minimal effect on neuropathyNot approved[461,462,463,464,465,466,467,468,469]
CardiologyDelcasertib (KAI-9803)PKCδ inhibitorNo benefitNot approved[470,471,472]
FlosequinanNonselective PKC inhibitorIncreased hospitalizations and HF mortality; early study terminationWithdrawn[422,473]
Volatile anestheticsPKCε activatorReduced troponin I, inotrope requirements, and length of hospitalizationApproved for other indications[474,475,476,477,478]
AdenosinePKCε activatorReduced MI, mortality, vasopressor requirementsApproved for other indications[479,480,481]
AcadesinePKCε activator, AMPK activatorNo reduction in death, MI, or strokeNot approved[482,483,484]
Bipolar disorderEndoxifenTamoxifen metabolite with four-fold increased PKC inhibitionImproved time to remissionNot approved[485]
NociceptionKAI-1678Inhibits PKCε translocationNo benefitNot approved[486,487]
InflammationSotrastaurin (AEB071)Nonselective PKC inhibitorWorse outcomes in transplant rejection, no benefit in malignancy or autoimmune trialsNot approved[488,489,490,491]
Alzheimer’s Disease Bryostatin Nonselective PKC activatorPrimary endpoint was not significantNot approved[492]
PKC, protein kinase C; HF, heart failure; MI, myocardial infarction; ER, estrogen receptor; AMPK, AMP-activated protein kinase; FLT3, fms-like tyrosine kinase 3; AML, acute myeloid leukemia; CHK1, checkpoint kinase 1; PMA, phorbol 12-myristate 13-acetate; PI3K, phosphoinositide 3-kinase. Adapted from refs. [382,420].

8. Summary

In this review, we discuss the extensive role PKC plays in cancer and CVDs, making these proteins highly sought-after therapeutic targets. We also emphasize the specific PKC isozymes related to specific conditions in these domains. The last section summarizes the current molecules (mainly SMKIs) used in the clinic as well as recent clinical trials that have been conducted. In particular, we discuss tamoxifen and midostaurin, which are both clinically active PKC inhibitors. Tamoxifen (Nolvadex) is a nonsteroidal antiestrogen used to treat estrogen-receptor-positive breast cancers, and midostaurin (Rydapt or Tauritmo) is a multitarget kinase inhibitor for the treatment of adult patients with newly diagnosed acute myeloid leukemia (AML). While both medications are available for human use, neither of them inhibits PKC selectively. In the subsequent article, we will elaborate on additional therapeutic approaches for PKC, focusing on targeting allosteric sites and their unique advantages.

Author Contributions

Writing—original draft, S.S. and M.Z.; Writing—review & editing, N.Q. and S.J.S.R.; Project administration, S.J.S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, J.L.; Yaron, T.M.; Huntsman, E.M.; Kerelsky, A.; Song, J.; Regev, A.; Lin, T.-Y.; Liberatore, K.; Cizin, D.M.; Cohen, B.M.; et al. An atlas of substrate specificities for the human serine/threonine kinome. Nature 2023, 613, 759–766. [Google Scholar] [CrossRef] [PubMed]
  2. Benn, C.L.; Dawson, L.A. Clinically Precedented Protein Kinases: Rationale for Their Use in Neurodegenerative Disease. Front. Aging Neurosci. 2020, 12, 242. [Google Scholar] [CrossRef] [PubMed]
  3. Patterson, H.; Nibbs, R.; McInnes, I.; Siebert, S. Protein kinase inhibitors in the treatment of inflammatory and autoimmune diseases. Clin. Exp. Immunol. 2014, 176, 1–10. [Google Scholar] [CrossRef] [PubMed]
  4. Lapenna, S.; Giordano, A. Cell cycle kinases as therapeutic targets for cancer. Nat. Rev. Drug Discov. 2009, 8, 547–566. [Google Scholar] [CrossRef]
  5. Tsai, C.-J.; Nussinov, R. The molecular basis of targeting protein kinases in cancer therapeutics. In Seminars in Cancer Biology; Elsevier: Amsterdam, The Netherlands, 2013. [Google Scholar]
  6. Amin, F.; Ahmed, A.; Feroz, A.; Khaki, P.S.S.; Khan, M.S.; Tabrez, S.; Zaidi, S.K.; Abdulaal, W.H.; Shamsi, A.; Khan, W.; et al. An update on the association of protein kinases with cardiovascular diseases. Curr. Pharm. Des. 2019, 25, 174–183. [Google Scholar] [CrossRef]
  7. Krebs, E.G.; Fischer, E.H. The phosphorylase b to a converting enzyme of rabbit skeletal muscle. Biochim. Biophys. Acta 1956, 20, 150–157. [Google Scholar] [CrossRef] [PubMed]
  8. Krebs, E.G.; Kent, A.B.; Fischer, E.H. The muscle phosphorylase b kinase reaction. J. Biol. Chem. 1958, 231, 73–83. [Google Scholar] [CrossRef]
  9. Cohen, P. Protein kinases-the major drug targets of the twenty-first century? Nat. Rev. Drug Discov. 2002, 1, 309–315. [Google Scholar] [CrossRef]
  10. Eglen, R.; Reisine, T. Drug discovery and the human kinome: Recent trends. Pharmacol. Ther. 2011, 130, 144–156. [Google Scholar] [CrossRef]
  11. Arencibia, J.M.; Pastor-Flores, D.; Bauer, A.F.; Schulze, J.O.; Biondi, R.M. AGC protein kinases: From structural mechanism of regulation to allosteric drug development for the treatment of human diseases. Biochim. Biophys. Acta 2013, 1834, 1302–1321. [Google Scholar] [CrossRef]
  12. Kenakin, T.P. Chapter 6—Enzymes as Drug Targets. In Pharmacology in Drug Discovery and Development, 2nd ed.; Kenakin, T.P., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 131–156. [Google Scholar]
  13. Attwood, M.M.; Fabbro, D.; Sokolov, A.V.; Knapp, S.; Schiöth, H.B. Trends in kinase drug discovery: Targets, indications and inhibitor design. Nat. Rev. Drug Discov. 2021, 20, 839–861. [Google Scholar] [CrossRef] [PubMed]
  14. Cohen, P.; Cross, D.; Jänne, P.A. Kinase drug discovery 20 years after imatinib: Progress and future directions. Nat. Rev. Drug Discov. 2021, 20, 551–569. [Google Scholar] [CrossRef] [PubMed]
  15. Fletcher, L. Approval heralds new generation of kinase inhibitors? Nat. Biotechnol. 2001, 19, 599–600. [Google Scholar] [CrossRef] [PubMed]
  16. Quintás-Cardama, A.; Cortes, J. Molecular biology of bcr-abl1–positive chronic myeloid leukemia. Blood 2009, 113, 1619–1630. [Google Scholar] [CrossRef] [PubMed]
  17. Hantschel, O. Structure, regulation, signaling, and targeting of abl kinases in cancer. Genes. Cancer 2012, 3, 436–446. [Google Scholar] [CrossRef] [PubMed]
  18. O’Brien, S.G.; Guilhot, F.; Larson, R.A.; Gathmann, I.; Baccarani, M.; Cervantes, F.; Cornelissen, J.J.; Fischer, T.; Hochhaus, A.; Hughes, T.; et al. Imatinib compared with interferon and low-dose cytarabine for newly diagnosed chronic-phase chronic myeloid leukemia. N. Engl. J. Med. 2003, 348, 994–1004. [Google Scholar] [CrossRef] [PubMed]
  19. Deininger, M.; O’Brien, S.G.; Guilhot, F.; Goldman, J.M.; Hochhaus, A.; Hughes, T.P.; Radich, J.P.; Hatfield, A.K.; Mone, M.; Filian, J.; et al. International randomized study of interferon vs. STI571 (IRIS) 8-year follow up: Sustained survival and low risk for progression or events in patients with newly diagnosed chronic myeloid leukemia in chronic phase (CML-CP) treated with imatinib. Blood 2009, 114, 1126. [Google Scholar] [CrossRef]
  20. Wu, P.; Nielsen, T.E.; Clausen, M.H. FDA-approved small-molecule kinase inhibitors. Trends Pharmacol. Sci. 2015, 36, 422–439. [Google Scholar] [CrossRef]
  21. Wu, P.; Nielsen, T.E.; Clausen, M.H. Small-molecule kinase inhibitors: An analysis of FDA-approved drugs. Drug Discov. Today 2016, 21, 5–10. [Google Scholar] [CrossRef]
  22. de la Torre, B.G.; Albericio, F. The Pharmaceutical Industry in 2021. An Analysis of FDA Drug Approvals from the Perspective of Molecules. Molecules 2022, 27, 1075. [Google Scholar] [CrossRef]
  23. Fischer, P.M. Approved and Experimental Small-Molecule Oncology Kinase Inhibitor Drugs: A Mid-2016 Overview. Med. Res. Rev. 2017, 37, 314–367. [Google Scholar] [CrossRef] [PubMed]
  24. Xie, Z.; Yang, X.; Duan, Y.; Han, J.; Liao, C. Small-molecule kinase inhibitors for the treatment of nononcologic diseases. J. Med. Chem. 2021, 64, 1283–1345. [Google Scholar] [CrossRef] [PubMed]
  25. Garnock-Jones, K.P. Ripasudil: First Global Approval. Drugs 2014, 74, 2211–2215. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, W.; Poh, A.L.; Tang, W.H.W. Novel Insights and Treatment Strategies for Right Heart Failure. Curr. Heart Fail Rep. 2018, 15, 141–155. [Google Scholar] [CrossRef]
  27. Lovly, C.M.; Shaw, A.T. Molecular pathways: Resistance to kinase inhibitors and implications for therapeutic strategies. Clin. Cancer Res. 2014, 20, 2249–2256. [Google Scholar] [CrossRef] [PubMed]
  28. Djikic, T.; Gagic, Z.; Nikolic, K. Design and Discovery of Kinase Inhibitors Using Docking Studies. In Molecular Docking for Computer-Aided Drug Design; Elsevier: Amsterdam, The Netherlands, 2021; pp. 337–365. [Google Scholar]
  29. Wilhelm, S.; Carter, C.; Lynch, M.; Lowinger, T.; Dumas, J.; Smith, R.A.; Schwartz, B.; Simantov, R.; Kelley, S. Discovery and development of sorafenib: A multikinase inhibitor for treating cancer. Nat. Rev. Drug Discov. 2006, 5, 835–844. [Google Scholar] [CrossRef] [PubMed]
  30. Gild, M.L.; Bullock, M.; Robinson, B.G.; Clifton-Bligh, R. Multikinase inhibitors: A new option for the treatment of thyroid cancer. Nat. Rev. Endocrinol. 2011, 7, 617–624. [Google Scholar] [CrossRef]
  31. Gild, M.L.; Tsang, V.H.M.; Clifton-Bligh, R.J.; Robinson, B.G. Multikinase inhibitors in thyroid cancer: Timing of targeted therapy. Nat. Rev. Endocrinol. 2021, 17, 225–234. [Google Scholar] [CrossRef]
  32. Smolinski, M.P.; Bu, Y.; Clements, J.; Gelman, I.H.; Hegab, T.; Cutler, D.L.; Fang, J.W.; Fetterly, G.; Kwan, R.; Barnett, A.; et al. Discovery of Novel Dual Mechanism of Action Src Signaling and Tubulin Polymerization Inhibitors (KX2-391 and KX2-361). J. Med. Chem. 2018, 61, 4704–4719. [Google Scholar] [CrossRef]
  33. Sakamoto, K.M.; Kim, K.B.; Kumagai, A.; Mercurio, F.; Crews, C.M.; Deshaies, R.J. Protacs: Chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proc. Natl. Acad. Sci. USA 2001, 98, 8554–8559. [Google Scholar] [CrossRef]
  34. Pettersson, M.; Crews, C.M. PROteolysis TArgeting Chimeras (PROTACs)—Past, present and future. Drug Discov. Today Technol. 2019, 31, 15–27. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, F.; Cai, M.; Shao, L.; Zhang, J. Targeting Protein Kinases Degradation by PROTACs. Front. Chem. 2021, 9, 679120. [Google Scholar] [CrossRef] [PubMed]
  36. Békés, M.; Langley, D.R.; Crews, C.M. PROTAC targeted protein degraders: The past is prologue. Nat. Rev. Drug Discov. 2022, 21, 181–200. [Google Scholar] [CrossRef] [PubMed]
  37. Qi, S.-M.; Dong, J.; Xu, Z.-Y.; Cheng, X.-D.; Zhang, W.-D.; Qin, J.-J. PROTAC: An Effective Targeted Protein Degradation Strategy for Cancer Therapy. Front. Pharmacol. 2021, 12, 692574. [Google Scholar] [CrossRef] [PubMed]
  38. Goulet, D.R.; Atkins, W.M. Considerations for the design of antibody-based therapeutics. J. Pharm. Sci. 2020, 109, 74–103. [Google Scholar] [CrossRef]
  39. Hudis, C.A. Trastuzumab—Mechanism of action and use in clinical practice. N. Engl. J. Med. 2007, 357, 39–51. [Google Scholar] [CrossRef]
  40. Jonker, D.J.; O’Callaghan, C.J.; Karapetis, C.S.; Zalcberg, J.R.; Tu, D.; Au, H.-J.; Berry, S.R.; Krahn, M.; Price, T.; Simes, R.J.; et al. Cetuximab for the treatment of colorectal cancer. N. Engl. J. Med. 2007, 357, 2040–2048. [Google Scholar] [CrossRef]
  41. Gharwan, H.; Groninger, H. Kinase inhibitors and monoclonal antibodies in oncology: Clinical implications. Nat. Rev. Clin. Oncol. 2016, 13, 209–227. [Google Scholar] [CrossRef]
  42. Sun, X.; Gao, H.; Yang, Y.; He, M.; Wu, Y.; Song, Y.; Tong, Y.; Rao, Y. PROTACs: Great opportunities for academia and industry. Signal Transduct. Target. Ther. 2019, 4, 64. [Google Scholar] [CrossRef]
  43. Li, X.; Song, Y. Proteolysis-targeting chimera (PROTAC) for targeted protein degradation and cancer therapy. J. Hematol. Oncol. 2020, 13, 50. [Google Scholar] [CrossRef]
  44. Imai, K.; Takaoka, A. Comparing antibody and small-molecule therapies for cancer. Nat. Rev. Cancer 2006, 6, 714–727. [Google Scholar] [CrossRef] [PubMed]
  45. Zhan, M.M.; Hu, X.Q.; Liu, X.X.; Ruan, B.F.; Xu, J.; Liao, C. From monoclonal antibodies to small molecules: The development of inhibitors targeting the PD-1/PD-L1 pathway. Drug Discov. Today 2016, 21, 1027–1036. [Google Scholar] [CrossRef] [PubMed]
  46. Lu, R.-M.; Hwang, Y.-C.; Liu, I.-J.; Lee, C.-C.; Tsai, H.-Z.; Li, H.-J.; Wu, H.-C. Development of therapeutic antibodies for the treatment of diseases. J. Biomed. Sci. 2020, 27, 1–30. [Google Scholar] [CrossRef] [PubMed]
  47. Hanks, S.K.; Quinn, A.M.; Hunter, T. The protein kinase family: Conserved features and deduced phylogeny of the catalytic domains. Science 1988, 241, 42–52. [Google Scholar] [CrossRef]
  48. Hanks, S.K.; Hunter, T. Protein kinases 6. The eukaryotic protein kinase superfamily: Kinase (catalytic) domain structure and classification. FASEB J. 1995, 9, 576–596. [Google Scholar] [CrossRef]
  49. Takai, Y.; Kishimoto, A.; Inoue, M.; Nishizuka, Y. Studies on a cyclic nucleotide-independent protein kinase and its proenzyme in mammalian tissues. I. Purification and characterization of an active enzyme from bovine cerebellum. J. Biol. Chem. 1977, 252, 7603–7609. [Google Scholar] [CrossRef]
  50. Takai, Y.; Kishimoto, A.; Iwasa, Y.; Kawahara, Y.; Mori, T.; Nishizuka, Y. Calcium-dependent activation of a multifunctional protein kinase by membrane phospholipids. J. Biol. Chem. 1979, 254, 3692–3695. [Google Scholar] [CrossRef]
  51. Hagiwara, M. Alternative splicing: A new drug target of the post-genome era. Biochim. Et Biophys. Acta (BBA) Proteins Proteom. 2005, 1754, 324–331. [Google Scholar] [CrossRef]
  52. Kim, J.D.; Seo, K.W.; Lee, E.A.; Quang, N.N.; Cho, H.R.; Kwon, B. A novel mouse PKCδ splice variant, PKCδIX, inhibits etoposide-induced apoptosis. Biochem. Biophys. Res. Commun. 2011, 410, 177–182. [Google Scholar] [CrossRef]
  53. Rosse, C.; Linch, M.; Kermorgant, S.; Cameron, A.J.; Boeckeler, K.; Parker, P.J. PKC and the control of localized signal dynamics. Nat. Rev. Mol. Cell Biol. 2010, 11, 103–112. [Google Scholar] [CrossRef]
  54. Kikkawa, U. The story of PKC: A discovery marked by unexpected twists and turns. IUBMB Life 2019, 71, 697–705. [Google Scholar] [CrossRef] [PubMed]
  55. Geraldes, P.; King, G.L. Activation of protein kinase C isoforms and its impact on diabetic complications. Circ. Res. 2010, 106, 1319–1331. [Google Scholar] [CrossRef] [PubMed]
  56. Faria, A.; Persaud, S.J. Cardiac oxidative stress in diabetes: Mechanisms and therapeutic potential. Pharmacol. Ther. 2017, 172, 50–62. [Google Scholar] [CrossRef]
  57. Spitaler, M.; Cantrell, D.A. Protein kinase C and beyond. Nat. Immunol. 2004, 5, 785–790. [Google Scholar] [CrossRef]
  58. Zanin-Zhorov, A.; Dustin, M.L.; Blazar, B.R. PKC-theta function at the immunological synapse: Prospects for therapeutic targeting. Trends Immunol. 2011, 32, 358–363. [Google Scholar] [CrossRef]
  59. Altman, A.; Kong, K.F. Protein kinase C inhibitors for immune disorders. Drug Discov. Today 2014, 19, 1217–1221. [Google Scholar] [CrossRef]
  60. Sipka, S.; Bíró, T.; Czifra, G.; Griger, Z.; Gergely, P.; Brugós, B.; Tarr, T. The role of protein kinase C isoenzymes in the pathogenesis of human autoimmune diseases. Clin. Immunol. 2022, 241, 109071. [Google Scholar] [CrossRef]
  61. Deka, S.J.; Trivedi, V. Potentials of PKC in cancer progression and anticancer drug development. Curr. Drug Discov. Technol. 2019, 16, 135–147. [Google Scholar] [CrossRef]
  62. Simonis, G.; Braun, M.U.; Kirrstetter, M.; Schön, S.P.; Strasser, R.H. Mechanisms of myocardial remodeling: Ramiprilat blocks the expressional upregulation of protein kinase C-epsilon in the surviving myocardium early after infarction. J. Cardiovasc. Pharmacol. 2003, 41, 780–787. [Google Scholar] [CrossRef]
  63. Koyanagi, T.; Noguchi, K.; Ootani, A.; Inagaki, K.; Robbins, R.C.; Mochly-Rosen, D. Pharmacological inhibition of epsilon PKC suppresses chronic inflammation in murine cardiac transplantation model. J. Mol. Cell Cardiol. 2007, 43, 517–522. [Google Scholar] [CrossRef]
  64. Marrocco, V.; Bogomolovas, J.; Ehler, E.; Dos Remedios, C.G.; Yu, J.; Gao, C.; Lange, S. PKC and PKN in heart disease. J. Mol. Cell. Cardiol. 2019, 128, 212–226. [Google Scholar] [CrossRef] [PubMed]
  65. Inagaki, K.; Churchill, E.; Mochly-Rosen, D. Epsilon protein kinase C as a potential therapeutic target for the ischemic heart. Cardiovasc. Res. 2006, 70, 222–230. [Google Scholar] [CrossRef] [PubMed]
  66. Ferreira, J.C.; Brum, P.C.; Mochly-Rosen, D. betaIIPKC and epsilonPKC isozymes as potential pharmacological targets in cardiac hypertrophy and heart failure. J. Mol. Cell Cardiol. 2011, 51, 479–484. [Google Scholar] [CrossRef] [PubMed]
  67. Ferreira, J.C.; Koyanagi, T.; Palaniyandi, S.S.; Fajardo, G.; Churchill, E.N.; Budas, G.; Disatnik, M.H.; Bernstein, D.; Brum, P.C.; Mochly-Rosen, D. Pharmacological inhibition of betaIIPKC is cardioprotective in late-stage hypertrophy. J. Mol. Cell. Cardiol. 2011, 51, 980–987. [Google Scholar] [CrossRef] [PubMed]
  68. Hardman, C.; Ho, S.; Shimizu, A.; Luu-Nguyen, Q.; Sloane, J.L.; Soliman, M.S.A.; Marsden, M.D.; Zack, J.A.; Wender, P.A. Synthesis and evaluation of designed PKC modulators for enhanced cancer immunotherapy. Nat. Commun. 2020, 11, 1879. [Google Scholar] [CrossRef] [PubMed]
  69. Kanev, G.K.; de Graaf, C.; de Esch, I.J.; Leurs, R.; Würdinger, T.; Westerman, B.A.; Kooistra, A.J. The Landscape of Atypical and Eukaryotic Protein Kinases. Trends Pharmacol. Sci. 2019, 40, 818–832. [Google Scholar] [CrossRef] [PubMed]
  70. Zhang, Y.; Zhang, H.; Ghosh, D.; Williams, R.O., 3rd. Just how prevalent are peptide therapeutic products? A critical review. Int. J. Pharm. 2020, 587, 119491. [Google Scholar] [CrossRef]
  71. Huang, L.; Jiang, S.; Shi, Y. Tyrosine kinase inhibitors for solid tumors in the past 20 years (2001–2020). J. Hematol. Oncol. 2020, 13, 143. [Google Scholar] [CrossRef]
  72. Wang, B.; Wu, H.; Hu, C.; Wang, H.; Liu, J.; Wang, W.; Liu, Q. An overview of kinase downregulators and recent advances in discovery approaches. Signal Transduct. Target. Ther. 2021, 6, 423. [Google Scholar] [CrossRef]
  73. Roskoski, R. Classification of small molecule protein kinase inhibitors based upon the structures of their drug-enzyme complexes. Pharmacol. Res. 2016, 103, 26–48. [Google Scholar] [CrossRef]
  74. Ferguson, F.M.; Gray, N.S. Kinase inhibitors: The road ahead. Nat. Rev. Drug Discov. 2018, 17, 353–377. [Google Scholar] [CrossRef] [PubMed]
  75. Kung, J.E.; Jura, N. Prospects for pharmacological targeting of pseudokinases. Nat. Rev. Drug Discov. 2019, 18, 501–526. [Google Scholar] [CrossRef] [PubMed]
  76. Roskoski, R. Properties of FDA-approved small molecule protein kinase inhibitors: A 2021 update. Pharmacol. Res. 2021, 165, 105463. [Google Scholar] [CrossRef] [PubMed]
  77. Szilveszter, K.P.; Németh, T.; Mócsai, A. Tyrosine Kinases in Autoimmune and Inflammatory Skin Diseases. Front. Immunol. 2019, 10, 1862. [Google Scholar] [CrossRef] [PubMed]
  78. Zarrin, A.A.; Bao, K.; Lupardus, P.; Vucic, D. Kinase inhibition in autoimmunity and inflammation. Nat. Rev. Drug Discov. 2021, 20, 39–63. [Google Scholar] [CrossRef] [PubMed]
  79. Yang, J.; Nie, J.; Ma, X.; Wei, Y.; Peng, Y.; Wei, X. Targeting PI3K in cancer: Mechanisms and advances in clinical trials. Mol. Cancer 2019, 18, 1–28. [Google Scholar] [CrossRef] [PubMed]
  80. Klaeger, S.; Heinzlmeir, S.; Wilhelm, M.; Polzer, H.; Vick, B.; Koenig, P.-A.; Reinecke, M.; Ruprecht, B.; Petzoldt, S.; Meng, C.; et al. The target landscape of clinical kinase drugs. Science 2017, 358, eaan4368. [Google Scholar] [CrossRef]
  81. Ayala-Aguilera, C.C.; Valero, T.; Lorente-Macías, Á.; Baillache, D.J.; Croke, S.; Unciti-Broceta, A. Small Molecule Kinase Inhibitor Drugs (1995–2021): Medical Indication, Pharmacology, and Synthesis. J. Med. Chem. 2022, 65, 1047–1131. [Google Scholar] [CrossRef]
  82. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  83. Bryant, P.; Pozzati, G.; Elofsson, A. Improved prediction of protein-protein interactions using AlphaFold2. Nat. Commun. 2022, 13, 1265. [Google Scholar] [CrossRef]
  84. Gould, C.M.; Newton, A.C. The life and death of protein kinase C. Curr. Drug Targets 2008, 9, 614–625. [Google Scholar] [CrossRef] [PubMed]
  85. Levin, D.E.; Fields, F.; Kunisawa, R.; Bishop, J.; Thorner, J. A candidate protein kinase C gene, PKC1, is required for the S. cerevisiae cell cycle. Cell 1990, 62, 213–224. [Google Scholar] [CrossRef] [PubMed]
  86. Watanabe, M.; Chen, C.; Levin, D. Saccharomyces cerevisiae PKC1 encodes a protein kinase C (PKC) homolog with a substrate specificity similar to that of mammalian PKC. J. Biol. Chem. 1994, 269, 16829–16836. [Google Scholar] [CrossRef] [PubMed]
  87. Giorgione, J.R.; Lin, J.H.; McCammon, J.A.; Newton, A.C. Increased membrane affinity of the C1 domain of protein kinase Cdelta compensates for the lack of involvement of its C2 domain in membrane recruitment. J. Biol. Chem. 2006, 281, 1660–1669. [Google Scholar] [CrossRef] [PubMed]
  88. Dries, D.R.; Gallegos, L.L.; Newton, A.C. A single residue in the C1 domain sensitizes novel protein kinase C isoforms to cellular diacylglycerol production. J. Biol. Chem. 2007, 282, 826–830. [Google Scholar] [CrossRef] [PubMed]
  89. Newton, A.C. Protein kinase C: Perfectly balanced. Crit. Rev. Biochem. Mol. Biol. 2018, 53, 208–230. [Google Scholar] [CrossRef] [PubMed]
  90. eranen, L.M.; Dutil, E.M.; Newton, A.C. Protein kinase C is regulated in vivo by three functionally distinct phosphorylations. Curr. Biol. 1995, 5, 1394–1403. [Google Scholar] [CrossRef]
  91. Tsutakawa, S.E.; Medzihradszky, K.F.; Flint, A.J.; Burlingame, A.L.; Koshland, D.E., Jr. Determination of in vivo phosphorylation sites in protein kinase C. J. Biol. Chem. 1995, 270, 26807–26812. [Google Scholar] [CrossRef]
  92. Gould, C.M.; Kannan, N.; Taylor, S.S.; Newton, A.C. The chaperones Hsp90 and Cdc37 mediate the maturation and stabilization of protein kinase C through a conserved PXXP motif in the C-terminal tail. J. Biol. Chem. 2009, 284, 4921–4935. [Google Scholar] [CrossRef]
  93. Guertin, D.A.; Stevens, D.M.; Thoreen, C.C.; Burds, A.A.; Kalaany, N.Y.; Moffat, J.; Brown, M.; Fitzgerald, K.J.; Sabatini, D.M. Ablation in mice of the mTORC components raptor, rictor, or mLST8 reveals that mTORC2 is required for signaling to Akt-FOXO and PKCα, but not S6K1. Dev. Cell 2006, 11, 859–871. [Google Scholar] [CrossRef]
  94. Baffi, T.R.; Lordén, G.; Wozniak, J.M.; Feichtner, A.; Yeung, W.; Kornev, A.P.; King, C.C.; Del Rio, J.C.; Limaye, A.J.; Bogomolovas, J.; et al. mTORC2 controls the activity of PKC and Akt by phosphorylating a conserved TOR interaction motif. Sci. Signal 2021, 14, eabe4509. [Google Scholar] [CrossRef] [PubMed]
  95. Newton, A.C. Protein kinase C as a tumor suppressor. Semin. Cancer Biol. 2018, 48, 18–26. [Google Scholar] [CrossRef] [PubMed]
  96. Balendran, A.; Hare, G.R.; Kieloch, A.; Williams, M.R.; Alessi, D.R. Further evidence that 3-phosphoinositide-dependent protein kinase-1 (PDK1) is required for the stability and phosphorylation of protein kinase C (PKC) isoforms. FEBS Lett. 2000, 484, 217–223. [Google Scholar] [CrossRef] [PubMed]
  97. Cameron, A.J.M.; Escribano, C.; Saurin, A.T.; Kostelecky, B.; Parker, P.J. PKC maturation is promoted by nucleotide pocket occupation independently of intrinsic kinase activity. Nat. Struct. Mol. Biol. 2009, 16, 624–630. [Google Scholar] [CrossRef] [PubMed]
  98. Newton, A.C. Regulation of the ABC kinases by phosphorylation: Protein kinase C as a paradigm. Biochem. J. 2003, 370 Pt 2, 361–371. [Google Scholar] [CrossRef]
  99. Newton, A. Regulation of Conventional and Novel Protein Kinase C Isozymes by Phosphorylation and Lipids. In Protein Kinase C in Cancer Signaling and Therapy; Kazanietz, M.G., Ed.; Humana Press: Totowa, NJ, USA, 2010; pp. 9–23. [Google Scholar]
  100. Freeley, M.; Kelleher, D.; Long, A. Regulation of Protein Kinase C function by phosphorylation on conserved and non-conserved sites. Cell Signal 2011, 23, 753–762. [Google Scholar] [CrossRef] [PubMed]
  101. Singh, R.K.; Kumar, S.; Tomar, M.S.; Verma, P.K.; Kumar, A.; Kumar, S.; Kumar, N.; Singh, J.P.; Acharya, A. Putative role of natural products as Protein Kinase C modulator in different disease conditions. Daru 2021, 29, 397–414. [Google Scholar] [CrossRef]
  102. Hornbeck, P.V.; Kornhauser, J.M.; Tkachev, S.; Zhang, B.; Skrzypek, E.; Murray, B.; Latham, V.; Sullivan, M. PhosphoSitePlus: A comprehensive resource for investigating the structure and function of experimentally determined post-translational modifications in man and mouse. Nucleic Acids Res. 2012, 40, D261–D270. [Google Scholar] [CrossRef]
  103. Wang, Y.; Wang, Y.; Zhang, H.; Gao, Y.; Huang, C.; Zhou, A.; Zhou, Y.; Li, Y. Sequential posttranslational modifications regulate PKC degradation. Mol. Biol. Cell 2016, 27, 410–420. [Google Scholar] [CrossRef]
  104. Drummond, M.L.; Prehoda, K.E. Molecular Control of Atypical Protein Kinase C: Tipping the Balance between Self-Renewal and Differentiation. J. Mol. Biol. 2016, 428, 1455–1464. [Google Scholar] [CrossRef]
  105. Violin, J.D.; Zhang, J.; Tsien, R.Y.; Newton, A.C. A genetically encoded fluorescent reporter reveals oscillatory phosphorylation by protein kinase C. J. Cell Biol. 2003, 161, 899–909. [Google Scholar] [CrossRef] [PubMed]
  106. Bartlett, P.J.; Young, K.W.; Nahorski, S.R.; Challiss, R.A.J. Single cell analysis and temporal profiling of agonist-mediated inositol 1, 4, 5-trisphosphate, Ca2+, diacylglycerol, and protein kinase C signaling using fluorescent biosensors. J. Biol. Chem. 2005, 280, 21837–21846. [Google Scholar] [CrossRef] [PubMed]
  107. Uchino, M.; Sakai, N.; Kashiwagi, K.; Shirai, Y.; Shinohara, Y.; Hirose, K.; Iino, M.; Yamamura, T.; Saito, N. Isoform-specific phosphorylation of metabotropic glutamate receptor 5 by protein kinase C (PKC) blocks Ca2+ oscillation and oscillatory translocation of Ca2+-dependent PKC. J. Biol. Chem. 2004, 279, 2254–2261. [Google Scholar] [CrossRef] [PubMed]
  108. Gallegos, L.L.; Kunkel, M.T.; Newton, A.C. Targeting protein kinase C activity reporter to discrete intracellular regions reveals spatiotemporal differences in agonist-dependent signaling. J. Biol. Chem. 2006, 281, 30947–30956. [Google Scholar] [CrossRef] [PubMed]
  109. Marín-Vicente, C.; Nicolás, F.E.; Gómez-Fernández, J.C.; Corbalán-García, S. The PtdIns (4, 5) P2 ligand itself influences the localization of PKCα in the plasma membrane of intact living cells. J. Mol. Biol. 2008, 377, 1038–1052. [Google Scholar] [CrossRef] [PubMed]
  110. Carrasco, S.; Merida, I. Diacylglycerol-dependent binding recruits PKCθ and RasGRP1 C1 domains to specific subcellular localizations in living T lymphocytes. Mol. Biol. Cell 2004, 15, 2932–2942. [Google Scholar] [CrossRef]
  111. Rahman, K. Studies on free radicals, antioxidants, and co-factors. Clin. Interv. Aging 2007, 2, 219–236. [Google Scholar]
  112. Antal, C.E.; Newton, A.C. Tuning the signalling output of protein kinase C. Biochem. Soc. Trans. 2014, 42, 1477–1483. [Google Scholar] [CrossRef]
  113. Konishi, H.; Tanaka, M.; Takemura, Y.; Matsuzaki, H.; Ono, Y.; Kikkawa, U.; Nishizuka, Y. Activation of protein kinase C by tyrosine phosphorylation in response to H2O2. Proc. Natl. Acad. Sci. USA 1997, 94, 11233–11237. [Google Scholar] [CrossRef]
  114. Konishi, H.; Yamauchi, E.; Taniguchi, H.; Yamamoto, T.; Matsuzaki, H.; Takemura, Y.; Ohmae, K.; Kikkawa, U.; Nishizuka, Y. Phosphorylation sites of protein kinase C delta in H2O2-treated cells and its activation by tyrosine kinase in vitro. Proc. Natl. Acad. Sci. USA 2001, 98, 6587–6592. [Google Scholar] [CrossRef]
  115. Adwan, T.S.; Ohm, A.M.; Jones, D.N.M.; Humphries, M.J.; Reyland, M.E. Regulated binding of importin-alpha to protein kinase Cdelta in response to apoptotic signals facilitates nuclear import. J. Biol. Chem. 2011, 286, 35716–35724. [Google Scholar] [CrossRef]
  116. Kaul, S.; Anantharam, V.; Yang, Y.; Choi, C.J.; Kanthasamy, A.; Kanthasamy, A.G. Tyrosine phosphorylation regulates the proteolytic activation of protein kinase Cdelta in dopaminergic neuronal cells. J. Biol. Chem. 2005, 280, 28721–28730. [Google Scholar] [CrossRef] [PubMed]
  117. Langeberg, L.K.; Scott, J.D. Signalling scaffolds and local organization of cellular behaviour. Nat. Rev. Mol. Cell Biol. 2015, 16, 232–244. [Google Scholar] [CrossRef] [PubMed]
  118. Finger, E.C.; Castellini, L.; Rankin, E.B.; Vilalta, M.; Krieg, A.J.; Jiang, D.; Banh, A.; Zundel, W.; Powell, M.B.; Giaccia, A.J. Hypoxic induction of AKAP12 variant 2 shifts PKA-mediated protein phosphorylation to enhance migration and metastasis of melanoma cells. Proc. Natl. Acad. Sci. USA 2015, 112, 4441–4446. [Google Scholar] [CrossRef] [PubMed]
  119. Hirsch, A.H.; Glantz, S.B.; Li, Y.I.N.G.; You, Y.; Rubin, C.S. Cloning and expression of an intron-less gene for AKAP 75, an anchor protein for the regulatory subunit of cAMP-dependent protein kinase II beta. J. Biol. Chem. 1992, 267, 2131–2134. [Google Scholar] [CrossRef] [PubMed]
  120. Wong, W.; Scott, J.D. AKAP signalling complexes: Focal points in space and time. Nat. Rev. Mol. Cell Biol. 2004, 5, 959–970. [Google Scholar] [CrossRef]
  121. Marin, W. A-kinase anchoring protein 1 (AKAP1) and its role in some cardiovascular diseases. J. Mol. Cell. Cardiol. 2020, 138, 99–109. [Google Scholar] [CrossRef] [PubMed]
  122. Kazanietz, M.G.; Lemmon, M.A. Protein Kinase C Regulation: C1 Meets C-tail. Structure 2011, 19, 144–146. [Google Scholar] [CrossRef]
  123. Scott, J.D.; Newton, A.C. Shedding light on local kinase activation. BMC Biol. 2012, 10, 61. [Google Scholar] [CrossRef]
  124. Mukherjee, A.; Roy, S.; Saha, B.; Mukherjee, D. Spatio-Temporal Regulation of PKC Isoforms Imparts Signaling Specificity. Front. Immunol. 2016, 7, 45. [Google Scholar] [CrossRef]
  125. Mochly-Rosen, D.; Khaner, H.; Lopez, J.; Smith, B. Intracellular receptors for activated protein kinase C. Identification of a binding site for the enzyme. J. Biol. Chem. 1991, 266, 14866–14868. [Google Scholar] [CrossRef] [PubMed]
  126. Mochly-Rosen, D.; Smith, B.L.; Chen, C.H.; Disatnik, M.H.; Ron, D. Interaction of protein kinase C with RACK1, a receptor for activated C-kinase: A role in beta protein kinase C mediated signal transduction. Biochem. Soc. Trans. 1995, 23, 596–600. [Google Scholar] [CrossRef] [PubMed]
  127. Mochly-Rosen, D.; Gordon, A.S. Anchoring proteins for protein kinase C: A means for isozyme selectivity. Faseb J. 1998, 12, 35–42. [Google Scholar] [CrossRef] [PubMed]
  128. Cunningham, A.D.; Qvit, N.; Mochly-Rosen, D. Peptides and peptidomimetics as regulators of protein-protein interactions. Curr. Opin. Struct. Biol. 2017, 44, 59–66. [Google Scholar] [CrossRef]
  129. Obsilova, V.; Obsil, T. The 14-3-3 Proteins as Important Allosteric Regulators of Protein Kinases. Int. J. Mol. Sci. 2020, 21, 8824. [Google Scholar] [CrossRef]
  130. Hoque, M.; Rentero, C.; Cairns, R.; Tebar, F.; Enrich, C.; Grewal, T. Annexins—Scaffolds modulating PKC localization and signaling. Cell Signal 2014, 26, 1213–1225. [Google Scholar] [CrossRef]
  131. Li, K.; Sun, P.; Wang, Y.; Gao, T.; Zheng, D.; Liu, A.; Ni, Y. Hsp90 interacts with Cdc37, is phosphorylated by PKA/PKC, and regulates Src phosphorylation in human sperm capacitation. Andrology 2021, 9, 185–195. [Google Scholar] [CrossRef]
  132. Chen, J.; Wu, M.; Zhang, X.; Zhang, W.; Zhang, Z.; Chen, L.; He, J.; Zheng, Y.; Chen, C.; Wang, F.; et al. Hepatitis B virus polymerase impairs interferon-α–induced STA T activation through inhibition of importin-α5 and protein kinase C-δ. Hepatology 2013, 57, 470–482. [Google Scholar] [CrossRef]
  133. Pidoux, G.; Taskén, K. Specificity and spatial dynamics of protein kinase A signaling organized by A-kinase-anchoring proteins. J. Mol. Endocrinol. 2010, 44, 271–284. [Google Scholar] [CrossRef]
  134. Limaye, A.J.; Bendzunas, G.N.; Kennedy, E.J. Targeted disruption of PKC from AKAP signaling complexes. RSC Chem. Biol. 2021, 2, 1227–1231. [Google Scholar] [CrossRef]
  135. Ron, D.; Chen, C.H.; Caldwell, J.; Jamieson, L.; Orr, E.; Mochly-Rosen, D. Cloning of an intracellular receptor for protein kinase C: A homolog of the beta subunit of G proteins. Proc. Natl. Acad. Sci. USA 1994, 91, 839–843. [Google Scholar] [CrossRef] [PubMed]
  136. Harrison-Lavoie, K.; Lewis, V.; Hynes, G.; Collison, K.; Nutland, E.; Willison, K. A 102 kDa subunit of a Golgi-associated particle has homology to beta subunits of trimeric G proteins. Embo J. 1993, 12, 2847–2853. [Google Scholar] [CrossRef] [PubMed]
  137. Stenbeck, G.; Harter, C.; Brecht, A.; Herrmann, D.; Lottspeich, F.; Orci, L.; Wieland, F. beta’-COP, a novel subunit of coatomer. Embo J. 1993, 12, 2841–2845. [Google Scholar] [CrossRef]
  138. Csukai, M.; Chen, C.H.; De Matteis, M.A.; Mochly-Rosen, D. The coatomer protein beta’-COP, a selective binding protein (RACK) for protein kinase Cepsilon. J. Biol. Chem. 1997, 272, 29200–29206. [Google Scholar] [CrossRef] [PubMed]
  139. Ron, D.; Mochly-Rosen, D. An Autoregulatory Region in Protein-Kinase-C—The Pseudoanchoring Site. Proc. Natl. Acad. Sci. USA 1995, 92, 492–496. [Google Scholar] [CrossRef] [PubMed]
  140. Ron, D.; Luo, J.; Mochly-Rosen, D. C2 region-derived peptides inhibit translocation and function of beta protein kinase C in vivo. J. Biol. Chem. 1995, 270, 24180–24187. [Google Scholar] [CrossRef]
  141. Churchill, E.N.; Qvit, N.; Mochly-Rosen, D. Rationally designed peptide regulators of protein kinase C. Trends Endocrinol. Metab. 2009, 20, 25–33. [Google Scholar]
  142. Kim, J.; Mochly-Rosen, D. Regulation of PKC by Protein-Protein Interactions in Cancer, in Protein Kinase C in Cancer Signaling and Therapy; Humana Press: Totowa, NJ, USA, 2010; pp. 79–103. [Google Scholar]
  143. Palaniyandi, S.S.; Sun, L.; Ferreira, J.C.B.; Mochly-Rosen, D. Protein kinase C in heart failure: A therapeutic target? Cardiovasc. Res. 2009, 82, 229–239. [Google Scholar] [CrossRef]
  144. Qvit, N.; Mochly-Rosen, D. Highly specific modulators of protein kinase C localization: Applications to heart failure. Drug Discov. Today: Dis. Mech. 2010, 7, e87–e93. [Google Scholar] [CrossRef]
  145. Tsunoda, S.; Sierralta, J.; Sun, Y.; Bodner, R.; Suzuki, E.; Becker, A.; Socolich, M.; Zuker, C.S. A multivalent PDZ-domain protein assembles signalling complexes in a G-protein-coupled cascade. Nature 1997, 388, 243–249. [Google Scholar] [CrossRef]
  146. Zhang, J.; Yang, P.L.; Gray, N.S. Targeting cancer with small molecule kinase inhibitors. Nat. Rev. Cancer 2009, 9, 28–39. [Google Scholar] [CrossRef] [PubMed]
  147. Bhullar, K.S.; Lagarón, N.O.; McGowan, E.M.; Parmar, I.; Jha, A.; Hubbard, B.P.; Rupasinghe, H.P.V. Kinase-targeted cancer therapies: Progress, challenges and future directions. Mol. Cancer 2018, 17, 48. [Google Scholar] [CrossRef] [PubMed]
  148. Dupont, C.A.; Riegel, K.; Pompaiah, M.; Juhl, H.; Rajalingam, K. Druggable genome and precision medicine in cancer: Current challenges. Febs J. 2021, 288, 6142–6158. [Google Scholar] [CrossRef] [PubMed]
  149. Nishizuka, Y. Intracellular signaling by hydrolysis of phospholipids and activation of protein kinase C. Science 1992, 258, 607–614. [Google Scholar] [CrossRef] [PubMed]
  150. Nishizuka, Y. The role of protein kinase C in cell surface signal transduction and tumour promotion. Nature 1984, 308, 693–698. [Google Scholar] [CrossRef]
  151. Garg, R.; Benedetti, L.G.; Abera, M.B.; Wang, H.; Abba, M.; Kazanietz, M.G. Protein kinase C and cancer: What we know and what we do not. Oncogene 2014, 33, 5225–5237. [Google Scholar] [CrossRef]
  152. Berenblum, I.; Shubik, P. The Role of Croton Oil Applications, Associated with a Single Painting of a Carcinogen, in Tumour Induction of the Mouse’s Skin. Br. J. Cancer 1947, 1, 379–382. [Google Scholar] [CrossRef]
  153. Black, A.R.; Black, J.D. Protein kinase C signaling and cell cycle regulation. Front. Immunol. 2012, 3, 423. [Google Scholar] [CrossRef]
  154. Gao, J.; Aksoy, B.A.; Dogrusoz, U.; Dresdner, G.; Gross, B.E.; Sumer, S.O.; Sun, Y.; Jacobsen, A.; Sinha, R.; Larsson, E.; et al. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci. Signal 2013, 6, pl1. [Google Scholar] [CrossRef]
  155. Antal, C.E.; Hudson, A.M.; Kang, E.; Zanca, C.; Wirth, C.; Stephenson, N.L.; Trotter, E.W.; Gallegos, L.L.; Miller, C.J.; Furnari, F.B.; et al. Cancer-associated protein kinase C mutations reveal kinase’s role as tumor suppressor. Cell 2015, 160, 489–502. [Google Scholar] [CrossRef]
  156. Parker, P.J.; Brown, S.J.; Calleja, V.; Chakravarty, P.; Cobbaut, M.; Linch, M.; Marshall, J.J.T.; Martini, S.; McDonald, N.Q.; Soliman, T.; et al. Equivocal, explicit and emergent actions of PKC isoforms in cancer. Nat. Rev. Cancer 2021, 21, 51–63. [Google Scholar] [CrossRef] [PubMed]
  157. Baudot, A.D.; Jeandel, P.Y.; Mouska, X.; Maurer, U.; Tartare-Deckert, S.; Raynaud, S.D.; Cassuto, J.P.; Ticchioni, M.; Deckert, M. The tyrosine kinase Syk regulates the survival of chronic lymphocytic leukemia B cells through PKCδ and proteasome-dependent regulation of Mcl-1 expression. Oncogene 2009, 28, 3261–3273. [Google Scholar] [CrossRef] [PubMed]
  158. Ruvolo, P.P.; Qiu, Y.; Coombes, K.R.; Zhang, N.; Neeley, E.S.; Ruvolo, V.R.; Hail, N.; Borthakur, G.; Konopleva, M.; Andreeff, M.; et al. Phosphorylation of GSK3α/β correlates with activation of AKT and is prognostic for poor overall survival in acute myeloid leukemia patients. BBA Clin. 2015, 4, 59–68. [Google Scholar] [CrossRef]
  159. Black, A.R.; Black, J.D. The complexities of PKCα signaling in cancer. Adv. Biol. Regul. 2021, 80, 100769. [Google Scholar] [CrossRef] [PubMed]
  160. Kaur, N.; Lum, M.A.; Lewis, R.E.; Black, A.R.; Black, J.D. A novel antiproliferative PKCα-Ras-ERK signaling axis in intestinal epithelial cells. J. Biol. Chem. 2022, 298, 102121. [Google Scholar] [CrossRef] [PubMed]
  161. Prévostel, C.; Alvaro, V.; De Boisvilliers, F.; Martin, A.; Jaffiol, C.; Joubert, D. The natural protein kinase C alpha mutant is present in human thyroid neoplasms. Oncogene 1995, 11, 669–674. [Google Scholar] [PubMed]
  162. Prevostel, C.; Alice, V.; Joubert, D.; Parker, P.J. Protein kinase C(alpha) actively downregulates through caveolae-dependent traffic to an endosomal compartment. J. Cell Sci. 2000, 113, 2575–2584. [Google Scholar] [CrossRef] [PubMed]
  163. Lahn, M.M.; Sundell, K.L. The role of protein kinase C-alpha (PKC-alpha) in melanoma. Melanoma Res. 2004, 14, 85–89. [Google Scholar] [CrossRef]
  164. Palazzo, E.; Kellett, M.D.; Cataisson, C.; Bible, P.W.; Bhattacharya, S.; Sun, H.-W.; Gormley, A.C.; Yuspa, S.H.; Morasso, M.I. A novel DLX3–PKC integrated signaling network drives keratinocyte differentiation. Cell Death Differ. 2017, 24, 717–730. [Google Scholar] [CrossRef]
  165. Isakov, N. Protein kinase C (PKC) isoforms in cancer, tumor promotion and tumor suppression. In Seminars in Cancer Biology; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar]
  166. Urtreger, A.J.; Kazanietz, M.G.; Bal de Kier Joffé, E.D. Contribution of individual PKC isoforms to breast cancer progression. IUBMB Life 2012, 64, 18–26. [Google Scholar] [CrossRef]
  167. Sadeghi, M.M.; Salama, M.F.; Hannun, Y.A. Protein Kinase C as a Therapeutic Target in Non-Small Cell Lung Cancer. Int. J. Mol. Sci. 2021, 22, 5527. [Google Scholar] [CrossRef] [PubMed]
  168. 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]
  169. 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 C beta II suppresses colorectal cancer by regulating IGF-1 mediated cell survival. Oncotarget 2016, 7, 20919–20933. [Google Scholar] [CrossRef] [PubMed]
  170. Engers, R.; Mrzyk, S.; Springer, E.; Fabbro, D.; Weissgerber, G.; Gerharz, 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] [PubMed]
  171. Li, J.; Gobe, G. Protein kinase C activation and its role in kidney disease. Nephrology 2006, 11, 428–434. [Google Scholar] [CrossRef] [PubMed]
  172. 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] [PubMed]
  173. 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] [PubMed]
  174. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Ahmedin, J.D.V.M. Cancer statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  175. Kamat, A.M.; Hahn, N.M.; Efstathiou, J.A.; Lerner, S.P.; Malmström, P.U.; Choi, W.; Guo, C.C.; Lotan, Y.; Kassouf, W. Bladder cancer. Lancet 2016, 388, 2796–2810. [Google Scholar] [CrossRef]
  176. Sanli, O.; Dobruch, J.; Knowles, M.A.; Burger, M.; Alemozaffar, M.; Nielsen, M.E.; Lotan, Y. Bladder cancer. Nat. Rev. Dis. Primers 2017, 3, 17022. [Google Scholar] [CrossRef]
  177. 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] [CrossRef] [PubMed]
  178. 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] [PubMed]
  179. 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] [CrossRef] [PubMed]
  180. Thandra, K.C.; Barsouk, A.; Saginala, K.; Padala, S.A.; Barsouk, A.; Rawla, P. Epidemiology of Non-Hodgkin’s Lymphoma. Med. Sci. 2021, 9, 5. [Google Scholar] [CrossRef] [PubMed]
  181. Cai, W.; Zeng, Q.; Zhang, X.; Ruan, W. Trends Analysis of Non-Hodgkin Lymphoma at the National, Regional, and Global Level, 1990–2019: Results from the Global Burden of Disease Study 2019. Front. Med. 2021, 8, 738693. [Google Scholar] [CrossRef] [PubMed]
  182. Su, T.T.; Guo, B.; Rawlings, D.J. Emerging roles for PKC isoforms in immune cell function. Mol. Interv. 2002, 2, 141–144. [Google Scholar] [CrossRef] [PubMed]
  183. McMahon, G. VEGF receptor signaling in tumor angiogenesis. Oncologist 2000, 5, 3–10. [Google Scholar] [CrossRef]
  184. Giles, F.J.; Vose, J.M.; Do, K.-A.; Johnson, M.M.; Manshouri, T.; Bociek, G.; Bierman, P.J.; O’Brien, S.M.; Kantarjian, H.M.; Armitage, J.O.; et al. Clinical relevance of circulating angiogenic factors in patients with non-Hodgkin’s lymphoma or Hodgkin’s lymphoma. Leuk. Res. 2004, 28, 595–604. [Google Scholar] [CrossRef]
  185. 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-beta or cyclin D2 predicts for inferior survival in diffuse large B-cell lymphoma. Mod. Pathol. 2005, 18, 1377–1384. [Google Scholar] [CrossRef]
  186. Berditchevski, F.; Fennell, E.; Murray, P.G. Calcium-dependent signalling in B-cell lymphomas. Oncogene 2021, 40, 6321–6328. [Google Scholar] [CrossRef]
  187. Cabanillas, M.E.; McFadden, D.G.; Durante, C. Thyroid cancer. Lancet 2016, 388, 2783–2795. [Google Scholar] [CrossRef] [PubMed]
  188. Rossi, E.D.; Pantanowitz, L.; Hornick, J.L. A worldwide journey of thyroid cancer incidence centred on tumour histology. Lancet Diabetes Endocrinol. 2021, 9, 193–194. [Google Scholar] [CrossRef] [PubMed]
  189. Prévostel, C.; Martin, A.; Alvaro, V.; Jaffiol, C.; Joubert, D. Protein kinase C alpha and tumorigenesis of the endocrine gland. Horm. Res. 1997, 47, 140–144. [Google Scholar] [PubMed]
  190. Alvaro, V.; Prévostel, C.; Joubert, D.; Slosberg, E.; Weinstein, B.I. Ectopic expression of a mutant form of PKCalpha originally found in human tumors: Aberrant subcellular translocation and effects on growth control. Oncogene 1997, 14, 677–685. [Google Scholar] [CrossRef] [PubMed]
  191. Chen, X.N.; Knauf, J.A.; Gonsky, R.; Wang, M.; Lai, E.H.; Chissoe, S.; Fagin, J.A.; Korenberg, J.R. From amplification to gene in thyroid cancer: A high-resolution mapped bacterial-artificial-chromosome resource for cancer chromosome aberrations guides gene discovery after comparative genome hybridization. Am. J. Hum. Genet. 1998, 63, 625–637. [Google Scholar] [CrossRef] [PubMed]
  192. 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] [PubMed]
  193. Zhang, S.; Gong, T.T.; Liu, F.H.; Jiang, Y.T.; Sun, H.; Ma, X.X.; Zhao, Y.H.; Wu, Q.J. Global, Regional, and National Burden of Endometrial Cancer, 1990–2017: Results from the Global Burden of Disease Study, 2017. Front. Oncol. 2019, 9, 1440. [Google Scholar] [CrossRef]
  194. Haughian, J.M.; Reno, E.M.; Thorne, A.M.; Bradford, A.P. Protein kinase C alpha-dependent signaling mediates endometrial cancer cell growth and tumorigenesis. Int. J. Cancer 2009, 125, 2556–2564. [Google Scholar] [CrossRef]
  195. Reno, E.M.; Haughian, J.M.; Dimitrova, I.K.; Jackson, T.A.; Shroyer, K.R.; Bradford, A.P. Analysis of protein kinase C delta (PKC delta) expression in endometrial tumors. Human. Pathol. 2008, 39, 21–29. [Google Scholar] [CrossRef]
  196. Koo, K.-H.; Jeong, W.-J.; Cho, Y.-H.; Park, J.-C.; Min, D.S.; Choi, K.-Y. K-Ras stabilization by estrogen via PKCδ is involved in endometrial tumorigenesis. Oncotarget 2015, 6, 21328–21340. [Google Scholar] [CrossRef]
  197. Kamisawa, T.; Wood, L.D.; Itoi, T.; Takaori, K. Pancreatic cancer. Lancet 2016, 388, 73–85. [Google Scholar] [CrossRef] [PubMed]
  198. Rawla, P.; Sunkara, T.; Gaduputi, V. Epidemiology of Pancreatic Cancer: Global Trends, Etiology and Risk Factors. World J. Oncol. 2019, 10, 10–27. [Google Scholar] [CrossRef] [PubMed]
  199. Hu, J.-X.; Zhao, C.-F.; Chen, W.-B.; Liu, Q.-C.; Li, Q.-W.; Lin, Y.-Y.; Gao, F. Pancreatic cancer: A review of epidemiology, trend, and risk factors. World J. Gastroenterol. 2021, 27, 4298–4321. [Google Scholar] [CrossRef] [PubMed]
  200. Evans, J.D.; Cornford, P.A.; Dodson, A.; Neoptolemos, J.P.; Foster, C.S. Expression Patterns of Protein Kinase C Isoenzymes Are Characteristically Modulated in Chronic Pancreatitis and Pancreatic Cancer. Am. J. Clin. Pathol. 2003, 119, 392–402. [Google Scholar] [CrossRef]
  201. Storz, P. Targeting protein kinase C subtypes in pancreatic cancer. Expert. Rev. Anticancer. Ther. 2015, 15, 433–438. [Google Scholar] [CrossRef]
  202. Scotti, M.L.; Bamlet, W.R.; Smyrk, T.C.; Fields, A.P.; Murray, N.R. Protein kinase Ciota is required for pancreatic cancer cell transformed growth and tumorigenesis. Cancer Res. 2010, 70, 2064–2074. [Google Scholar] [CrossRef]
  203. Mauro, L.V.; Grossoni, V.C.; Urtreger, A.J.; Yang, C.; Colombo, L.L.; Morandi, A.; Pallotta, M.G.; Kazanietz, M.G.; de Kier Joffé, E.D.B.; Puricelli, L.L. PKCdelta promotes tumoral progression of human ductal pancreatic cancer. Pancreas 2010, 39, e31–e41. [Google Scholar] [CrossRef]
  204. Dowling, C.M.; Hayes, S.L.; Phelan, J.J.; Cathcart, M.C.; Finn, S.P.; Mehigan, B.; McCormick, P.; Coffey, J.C.; O’sullivan, J.; Kiely, P.A. Expression of protein kinase C gamma promotes cell migration in colon cancer. Oncotarget 2017, 8, 72096–72107. [Google Scholar] [CrossRef]
  205. Zhang, L.L.; Cao, F.F.; Wang, Y.; Meng, F.L.; Zhang, Y.; Zhong, D.S.; Zhou, Q.H. The protein kinase C (PKC) inhibitors combined with chemotherapy in the treatment of advanced non-small cell lung cancer: Meta-analysis of randomized controlled trials. Clin. Transl. Oncol. 2015, 17, 371–377. [Google Scholar] [CrossRef]
  206. Soares, A.C.; Fonseca, D.A. Cardiovascular diseases: A therapeutic perspective around the clock. Drug Discov. Today 2020, 25, 1086–1098. [Google Scholar] [CrossRef]
  207. Virani, S.S.; Alonso, A.; Aparicio, H.J.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Cheng, S.; Delling, F.N.; et al. Heart disease and stroke statistics—2021 update: A report from the American Heart Association. Circulation 2021, 143, e254–e743. [Google Scholar] [CrossRef] [PubMed]
  208. Kumar, R.; Singh, V.P.; Baker, K.M. Kinase inhibitors for cardiovascular disease. J. Mol. Cell. Cardiol. 2007, 42, 1–11. [Google Scholar] [CrossRef] [PubMed]
  209. Fuller, S.J.; Osborne, S.A.; Leonard, S.J.; Hardyman, M.A.; Vaniotis, G.; Allen, B.G.; Sugden, P.H.; Clerk, A. Cardiac protein kinases: The cardiomyocyte kinome and differential kinase expression in human failing hearts. Cardiovasc. Res. 2015, 108, 87–98. [Google Scholar] [CrossRef] [PubMed]
  210. Chen, J.; Li, Y.; Du, C.; Wei, T.; Shan, T.; Wang, L. Protein kinases in cardiovascular diseases. Chin. Med. J. 2022, 135, 557–570. [Google Scholar] [CrossRef] [PubMed]
  211. Churchill, E.; Budas, G.; Vallentin, A.; Koyanagi, T.; Mochly-Rosen, D. PKC isozymes in chronic cardiac disease: Possible therapeutic targets? Annu. Rev. Pharmacol. Toxicol. 2008, 48, 569–599. [Google Scholar] [CrossRef] [PubMed]
  212. Steinberg, S.F. Cardiac actions of protein kinase C isoforms. Physiology 2012, 27, 130–139. [Google Scholar] [CrossRef]
  213. Pucéat, M.; Hilal-Dandan, R.; Strulovici, B.; Brunton, L.; Brown, J. Differential regulation of protein kinase C isoforms in isolated neonatal and adult rat cardiomyocytes. J. Biol. Chem. 1994, 269, 16938–16944. [Google Scholar] [CrossRef] [PubMed]
  214. Rybin, V.O.; Steinberg, S.F. Protein kinase C isoform expression and regulation in the developing rat heart. Circ. Res. 1994, 74, 299–309. [Google Scholar] [CrossRef]
  215. Bowling, N.; Walsh, R.A.; Song, G.; Estridge, T.; Sandusky, G.E.; Fouts, R.L.; Mintze, K.; Pickard, T.; Roden, R.; Bristow, M.R.; et al. Increased protein kinase C activity and expression of Ca2+-sensitive isoforms in the failing human heart. Circulation 1999, 99, 384–391. [Google Scholar] [CrossRef]
  216. Simonis, G.; Briem, S.K.; Schoen, S.P.; Bock, M.; Marquetant, R.; Strasser, R.H. Protein kinase C in the human heart: Differential regulation of the isoforms in aortic stenosis or dilated cardiomyopathy. Mol. Cell. Biochem. 2007, 305, 103–111. [Google Scholar] [CrossRef]
  217. Singh, R.M.; Cummings, E.; Pantos, C.; Singh, J. Protein kinase C and cardiac dysfunction: A review. Heart Fail. Rev. 2017, 22, 843–859. [Google Scholar] [CrossRef]
  218. Weeks, K.L.; McMullen, J.R. Divergent Effects of PKC (Protein Kinase C) α in the Human and Animal Heart? Therapeutic Implications for PKC Inhibitors in Cardiac Patients. Circ. Genom. Precis. Med. 2018, 11, e002104. [Google Scholar] [CrossRef] [PubMed]
  219. Wetsel, W.; Khan, W.; Merchenthaler, I.; Rivera, H.; Halpern, A.; Phung, H.; Negro-Vilar, A.; Hannun, Y. Tissue and cellular distribution of the extended family of protein kinase C isoenzymes. J. Cell Biol. 1992, 117, 121–133. [Google Scholar] [CrossRef] [PubMed]
  220. Bogoyevitch, M.A.; Parker, P.J.; Sugden, P.H. Characterization of protein kinase C isotype expression in adult rat heart. Protein kinase C-epsilon is a major isotype present, and it is activated by phorbol esters, epinephrine, and endothelin. Circ. Res. 1993, 72, 757–767. [Google Scholar] [CrossRef]
  221. Talosi, L.; Kranias, E.G. Effect of alpha-adrenergic stimulation on activation of protein kinase C and phosphorylation of proteins in intact rabbit hearts. Circ. Res. 1992, 70, 670–678. [Google Scholar] [CrossRef] [PubMed]
  222. Qiu, Y.; Ping, P.; Tang, X.L.; Manchikalapudi, S.; Rizvi, A.; Zhang, J.; Takano, H.; Wu, W.J.; Teschner, S.; Bolli, R. Direct evidence that protein kinase C plays an essential role in the development of late preconditioning against myocardial stunning in conscious rabbits and that epsilon is the isoform involved. J. Clin. Investig. 1998, 101, 2182–2198. [Google Scholar] [CrossRef] [PubMed]
  223. Karamchand, P.; Ball, N.A.; Dorn, G.W.; Walsh, R.A. Left ventricular stretch stimulates angiotensin II--mediated phosphatidylinositol hydrolysis and protein kinase C epsilon isoform translocation in adult guinea pig hearts. Circ. Res. 1997, 81, 643–650. [Google Scholar]
  224. Cai, J.J.; Lee, H.C. Protein kinase C isozyme-specific modulation of cyclic AMP-dependent phosphodiesterase in hypertrophic cardiomyopathic hamster hearts. Mol. Pharmacol. 1996, 49, 81–88. [Google Scholar]
  225. Domenech, R.J.; Macho, P.; Vélez, D.; Sánchez, G.; Liu, X.; Dhalla, N. Tachycardia preconditions infarct size in dogs: Role of adenosine and protein kinase C. Circulation 1998, 97, 786–794. [Google Scholar] [CrossRef]
  226. Goldberg, M.; Steinberg, S.F. Tissue-specific developmental regulation of protein kinase C isoforms. Biochem. Pharmacol. 1996, 51, 1089–1093. [Google Scholar] [CrossRef]
  227. Chen, C.-H.; Budas, G.R.; Churchill, E.N.; Disatnik, M.-H.; Hurley, T.D.; Mochly-Rosen, D. Activation of aldehyde dehydrogenase-2 reduces ischemic damage to the heart. Science 2008, 321, 1493–1495. [Google Scholar] [CrossRef] [PubMed]
  228. Ferreira, J.C.B.; Mochly-Rosen, D.; Boutjdir, M. Regulation of cardiac excitability by protein kinase C isozymes. Front. Biosci. (Sch. Ed.) 2012, 4, 532–546. [Google Scholar] [CrossRef] [PubMed]
  229. Churchill, E.N.; Mochly-Rosen, D. The roles of PKCdelta and epsilon isoenzymes in the regulation of myocardial ischaemia/reperfusion injury. Biochem. Soc. Trans. 2007, 35 Pt 5, 1040–1042. [Google Scholar] [CrossRef] [PubMed]
  230. Qvit, N.; Disatnik, M.-H.; Sho, E.; Mochly-Rosen, D. Selective phosphorylation inhibitor of delta protein kinase C-pyruvate dehydrogenase kinase protein-protein interactions: Application for myocardial injury in vivo. J. Am. Chem. Soc. 2016, 138, 7626–7635. [Google Scholar] [CrossRef]
  231. Miao, L.-N.; Pan, D.; Shi, J.; Du, J.-P.; Chen, P.-F.; Gao, J.; Yu, Y.; Shi, D.-Z.; Guo, M. Role and Mechanism of PKC-δ for Cardiovascular Disease: Current Status and Perspective. Front. Cardiovasc. Med. 2022, 9, 816369. [Google Scholar] [CrossRef] [PubMed]
  232. Paumelle, R.; Blanquart, C.; Briand, O.; Barbier, O.; Duhem, C.; Woerly, G.; Percevault, F.; Fruchart, J.C.; Dombrowicz, D.; Glineur, C.; et al. Acute antiinflammatory properties of statins involve peroxisome proliferator-activated receptor-alpha via inhibition of the protein kinase C signaling pathway. Circ. Res. 2006, 98, 361–369. [Google Scholar] [CrossRef] [PubMed]
  233. Ruboxistaurin: LY 333531. Drugs R D 2007, 8, 193–199. [CrossRef]
  234. Wei, L.; Yin, Z.; Yuan, Y.; Hwang, A.; Lee, A.; Sun, D.; Li, F.; Di, C.; Zhang, R.; Cao, F.; et al. A PKC-β inhibitor treatment reverses cardiac microvascular barrier dysfunction in diabetic rats. Microvasc. Res. 2010, 80, 158–165. [Google Scholar] [CrossRef]
  235. Connelly, K.A.; Kelly, D.J.; Zhang, Y.; Prior, D.L.; Advani, A.; Cox, A.J.; Thai, K.; Krum, H.; Gilbert, R.E. Inhibition of Protein Kinase C–β by Ruboxistaurin Preserves Cardiac Function and Reduces Extracellular Matrix Production in Diabetic Cardiomyopathy. Circ. Heart Fail. 2009, 2, 129–137. [Google Scholar] [CrossRef]
  236. Ladage, D.; Tilemann, L.; Ishikawa, K.; Correll, R.N.; Kawase, Y.; Houser, S.R.; Molkentin, J.D.; Hajjar, R.J. Inhibition of PKCα/β with ruboxistaurin antagonizes heart failure in pigs after myocardial infarction injury. Circ. Res. 2011, 109, 1396–1400. [Google Scholar] [CrossRef]
  237. Zatta, A.J.; Kin, H.; Lee, G.; Wang, N.; Jiang, R.; Lust, R.; Reeves, J.G.; Mykytenko, J.; Guyton, R.A.; Zhao, Z.-Q.; et al. Infarct-sparing effect of myocardial postconditioning is dependent on protein kinase C signalling. Cardiovasc. Res. 2006, 70, 315–324. [Google Scholar] [CrossRef] [PubMed]
  238. Soltoff, S.P. Rottlerin: An inappropriate and ineffective inhibitor of PKCdelta. Trends Pharmacol. Sci. 2007, 28, 453–458. [Google Scholar] [CrossRef] [PubMed]
  239. Fan, H.C.; Fernández-Hernando, C.; Lai, J.H. Protein kinase C isoforms in atherosclerosis: Pro- or anti-inflammatory? Biochem. Pharmacol. 2014, 88, 139–149. [Google Scholar] [CrossRef] [PubMed]
  240. Lien, C.-F.; Chen, S.-J.; Tsai, M.-C.; Lin, C.-S. Potential Role of Protein Kinase C in the Pathophysiology of Diabetes-Associated Atherosclerosis. Front. Pharmacol. 2021, 12, 716332. [Google Scholar] [CrossRef]
  241. Patel, J. Heart failure population health considerations. Am. J. Manag. Care 2021, 27 (Suppl. 9), S191–S195. [Google Scholar]
  242. Vlahos, C.J.; McDowell, S.A.; Clerk, A. Kinases as therapeutic targets for heart failure. Nat. Rev. Drug Discov. 2003, 2, 99–113. [Google Scholar] [CrossRef]
  243. Aslam, N. Increase in PKCα Activity during Heart Failure Despite the Stimulation of PKCα Braking Mechanism. Int. J. Mol. Sci. 2020, 21, 2561. [Google Scholar] [CrossRef]
  244. Liu, Q.; Molkentin, J.D. Protein kinase Cα as a heart failure therapeutic target. J. Mol. Cell Cardiol. 2011, 51, 474–478. [Google Scholar] [CrossRef]
  245. Sheng, J.; Chen, Y.; Chang, H.; Wang, Y.; Jiao, B.; Yu, Z. Multisite phosphorylation of Bcl-2 via protein kinase Cδ facilitates apoptosis of hypertrophic cardiomyocytes. Clin. Exp. Pharmacol. Physiol. 2014, 41, 891–901. [Google Scholar] [CrossRef]
  246. Fryer, L.; Holness, M.; Decock, J.; Sugden, M. Cardiac protein kinase C expression in two models of cardiac hypertrophy associated with an activated cardiac renin-angiotensin system: Effects of experimental hyperthyroidism and genetic hypertension (the mRen-2 rat). J. Endocrinol. 1998, 158, 27–33. [Google Scholar] [CrossRef]
  247. Cain, A.E.; Khalil, R.A. Pathophysiology of essential hypertension: Role of the pump, the vessel, and the kidney. Semin. Nephrol. 2002, 22, 3–16. [Google Scholar] [CrossRef] [PubMed]
  248. Salamanca, D.A.; Khalil, R.A. Protein kinase C isoforms as specific targets for modulation of vascular smooth muscle function in hypertension. Biochem. Pharmacol. 2005, 70, 1537–1547. [Google Scholar] [CrossRef] [PubMed]
  249. Qiao, X.; Khalil, R.A. Role of Protein Kinase C and Related Pathways in Vascular Smooth Muscle Contraction and Hypertension. Neurovascular Med. Pursuing Cell. Longev. Healthy Aging 2008, 21. [Google Scholar] [CrossRef]
  250. Ringvold, H.C.; Khalil, R.A. Protein Kinase C as Regulator of Vascular Smooth Muscle Function and Potential Target in Vascular Disorders. Adv. Pharmacol. 2017, 78, 203–301. [Google Scholar]
  251. Khalil, R.A.; Lajoie, C.; Morgan, K.G. In situ determination of [Ca2+] i threshold for translocation of the alpha-protein kinase C isoform. Am. J. Physiol.-Cell Physiol. 1994, 266, C1544–C1551. [Google Scholar] [CrossRef]
  252. Liou, Y.M.; Morgan, K.G. Redistribution of protein kinase C isoforms in association with vascular hypertrophy of rat aorta. Am. J. Physiol. 1994, 267, C980–C989. [Google Scholar] [CrossRef]
  253. Chen, Y.; Zhang, Y.; Shan, M.; Zhou, Y.; Huang, Y.; Shi, L. Aerobic exercise-induced inhibition of PKCα/CaV1.2 pathway enhances the vasodilation of mesenteric arteries in hypertension. Arch. Biochem. Biophys. 2019, 678, 108191. [Google Scholar] [CrossRef]
  254. Inagaki, K.; Iwanaga, Y.; Sarai, N.; Onozawa, Y.; Takenaka, H.; Mochly-Rosen, D.; Kihara, Y. Tissue angiotensin II during progression or ventricular hypertrophy to heart failure in hypertensive rats; differential effects on PKC epsilon and PKC beta. J. Mol. Cell. Cardiol. 2002, 34, 1377–1385. [Google Scholar] [CrossRef]
  255. Palaniyandi, S.S.; Ferreira, J.C.B.; Brum, P.C.; Mochly-Rosen, D. PKC beta II inhibition attenuates myocardial infarction induced heart failure and is associated with a reduction of fibrosis and pro-inflammatory responses. J. Cell. Mol. Med. 2011, 15, 1769–1777. [Google Scholar] [CrossRef]
  256. Palaniyandi, S.S.; Inagaki, K.; Mochly-Rosen, D. Mast cells and epsilonPKC: A role in cardiac remodeling in hypertension-induced heart failure. J. Mol. Cell. Cardiol. 2008, 45, 779–786. [Google Scholar] [CrossRef]
  257. Phipps, M.S.; Cronin, C.A. Management of acute ischemic stroke. BMJ 2020, 368, l6983. [Google Scholar] [CrossRef] [PubMed]
  258. Crumrine, R.C.; Dubyak, G.; LaManna, J.C. Decreased protein kinase C activity during cerebral ischemia and after reperfusion in the adult rat. J. Neurochem. 1990, 55, 2001–2007. [Google Scholar] [CrossRef] [PubMed]
  259. Domańska-Janik, K.; Zalewska, T. Effect of brain ischemia on protein kinase C. J. Neurochem. 1992, 58, 1432–1439. [Google Scholar] [CrossRef] [PubMed]
  260. Chou, W.-H.; Messing, R.O. Protein Kinase C Isozymes in Stroke. Trends Cardiovasc. Med. 2005, 15, 47–51. [Google Scholar] [CrossRef] [PubMed]
  261. Bright, R.; Raval, A.P.; Dembner, J.M.; Pérez-Pinzón, M.A.; Steinberg, G.K.; Yenari, M.A.; Mochly-Rosen, D. Protein kinase C delta mediates cerebral reperfusion injury in vivo. J. Neurosci. 2004, 24, 6880–6888. [Google Scholar] [CrossRef] [PubMed]
  262. Bright, R.; Mochly-Rosen, D. The role of protein kinase C in cerebral ischemic and reperfusion injury. Stroke 2005, 36, 2781–2790. [Google Scholar] [CrossRef] [PubMed]
  263. Grune, J.; Yamazoe, M.; Nahrendorf, M. Electroimmunology and cardiac arrhythmia. Nat. Rev. Cardiol. 2021, 18, 547–564. [Google Scholar] [CrossRef]
  264. Kida, K.; Ichinose, F. Preventing ischemic brain injury after sudden cardiac arrest using NO inhalation. Crit. Care 2014, 18, 1–6. [Google Scholar] [CrossRef]
  265. Damluji, A.A.; Al-Damluji, M.S.; Pomenti, S.; Zhang, T.J.; Cohen, M.G.; Mitrani, R.D.; Moscucci, M.; Myerburg, R.J. Health Care Costs After Cardiac Arrest in the United States. Circ. Arrhythmia Electrophysiol. 2018, 11, e005689. [Google Scholar] [CrossRef]
  266. Perez-Pinzon, M.A.; Raval, A.P.; Dave, K.R. Protein kinase C and synaptic dysfunction after cardiac arrest. Pathophysiology 2005, 12, 29–34. [Google Scholar] [CrossRef]
  267. Lu, L.; Liu, M.; Sun, R.; Zheng, Y.; Zhang, P. Myocardial Infarction: Symptoms and Treatments. Cell Biochem. Biophys. 2015, 72, 865–867. [Google Scholar] [CrossRef] [PubMed]
  268. Benjamin, E.J.; Muntner, P.; Alonso, A.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Das, S.R.; et al. Heart Disease and Stroke Statistics-2019 Update: A Report From the American Heart Association. Circulation 2019, 139, e56–e528. [Google Scholar] [CrossRef] [PubMed]
  269. Chen, L.; Hahn, H.; Wu, G.; Chen, C.H.; Liron, T.; Schechtman, D.; Cavallaro, G.; Banci, L.; Guo, Y.; Bolli, R.; et al. Opposing cardioprotective actions and parallel hypertrophic effects of delta PKC and epsilon PKC. Proc. Natl. Acad. Sci. USA 2001, 98, 11114–11119. [Google Scholar] [CrossRef] [PubMed]
  270. Churchill, E.N.; Ferreira, J.C.; Brum, P.C.; Szweda, L.I.; Mochly-Rosen, D. Ischaemic preconditioning improves proteasomal activity and increases the degradation of deltaPKC during reperfusion. Cardiovasc. Res. 2010, 85, 385–394. [Google Scholar] [CrossRef] [PubMed]
  271. Ishii, H.; Jirousek, M.R.; Koya, D.; Takagi, C.; Xia, P.; Clermont, A.; Bursell, S.E.; Kern, T.S.; Ballas, L.M.; Heath, W.F.; et al. Amelioration of vascular dysfunctions in diabetic rats by an oral PKC beta inhibitor. Science 1996, 272, 728–731. [Google Scholar] [CrossRef] [PubMed]
  272. Armstrong, D.; Zidovetzki, R. Amplification of diacylglycerol activation of protein kinase C by cholesterol. Biophys. J. 2008, 94, 4700–4710. [Google Scholar] [CrossRef]
  273. Gineste, R.; Sirvent, A.; Paumelle, R.; Helleboid, S.; Aquilina, A.; Darteil, R.; Hum, D.W.; Fruchart, J.-C.; Staels, B. Phosphorylation of farnesoid X receptor by protein kinase C promotes its transcriptional activity. Mol. Endocrinol. 2008, 22, 2433–2447. [Google Scholar] [CrossRef]
  274. Huang, W.; Bansode, R.; Mehta, M.; Mehta, K.D. Loss of protein kinase Cβ function protects mice against diet-induced obesity and development of hepatic steatosis and insulin resistance. Hepatology 2009, 49, 1525–1536. [Google Scholar] [CrossRef]
  275. Araki, T.; Hayashi, M.; Saruta, T. Cloning and characterization of a novel gene promoting ureteric bud branching in the metanephros. Kidney Int. 2003, 64, 1968–1977. [Google Scholar] [CrossRef]
  276. Lang, D.; Terstesse, M.; Dohle, F.; Bangen, P.; Banas, B.; Pauels, H.-G.; Heidenreich, S. Protein kinase C (PKC) dependent induction of tissue factor (TF) by mesangial cells in response to inflammatory mediators and release during apoptosis. Br. J. Pharmacol. 2002, 137, 1116–1124. [Google Scholar] [CrossRef]
  277. Noh, H.; King, G. The role of protein kinase C activation in diabetic nephropathy. Kidney Int. 2007, 72, S49–S53. [Google Scholar] [CrossRef] [PubMed]
  278. Tuttle, K.R. Protein kinase C-beta inhibition for diabetic kidney disease. Diabetes Res. Clin. Pract. 2008, 82 (Suppl. 1), S70–S74. [Google Scholar] [CrossRef] [PubMed]
  279. Brenner, W.; Färber, G.; Herget, T.; Wiesner, C.; Hengstler, J.G.; Thüroff, J.W. Protein kinase C eta is associated with progression of renal cell carcinoma (RCC). Anticancer Res. 2003, 23, 4001–4006. [Google Scholar] [PubMed]
  280. Dempsey, E.C.; Newton, A.C.; Mochly-Rosen, D.; Fields, A.P.; Reyland, M.E.; Insel, P.A.; Messing, R.O.; Jackson, W.F.; Boerman, E.M.; Gottipati, K.R.; et al. Protein kinase C isozymes and the regulation of diverse cell responses. Am. J. Physiol. Lung Cell Mol. Physiol. 2000, 279, L429–L438. [Google Scholar] [CrossRef] [PubMed]
  281. Dempsey, E.C.; Cool, C.D.; Littler, C.M. Lung disease and PKCs. Pharmacol. Res. 2007, 55, 545–559. [Google Scholar] [CrossRef] [PubMed]
  282. Li, G.; Zhang, Y.; Fan, Z. Cellular Signal Transduction Pathways Involved in Acute Lung Injury Induced by Intestinal Ischemia-Reperfusion. Oxidative Med. Cell. Longev. 2021, 2021, 9985701. [Google Scholar] [CrossRef] [PubMed]
  283. Vachier, I.; Chanez, P.; Radeau, T.; Le Doucen, C.; Leger, C.; Godard, P. Cellular protein kinase C activity in asthma. Am. J. Respir. Crit. Care Med. 1997, 155, 1211–1216. [Google Scholar] [CrossRef]
  284. Do, J.-S.; Park, K.-S.; Seo, H.-J.; Kim, J.-H.; Hwang, J.-K.; Song, E.-J.; Nam, S.-Y. Therapeutic target validation of protein kinase C(PKC)-zeta for asthma using a mouse model. Int. J. Mol. Med. 2009, 23, 561–566. [Google Scholar]
  285. Mousavi, S.R.; Ahmadi, A.; Jamalkandi, S.A.; Salimian, J. Involvement of microRNAs in physiological and pathological processes in asthma. J. Cell Physiol. 2019, 234, 21547–21559. [Google Scholar] [CrossRef]
  286. Abdel-Halim, M.; Darwish, S.S.; ElHady, A.K.; Hoppstädter, J.; Abadi, A.H.; Hartmann, R.W.; Kiemer, A.K.; Engel, M. Pharmacological inhibition of protein kinase C (PKC)ζ downregulates the expression of cytokines involved in the pathogenesis of chronic obstructive pulmonary disease (COPD). Eur. J. Pharm. Sci. 2016, 93, 405–409. [Google Scholar] [CrossRef]
  287. Wang, J.; Sun, L.; Nie, Y.; Duan, S.; Zhang, T.; Wang, W.; Ye, R.D.; Hou, S.; Qian, F. Protein Kinase C δ (PKCδ) Attenuates Bleomycin Induced Pulmonary Fibrosis via Inhibiting NF-κB Signaling Pathway. Front. Physiol. 2020, 11, 367. [Google Scholar] [CrossRef] [PubMed]
  288. Mondrinos, M.J.; Kennedy, P.A.; Lyons, M.; Deutschman, C.S.; Kilpatrick, L.E. Protein kinase C and acute respiratory distress syndrome. Shock 2013, 39, 467–479. [Google Scholar] [CrossRef] [PubMed]
  289. Hrenak, J.; Simko, F. Renin–Angiotensin System: An Important Player in the Pathogenesis of Acute Respiratory Distress Syndrome. Int. J. Mol. Sci. 2020, 21, 8038. [Google Scholar] [CrossRef] [PubMed]
  290. Lee, P.-C.; Fang, Y.-F.; Yamaguchi, H.; Wang, W.-J.; Chen, T.-C.; Hong, X.; 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] [PubMed]
  291. Inoue, M.; Kishimoto, A.; Takai, Y.; Nishizuka, Y. Studies on a cyclic nucleotide-independent protein kinase and its proenzyme in mammalian tissues. II. Proenzyme and its activation by calcium-dependent protease from rat brain. J. Biol. Chem. 1977, 252, 7610–7616. [Google Scholar] [CrossRef] [PubMed]
  292. Arts, B.; Jabben, N.; Krabbendam, L.; van Os, J. Meta-analyses of cognitive functioning in euthymic bipolar patients and their first-degree relatives. Psychol. Med. 2008, 38, 771–785. [Google Scholar] [CrossRef] [PubMed]
  293. Forbes, N.F.; Carrick, L.A.; McIntosh, A.M.; Lawrie, S.M. Working memory in schizophrenia: A meta-analysis. Psychol. Med. 2009, 39, 889–905. [Google Scholar] [CrossRef] [PubMed]
  294. Pandey, G.N.; Rizavi, H.S.; Ren, X. Protein and mRNA expression of protein kinase C (PKC) in the postmortem brain of bipolar and schizophrenic subjects. J. Psychiatr. Res. 2020, 130, 362–371. [Google Scholar] [CrossRef]
  295. Young, K.W.; Garro, M.A.; Challiss, R.J.; Nahorski, S.R. NMDA-receptor regulation of muscarinic-receptor stimulated inositol 1, 4, 5-trisphosphate production and protein kinase C activation in single cerebellar granule neurons. J. Neurochem. 2004, 89, 1537–1546. [Google Scholar] [CrossRef]
  296. Choi, B.; Chae, H.; Park, T.; Oh, J.; Lim, J.; Kang, S.; Ha, H.; Kim, K. Protein kinase C regulates the activity and stability of serotonin N-acetyltransferase. J. Neurochem. 2004, 90, 442–454. [Google Scholar] [CrossRef]
  297. Poole, A.W.; Pula, G.; Hers, I.; Crosby, D.; Jones, M.L. PKC-interacting proteins: From function to pharmacology. Trends Pharmacol. Sci. 2004, 25, 528–535. [Google Scholar] [CrossRef] [PubMed]
  298. Sun, M.K.; Alkon, D.L. Protein kinase C activators as synaptogenic and memory therapeutics. Arch. Der Pharm. An. Int. J. Pharm. Med. Chem. 2009, 342, 689–698. [Google Scholar] [CrossRef] [PubMed]
  299. Kim, P.M.; Kornberg, M.D. Targeting PKC in microglia to promote remyelination and repair in the CNS. Curr. Opin. Pharmacol. 2022, 62, 103–108. [Google Scholar] [CrossRef] [PubMed]
  300. Amadio, M.; Battaini, F.; Pascale, A. The different facets of protein kinases C: Old and new players in neuronal signal transduction pathways. Pharmacol. Res. 2006, 54, 317–325. [Google Scholar] [CrossRef] [PubMed]
  301. Govoni, S.; Amadio, M.; Battaini, F.; Pascale, A. Senescence of the brain: Focus on cognitive kinases. Curr. Pharm. Des. 2010, 16, 660–671. [Google Scholar] [CrossRef] [PubMed]
  302. Geribaldi-Doldán, N.; Gómez-Oliva, R.; Domínguez-García, S.; Nunez-Abades, P.; Castro, C. Protein Kinase C: Targets to Regenerate Brain Injuries? Front. Cell Dev. Biol. 2019, 7, 39. [Google Scholar] [CrossRef] [PubMed]
  303. Zhang, D.; Anantharam, V.; Kanthasamy, A.; Kanthasamy, A.G. Neuroprotective effect of protein kinase Cδ inhibitor rottlerin in cell culture and animal models of Parkinson’s disease. J. Pharmacol. Exp. Ther. 2007, 322, 913–922. [Google Scholar] [CrossRef]
  304. Burguillos, M.A.; Deierborg, T.; Kavanagh, E.; Persson, A.; Hajji, N.; Garcia-Quintanilla, A.; Cano, J.; Brundin, P.; Englund, E.; Venero, J.L.; et al. Caspase signalling controls microglia activation and neurotoxicity. Nature 2011, 472, 319–324. [Google Scholar] [CrossRef]
  305. Kaleli, H.N.; Ozer, E.; Kaya, V.O.; Kutlu, O. Protein kinase C isozymes and autophagy during neurodegenerative disease progression. Cells 2020, 9, 553. [Google Scholar] [CrossRef]
  306. Laperle, A.H.; Sances, S.; Yucer, N.; Dardov, V.J.; Garcia, V.J.; Ho, R.; Fulton, A.N.; Jones, M.R.; Roxas, K.M.; Avalos, P.; et al. iPSC modeling of young-onset Parkinson’s disease reveals a molecular signature of disease and novel therapeutic candidates. Nat. Med. 2020, 26, 289–299. [Google Scholar] [CrossRef]
  307. Giusto, E.; Yacoubian, T.A.; Greggio, E.; Civiero, L. Pathways to Parkinson’s disease: A spotlight on 14-3-3 proteins. NPJ Park. Dis. 2021, 7, 85. [Google Scholar] [CrossRef] [PubMed]
  308. Sun, M.-K.; Alkon, D.L. Activation of Protein Kinase C Isozymes for the Treatment of Dementias. In Advances in Pharmacology; Michaelis, E.K., Michaelis, M.L., Eds.; Academic Press: Cambridge, MA, USA, 2012; pp. 273–302. [Google Scholar]
  309. Sun, M.-K.; Alkon, D.L. Chapter Two—The “Memory Kinases”: Roles of PKC Isoforms in Signal Processing and Memory Formation. In Progress in Molecular Biology and Translational Science; Khan, Z.U., Muly, E.C., Eds.; Academic Press: Cambridge, MA, USA, 2014; pp. 31–59. [Google Scholar]
  310. Sweitzer, S.M.; Wong, S.M.; Peters, M.C.; Mochly-Rosen, D.; Yeomans, D.C.; Kendig, J.J. Protein kinase C epsilon and gamma: Involvement in formalin-induced nociception in neonatal rats. J. Pharmacol. Exp. Ther. 2004, 309, 616–625. [Google Scholar] [CrossRef] [PubMed]
  311. Velzquez, K.T.; Mohammad, H.; Sweitzer, S.M. Protein kinase C in pain: Involvement of multiple isoforms. Pharmacol. Res. 2007, 55, 578–589. [Google Scholar] [CrossRef] [PubMed]
  312. Zhao, C.; Leitges, M.; Gereau, R.W. Isozyme-specific effects of protein kinase C in pain modulation. Anesthesiology 2011, 115, 1261–1270. [Google Scholar] [CrossRef] [PubMed]
  313. Giraud, F.; Pereira, E.; Anizon, F.; Moreau, P. Recent Advances in pain management: Relevant protein kinases and their inhibitors. Molecules 2021, 26, 2696. [Google Scholar] [CrossRef] [PubMed]
  314. Garrido, J.L.; Godoy, J.; Alvarez, A.; Bronfman, M.; Inestrosa, N.C. Protein kinase C inhibits amyloid beta peptide neurotoxicity by acting on members of the Wnt pathway. Faseb J. 2002, 16, 1982–1984. [Google Scholar] [CrossRef]
  315. Alkon, D.L.; Sun, M.K.; Nelson, T.J. PKC signaling deficits: A mechanistic hypothesis for the origins of Alzheimer’s disease. Trends Pharmacol. Sci. 2007, 28, 51–60. [Google Scholar] [CrossRef]
  316. Talman, V.; Pascale, A.; Jäntti, M.; Amadio, M.; Tuominen, R.K. Protein Kinase C Activation as a Potential Therapeutic Strategy in Alzheimer’s Disease: Is there a Role for Embryonic Lethal Abnormal Vision-like Proteins? Basic Clin. Pharmacol. Toxicol. 2016, 119, 149–160. [Google Scholar] [CrossRef]
  317. Callender, J.A.; Newton, A.C. Conventional protein kinase C in the brain: 40 years later. Neuronal Signal 2017, 1, Ns20160005. [Google Scholar] [CrossRef]
  318. Du, Y.; Zhao, Y.; Li, C.; Zheng, Q.; Tian, J.; Li, Z.; Huang, T.Y.; Zhang, W.; Xu, H. Inhibition of PKCδ reduces amyloid-β levels and reverses Alzheimer disease phenotypes. J. Exp. Med. 2018, 215, 1665–1677. [Google Scholar] [CrossRef]
  319. Chen, W.-H.; Chang, Y.-T.; Chen, Y.-C.; Cheng, S.-J.; Chen, C.-C. Spinal protein kinase C/extracellular signal–regulated kinase signal pathway mediates hyperalgesia priming. Pain 2018, 159, 907–918. [Google Scholar] [CrossRef] [PubMed]
  320. He, Y.; Wang, Z.J. Spinal and afferent PKC signaling mechanisms that mediate chronic pain in sickle cell disease. Neurosci. Lett. 2019, 706, 56–60. [Google Scholar] [CrossRef] [PubMed]
  321. Kopach, O.; Krotov, V.; Shysh, A.; Sotnic, A.; Viatchenko-Karpinski, V.; Dosenko, V.; Voitenko, N. Spinal PKCα inhibition and gene-silencing for pain relief: AMPAR trafficking at the synapses between primary afferents and sensory interneurons. Sci. Rep. 2018, 8, 10285. [Google Scholar] [CrossRef] [PubMed]
  322. Zygmunt, P.M.; Petersson, J.; Andersson, D.A.; Chuang, H.-H.; Sørgård, M.; Di Marzo, V.; Julius, D.; Högestätt, E.D. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature 1999, 400, 452–457. [Google Scholar] [CrossRef] [PubMed]
  323. Gross, E.R.; Urban, T.J.; Hsu, A.K.; Qvit, N.; Gross, G.J.; Mochly-Rosen, D. TRPV1 Mediates Remote and Direct Cardioprotection. Am. Heart Assoc. 2013. [Google Scholar] [CrossRef]
  324. Fang, J.; Wang, S.; Zhou, J.; Shao, X.; Sun, H.; Liang, Y.; He, X.; Jiang, Y.; Liu, B.; Jin, X.; et al. Electroacupuncture Regulates Pain Transition Through Inhibiting PKCε and TRPV1 Expression in Dorsal Root Ganglion. Front. Neurosci. 2021, 15, 685715. [Google Scholar] [CrossRef]
  325. Chuang, H.-H.; Prescott, E.D.; Kong, H.; Shields, S.; Jordt, S.-E.; Basbaum, A.I.; Chao, M.V.; Julius, D. Bradykinin and nerve growth factor release the capsaicin receptor from PtdIns(4,5)P2-mediated inhibition. Nature 2001, 411, 957–962. [Google Scholar] [CrossRef] [PubMed]
  326. García-Bernal, F.; Geribaldi-Doldán, N.; Domínguez-García, S.; Carrasco, M.; Murillo-Carretero, M.; Delgado-Ariza, A.; Díez-Salguero, M.; Verástegui, C.; Castro, C. Protein Kinase C Inhibition Mediates Neuroblast Enrichment in Mechanical Brain Injuries. Front. Cell. Neurosci. 2018, 12, 462. [Google Scholar] [CrossRef]
  327. Brennan, H.A.; Vu, M.A.T.; Maciejewski, P.K.; van Dyck, C.H.; Gottron, M.; Arnsten, A.F. Inhibition of protein kinase C signaling protects prefrontal cortex dendritic spines and cognition from the effects of chronic stress. Proc. Natl. Acad. Sci. USA 2009, 106, 17957–17962. [Google Scholar]
  328. Turner, R.S.; Raynor, R.L.; Mazzei, G.J.; Girard, P.R.; Kuo, J.F. Developmental studies of phospholipid-sensitive Ca2+-dependent protein kinase and its substrates and of phosphoprotein phosphatases in rat brain. Proc. Natl. Acad. Sci. USA 1984, 81, 3143–3147. [Google Scholar] [CrossRef]
  329. Corbit, K.C.; Soh, J.-W.; Yoshida, K.; Eves, E.M.; Weinstein, I.B.; Rosner, M.R. Different protein kinase C isoforms determine growth factor specificity in neuronal cells. Mol. Cell Biol. 2000, 20, 5392–5403. [Google Scholar] [CrossRef] [PubMed]
  330. Müller-Oerlinghausen, B.; Berghöfer, A.; Bauer, M. Bipolar disorder. Lancet 2002, 359, 241–247. [Google Scholar] [CrossRef] [PubMed]
  331. Manji, H.K.; Lenox, R.H. The nature of bipolar disorder. J. Clin. Psychiatry 2000, 61 (Suppl. 13), 42–57. [Google Scholar] [PubMed]
  332. Zarate, C.A.; Manji, H.K. Protein kinase C inhibitors: Rationale for use and potential in the treatment of bipolar disorder. CNS Drugs 2009, 23, 569–582. [Google Scholar] [CrossRef] [PubMed]
  333. Mann, J.J.; Currier, D.; Quiroz, J.A.; Manji, H.K. Chapter 60—Neurobiology of Severe Mood and Anxiety Disorders. In Basic Neurochemistry, 8th ed.; Brady, S.T., Siegel, G.J., Albers, R.W., Price, D.L., Eds.; Academic Press: New York, NY, USA, 2012; pp. 1021–1036. [Google Scholar] [CrossRef]
  334. Saxena, A.; Scaini, G.; Bavaresco, D.V.; Leite, C.; Valvassoria, S.S.; Carvalho, A.F.; Quevedo, J. Role of protein kinase C in bipolar disorder: A review of the current literature. Complex. Psychiatry 2017, 3, 108–124. [Google Scholar] [CrossRef] [PubMed]
  335. Manji, H.K.; Lenox, R.H. Protein kinase C signaling in the brain: Molecular transduction of mood stabilization in the treatment of manic-depressive illness. Biol. Psychiatry 1999, 46, 1328–1351. [Google Scholar] [CrossRef] [PubMed]
  336. Denning, M.F. Epidermal keratinocytes: Regulation of multiple cell phenotypes by multiple protein kinase C isoforms. Int. J. Biochem. Cell Biol. 2004, 36, 1141–1146. [Google Scholar] [CrossRef] [PubMed]
  337. D’Costa, A.M.; Robinson, J.K.; Maududi, T.; Chaturvedi, V.; Nickoloff, B.J.; Denning, M.F. The proapoptotic tumor suppressor protein kinase C-delta is lost in human squamous cell carcinomas. Oncogene 2006, 25, 378–386. [Google Scholar] [CrossRef]
  338. Jansen, A.P.; Verwiebe, E.G.; Dreckschmidt, N.E.; Wheeler, D.L.; Oberley, T.D.; Verma, A.K. Protein kinase C-epsilon transgenic mice: A unique model for metastatic squamous cell carcinoma. Cancer Res. 2001, 61, 808–812. [Google Scholar]
  339. Tibudan, S.S.; Wang, Y.; Denning, M.F. Activation of Protein Kinase C Triggers Irreversible Cell CycleWithdrawal In Human Keratinocytes. J. Investig. Dermatol. 2002, 119, 1282–1289. [Google Scholar] [CrossRef]
  340. Segrelles, C.; Moral, M.; Lara, M.F.; Ruiz, S.; Santos, M.; Leis, H.; García-Escudero, R.; Martínez-Cruz, A.B.; Martínez-Palacio, J.; Hernández, P.; et al. Molecular determinants of Akt-induced keratinocyte transformation. Oncogene 2006, 25, 1174–1185. [Google Scholar] [CrossRef] [PubMed]
  341. Maioli, E.; Valacchi, G. Rottlerin: Bases for a possible usage in psoriasis. Curr. Drug Metab. 2010, 11, 425–430. [Google Scholar] [CrossRef] [PubMed]
  342. Skvara, H.; Dawid, M.; Kleyn, E.; Wolff, B.; Meingassner, J.G.; Knight, H.; Dumortier, T.; Kopp, T.; Fallahi, N.; Stary, G.; et al. The PKC inhibitor AEB071 may be a therapeutic option for psoriasis. J. Clin. Investig. 2008, 118, 3151–3159. [Google Scholar] [CrossRef] [PubMed]
  343. Soloff, R.S.; Katayama, C.; Lin, M.Y.; Feramisco, J.R.; Hedrick, S.M. Targeted deletion of protein kinase C lambda reveals a distribution of functions between the two atypical protein kinase C isoforms. J. Immunol. 2004, 173, 3250–3260. [Google Scholar] [CrossRef] [PubMed]
  344. Varin, M.-M.; Guerrier, T.; Devauchelle-Pensec, V.; Jamin, C.; Youinou, P.; Pers, J.-O. In Sjögren’s syndrome, B lymphocytes induce epithelial cells of salivary glands into apoptosis through protein kinase C delta activation. Autoimmun. Rev. 2012, 11, 252–258. [Google Scholar] [CrossRef] [PubMed]
  345. Mohanraj, M.; Sekar, P.; Liou, H.-H.; Chang, S.-F.; Lin, W.-W. The mycobacterial adjuvant analogue TDB attenuates neuroinflammation via Mincle-independent PLC-γ1/PKC/ERK signaling and microglial polarization. Mol. Neurobiol. 2019, 56, 1167–1187. [Google Scholar] [CrossRef]
  346. Birchall, A.M.; Bishop, J.; Bradshaw, D.; Cline, A.; Coffey, J.; Elliott, L.H.; Gibson, V.M.; Greenham, A.; Hallam, T.J.; Harris, W. Ro 32-0432, a selective and orally active inhibitor of protein kinase C prevents T-cell activation. J. Pharmacol. Exp. Ther. 1994, 268, 922–929. [Google Scholar]
  347. Zhong, C.; Wu, Y.; Chang, H.; Liu, C.; Zhou, L.; Zou, J.; Qi, Z. Effect of PKC inhibitor on experimental autoimmune myocarditis in Lewis rats. Oncotarget 2017, 8, 54187–54198. [Google Scholar] [CrossRef]
  348. DiMartino, M.; Wolff, C.; Patil, A.; Nambi, P. Effects of a protein kinase C inhibitor (PKCI) on the development of adjuvant-induced arthritis (AA) in rats. Inflamm. Res. 1995, 44 (Suppl. 2), S123–S124. [Google Scholar] [CrossRef]
  349. Leppänen, T.; Tuominen, R.K.; Moilanen, E. Protein Kinase C and its Inhibitors in the Regulation of Inflammation: Inducible Nitric Oxide Synthase as an Example. Basic. Clin. Pharmacol. Toxicol. 2014, 114, 37–43. [Google Scholar] [CrossRef]
  350. Jacobson, P.B.; Kuchera, S.L.; Metz, A.; Schächtele, C.; Imre, K.; Schrier, D.J. Anti-inflammatory properties of Gö 6850: A selective inhibitor of protein kinase C. J. Pharmacol. Exp. Ther. 1995, 275, 995–1002. [Google Scholar] [PubMed]
  351. Le, M.Q.; Kim, M.S.; Song, Y.S.; Ryu, H.W.; Oh, S.R.; Yoon, D.Y. 6-O-Veratroyl catalpol suppresses pro-inflammatory cytokines via regulation of extracellular signal-regulated kinase and nuclear factor-κB in human monocytic cells. Biochimie 2015, 119, 52–59. [Google Scholar] [CrossRef] [PubMed]
  352. Mineo, C.; Anderson, R.G. Potocytosis. Robert Feulgen Lecture. Histochem. Cell Biol. 2001, 116, 109–118. [Google Scholar] [CrossRef] [PubMed]
  353. Empig, C.J.; Goldsmith, M.A. Association of the caveola vesicular system with cellular entry by filoviruses. J. Virol. 2002, 76, 5266–5270. [Google Scholar] [CrossRef] [PubMed]
  354. Stuart, A.D.; Eustace, H.E.; McKee, T.A.; Brown, T.D.K. A novel cell entry pathway for a DAF-using human enterovirus is dependent on lipid rafts. J. Virol. 2002, 76, 9307–9322. [Google Scholar] [CrossRef] [PubMed]
  355. Marjomäki, V.; Pietiäinen, V.; Matilainen, H.; Upla, P.; Ivaska, J.; Nissinen, L.; Reunanen, H.; Huttunen, P.; Hyypiä, T.; Heino, J. Internalization of echovirus 1 in caveolae. J. Virol. 2002, 76, 1856–1865. [Google Scholar] [CrossRef]
  356. Mañes, S.; del Real, G.; Lacalle, R.A.; Lucas, P.; Gómez-Moutón, C.; Sánchez-Palomino, S.; Delgado, R.; Alcamí, J.; Mira, E.; Martínez-A, C. Membrane raft microdomains mediate lateral assemblies required for HIV-1 infection. EMBO Rep. 2000, 1, 190–196. [Google Scholar] [CrossRef]
  357. Guyader, M.; Kiyokawa, E.; Abrami, L.; Turelli, P.; Trono, D. Role for human immunodeficiency virus type 1 membrane cholesterol in viral internalization. J. Virol. 2002, 76, 10356–10364. [Google Scholar] [CrossRef]
  358. Warrilow, D.; Gardner, J.; Darnell, G.A.; Suhrbier, A.; Harrich, D. HIV Type 1 Inhibition by Protein Kinase C Modulatory Compounds. AIDS Res. Human. Retroviruses 2006, 22, 854–864. [Google Scholar] [CrossRef]
  359. Contreras, X.; Mzoughi, O.; Gaston, F.; Peterlin, M.B.; Bahraoui, E. Protein kinase C-delta regulates HIV-1 replication at an early post-entry step in macrophages. Retrovirology 2012, 9, 37. [Google Scholar] [CrossRef]
  360. Dang, Y.-F.; Qiu, T.-X.; Song, D.-W.; Liu, L. PMA-triggered PKCε activity enhances Nrf2-mediated antiviral response on fish rhabdovirus infection. Fish. Shellfish. Immunol. 2019, 87, 871–878. [Google Scholar] [CrossRef] [PubMed]
  361. Souza-Souza, K.F.C.; Gonçalves-De-Albuquerque, C.F.; Cirne-Santos, C.; Paixão, I.C.N.P.; Burth, P. Alphavirus Replication: The Role of Cardiac Glycosides and Ion Concentration in Host Cells. Biomed. Res. Int. 2020, 2020, 2813253. [Google Scholar] [CrossRef] [PubMed]
  362. Yousuf, M.A.; Lee, J.S.; Zhou, X.; Ramke, M.; Lee, J.Y.; Chodosh, J.; Rajaiya, J. Protein Kinase C Signaling in Adenoviral Infection. Biochemistry 2016, 55, 5938–5946. [Google Scholar] [CrossRef] [PubMed]
  363. Constantinescu, S.N.; Cernescu, C.D.; Popescu, L.M. Effects of protein kinase C inhibitors on viral entry and infectivity. FEBS Lett. 1991, 292, 31–33. [Google Scholar] [PubMed]
  364. Park, R.; Baines Joel, D. Herpes Simplex Virus Type 1 Infection Induces Activation and Recruitment of Protein Kinase C to the Nuclear Membrane and Increased Phosphorylation of Lamin B. J. Virol. 2006, 80, 494–504. [Google Scholar] [CrossRef]
  365. Sieczkarski, S.B.; Brown, H.A.; Whittaker, G.R. Role of protein kinase C βII in influenza virus entry via late endosomes. J. Virol. 2003, 77, 460–469. [Google Scholar] [CrossRef]
  366. Mondal, A.; Dawson, A.R.; Potts, G.K.; Freiberger, E.C.; Baker, S.F.; Moser, L.A.; Bernard, K.A.; Coon, J.J.; Mehle, A. Influenza virus recruits host protein kinase C to control assembly and activity of its replication machinery. eLife 2017, 6, e26910. [Google Scholar] [CrossRef]
  367. Noppakunmongkolchai, W.; Poyomtip, T.; Jittawuttipoka, T.; Luplertlop, N.; Sakuntabhai, A.; Chimnaronk, S.; Jirawatnotai, S.; Tohtong, R. Inhibition of protein kinase C promotes dengue virus replication. Virol. J. 2016, 13, 35. [Google Scholar] [CrossRef]
  368. Monick, M.M.; Staber, J.M.; Thomas, K.W.; Hunninghake, G.W. Respiratory Syncytial Virus Infection Results in Activation of Multiple Protein Kinase C Isoforms Leading to Activation of Mitogen-Activated Protein Kinase. J. Immunol. 2001, 166, 2681. [Google Scholar] [CrossRef]
  369. San-Juan-Vergara, H.; Peeples, M.E.; Lockey, R.F.; Mohapatra, S.S. Protein kinase C-α activity is required for respiratory syncytial virus fusion to human bronchial epithelial cells. J. Virol. 2004, 78, 13717–13726. [Google Scholar] [CrossRef]
  370. Yousuf, M.A.; Zhou, X.; Mukherjee, S.; Chintakuntlawar, A.V.; Lee, J.Y.; Ramke, M.; Chodosh, J.; Rajaiya, J. Caveolin-1 associated adenovirus entry into human corneal cells. PLoS ONE 2013, 8, e77462. [Google Scholar] [CrossRef] [PubMed]
  371. Crane, J.K.; Vezina, C.M. Externalization of host cell protein kinase C during enteropathogenic Escherichia coli infection. Cell Death Differ. 2005, 12, 115–127. [Google Scholar] [CrossRef] [PubMed]
  372. Gauron, M.C.; Newton, A.C.; Colombo, M.I. PKCα Is Recruited to Staphylococcus aureus-Containing Phagosomes and Impairs Bacterial Replication by Inhibition of Autophagy. Front. Immunol. 2021, 12, 662987. [Google Scholar] [CrossRef] [PubMed]
  373. Bhalla, M.; Van Ngo, H.; Gyanwali, G.C.; Ireton, K. The Host Scaffolding Protein Filamin A and the Exocyst Complex Control Exocytosis during InlB-Mediated Entry of Listeria monocytogenes. Infect. Immun. 2019, 87, e00689-18. [Google Scholar] [CrossRef] [PubMed]
  374. Sah, P.; Nelson, N.H.; Shaw, J.H.; Lutter, E.I. Chlamydia trachomatis recruits protein kinase C during infection. Pathog. Dis. 2019, 77, ftz061. [Google Scholar] [CrossRef] [PubMed]
  375. LaFayette, S.L.; Collins, C.; Zaas, A.K.; Schell, W.A.; Betancourt-Quiroz, M.; Gunatilaka, A.A.L.; Perfect, J.R.; Cowen, L.E. PKC signaling regulates drug resistance of the fungal pathogen Candida albicans via circuitry comprised of Mkc1, calcineurin, and Hsp90. PLoS Pathog. 2010, 6, e1001069. [Google Scholar] [CrossRef] [PubMed]
  376. Schwegmann, A.; Guler, R.; Cutler, A.J.; Arendse, B.; Horsnell, W.G.C.; Flemming, A.; Kottmann, A.H.; Ryan, G.; Hide, W.; Leitges, M.; et al. Protein kinase C δ is essential for optimal macrophage-mediated phagosomal containment of Listeria monocytogenes. Proc. Natl. Acad. Sci. USA 2007, 104, 16251–16256. [Google Scholar] [CrossRef] [PubMed]
  377. Parihar, S.P.; Ozturk, M.; Marakalala, M.J.; Loots, D.T.; Hurdayal, R.; Maasdorp, D.B.; Van Reenen, M.; E Zak, D.; Darboe, F.; Penn-Nicholson, A.; et al. Protein kinase C-delta (PKCδ), a marker of inflammation and tuberculosis disease progression in humans, is important for optimal macrophage killing effector functions and survival in mice. Mucosal Immunol. 2018, 11, 496–511. [Google Scholar] [CrossRef]
  378. St-Denis, A.; Caouras, V.; Gervais, F.; Descoteaux, A. Role of protein kinase C-α in the control of infection by intracellular pathogens in macrophages. J. Immunol. 1999, 163, 5505–5511. [Google Scholar] [CrossRef]
  379. Guler, R.; Afshar, M.; Arendse, B.; Parihar, S.P.; Revaz-Breton, M.; Leitges, M.; Schwegmann, A.; Brombacher, F. PKCδ regulates IL-12p40/p70 production by macrophages and dendritic cells, driving a type 1 healer phenotype in cutaneous leishmaniasis. Eur. J. Immunol. 2011, 41, 706–715. [Google Scholar] [CrossRef]
  380. Bhattacharyya, S.; Ghosh, S.; Sen, P.; Roy, S.; Majumdar, S. Selective impairment of protein kinase C isotypes in murine macrophage by Leismania donovani. Mol. Cell. Biochem. 2001, 216, 47–57. [Google Scholar] [CrossRef] [PubMed]
  381. Pandey, D.; Gratton, J.-P.; Rafikov, R.; Black, S.M.; Fulton, D.J.R. Calcium/calmodulin-dependent kinase II mediates the phosphorylation and activation of NADPH oxidase 5. Mol. Pharmacol. 2011, 80, 407–415. [Google Scholar] [CrossRef] [PubMed]
  382. Mochly-Rosen, D.; Das, K.; Grimes, K.V. Protein kinase C, an elusive therapeutic target? Nat. Rev. Drug Discov. 2012, 11, 937–957. [Google Scholar] [CrossRef] [PubMed]
  383. Rahimova, N.; Cooke, M.; Zhang, S.; Baker, M.J.; Kazanietz, M.G. The PKC universe keeps expanding: From cancer initiation to metastasis. Adv. Biol. Regul. 2020, 78, 100755. [Google Scholar] [CrossRef] [PubMed]
  384. Lordén, G.; Newton, A.C. Conventional protein kinase C in the brain: Repurposing cancer drugs for neurodegenerative treatment? Neuronal Signal 2021, 5, Ns20210036. [Google Scholar] [CrossRef]
  385. Ghoreschi, K.; Laurence, A.; O’shea, J.J. Selectivity and therapeutic inhibition of kinases: To be or not to be? Nat. Immunol. 2009, 10, 356–360. [Google Scholar] [CrossRef]
  386. Wu-Zhang, A.X.; Newton, A.C. Protein kinase C pharmacology: Refining the toolbox. Biochem. J. 2013, 452, 195–209. [Google Scholar] [CrossRef]
  387. Echols, N.; Harrison, P.; Balasubramanian, S.; Luscombe, N.M.; Bertone, P.; Zhang, Z.; Gerstein, M. Comprehensive analysis of amino acid and nucleotide composition in eukaryotic genomes, comparing genes and pseudogenes. Nucleic Acids Res. 2002, 30, 2515–2523. [Google Scholar] [CrossRef]
  388. Ubersax, J.A.; Ferrell, J.E., Jr. Mechanisms of specificity in protein phosphorylation. Nat. Rev. Mol. Cell Biol. 2007, 8, 530–541. [Google Scholar] [CrossRef]
  389. Karaman, M.W.; Herrgard, S.; Treiber, D.K.; Gallant, P.; Atteridge, C.E.; Campbell, B.T.; Chan, K.W.; Ciceri, P.; Davis, M.I.; Edeen, P.T.; et al. A quantitative analysis of kinase inhibitor selectivity. Nat. Biotechnol. 2008, 26, 127–132. [Google Scholar] [CrossRef]
  390. Lee, K.-W.; Kim, S.G.; Kim, H.-P.; Kwon, E.; You, J.; Choi, H.-J.; Park, J.-H.; Kang, B.-C.; Im, S.-A.; Kim, T.-Y.; et al. Enzastaurin, a protein kinase C beta inhibitor, suppresses signaling through the ribosomal S6 kinase and bad pathways and induces apoptosis in human gastric cancer cells. Cancer Res. 2008, 68, 1916–1926. [Google Scholar] [CrossRef] [PubMed]
  391. Bourhill, T.; Narendran, A.; Johnston, R.N. Enzastaurin: A lesson in drug development. Crit. Rev. Oncol./Hematol. 2017, 112, 72–79. [Google Scholar] [CrossRef]
  392. Blumberg, P.M.; Jaken, S.; König, B.; Sharkey, N.A.; Leach, K.L.; Jeng, A.Y.; Yeh, E. Mechanism of action of the phorbol ester tumor promoters: Specific receptors for lipophilic ligands. Biochem. Pharmacol. 1984, 33, 933–940. [Google Scholar] [CrossRef] [PubMed]
  393. Marquez, V.E.; Nacro, K.; Benzaria, S.; Lee, J.; Sharma, R.; Teng, K.; Milne, G.W.; Bienfait, B.; Wang, S.; Lewin, N.E.; et al. The transition from a pharmacophore-guided approach to a receptor-guided approach in the design of potent protein kinase C ligands. Pharmacol. Ther. 1999, 82, 251–261. [Google Scholar] [CrossRef] [PubMed]
  394. Kang, J.-H.; Kim, S.Y.; Lee, J.; Marquez, V.E.; Lewin, N.E.; Pearce, L.V.; Blumberg, P.M. Macrocyclic diacylglycerol-bis-lactones as conformationally constrained analogues of diacylglycerol-lactones. Interactions with protein kinase C. J. Med. Chem. 2004, 47, 4000–4007. [Google Scholar] [CrossRef] [PubMed]
  395. Cooke, M.; Zhou, X.; Casado-Medrano, V.; Lopez-Haber, C.; Baker, M.J.; Garg, R.; Ann, J.; Lee, J.; Blumberg, P.M.; Kazanietz, M.G. Characterization of AJH-836, a diacylglycerol-lactone with selectivity for novel PKC isozymes. J. Biol. Chem. 2018, 293, 8330–8341. [Google Scholar] [CrossRef] [PubMed]
  396. Garcia-Bermejo, M.L.; Leskow, F.C.; Fujii, T.; Wang, Q.; Blumberg, P.M.; Ohba, M.; Kuroki, T.; Han, K.C.; Lee, J.; Marquez, V.E.; et al. Diacylglycerol (DAG)-lactones, a new class of protein kinase C (PKC) agonists, induce apoptosis in LNCaP prostate cancer cells by selective activation of PKCalpha. J. Biol. Chem. 2002, 277, 645–655. [Google Scholar] [CrossRef] [PubMed]
  397. Sun, M.-K.; Alkon, D.L. Bryostatin-1: Pharmacology and therapeutic potential as a CNS drug. CNS Drug Rev. 2006, 12, 1–8. [Google Scholar] [CrossRef]
  398. Ruan, B.-F.; Zhu, H.-L. The chemistry and biology of the bryostatins: Potential PKC inhibitors in clinical development. Curr. Med. Chem. 2012, 19, 2652–2664. [Google Scholar] [CrossRef]
  399. Mackay, H.J.; Twelves, C.J. Targeting the protein kinase C family: Are we there yet? Nat. Rev. Cancer 2007, 7, 554–562. [Google Scholar] [CrossRef]
  400. Raghuvanshi, R.; Bharate, S.B. Preclinical and clinical studies on bryostatins, a class of marine-derived protein kinase C modulators: A mini-review. Curr. Top. Med. Chem. 2020, 20, 1124–1135. [Google Scholar] [CrossRef] [PubMed]
  401. Geiger, T.; Müller, M.; Dean, N.M.; Fabbro, D. Antitumor activity of a PKC-alpha antisense oligonucleotide in combination with standard chemotherapeutic agents against various human tumors transplanted into nude mice. Anticancer Drug Des. 1998, 13, 35–45. [Google Scholar] [PubMed]
  402. Wang, C.; Wang, X.; Liang, H.; Wang, T.; Yan, X.; Cao, M.; Wang, N.; Zhang, S.; Zen, K.; Zhang, C.; et al. miR-203 Inhibits Cell Proliferation and Migration of Lung Cancer Cells by Targeting PKCα. PLoS ONE 2013, 8, e73985. [Google Scholar] [CrossRef] [PubMed]
  403. Fabbro, D.; Ruetz, S.; Bodis, S.; Pruschy, M.; Csermak, K.; Man, A.; Campochiaro, P.; Wood, J.; O’Reilly, T.; Meyer, T. PKC412--a protein kinase inhibitor with a broad therapeutic potential. Anticancer Drug Des. 2000, 15, 17–28. [Google Scholar] [PubMed]
  404. Tremblay, G.; Dolph, M.; Patel, S.; Brandt, P.; Forsythe, A. Cost-effectiveness analysis for midostaurin versus standard of care in acute myeloid leukemia in the United Kingdom. Cost. Eff. Resour. Alloc. 2018, 16, 1–13. [Google Scholar] [CrossRef] [PubMed]
  405. Medicine, C.T.g.U.S.N.L.o. Available online: https://clinicaltrials.gov/ct2/results?cond=&term=Midostaurin&cntry=&state=&city=&dist= (accessed on 15 July 2022).
  406. Crump, M.; Leppä, S.; Fayad, L.; Lee, J.J.; Di Rocco, A.; Ogura, M.; Hagberg, H.; Schnell, F.; Rifkin, R.; Mackensen, A.; et al. Randomized, Double-Blind, Phase III Trial of Enzastaurin Versus Placebo in Patients Achieving Remission After First-Line Therapy for High-Risk Diffuse Large B-Cell Lymphoma. J. Clin. Oncol. 2016, 34, 2484–2492. [Google Scholar] [CrossRef] [PubMed]
  407. 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]
  408. Grønberg, B.H.; Ciuleanu, T.; Fløtten, Ø.; Knuuttila, A.; Abel, E.; Langer, S.W.; Krejcy, K.; Liepa, A.M.; Munoz, M.; Hahka-Kemppinen, M.; et al. A placebo-controlled, randomized phase II study of maintenance enzastaurin following whole brain radiation therapy in the treatment of brain metastases from lung cancer. Lung Cancer 2012, 78, 63–69. [Google Scholar] [CrossRef]
  409. Usha, L.; Sill, M.W.; Darcy, K.M.; Benbrook, D.M.; Hurteau, J.A.; Michelin, D.P.; Mannel, R.S.; Hanjani, P.; De Geest, K.; Godwin, A.K. A Gynecologic Oncology Group phase II trial of the protein kinase C-beta inhibitor, enzastaurin and evaluation of markers with potential predictive and prognostic value in persistent or recurrent epithelial ovarian and primary peritoneal malignancies. Gynecol. Oncol. 2011, 121, 455–461. [Google Scholar] [CrossRef]
  410. Morschhauser, F.; Seymour, J.F.; Kluin-Nelemans, H.C.; Grigg, A.; Wolf, M.; Pfreundschuh, M.; Tilly, H.; Raemaekers, J.; van’t Veer, M.B.; Milpied, N.; et al. A phase II study of enzastaurin, a protein kinase C beta inhibitor, in patients with relapsed or refractory mantle cell lymphoma. Ann. Oncol. 2008, 19, 247–253. [Google Scholar] [CrossRef]
  411. Querfeld, C.; Kuzel, T.M.; Kim, Y.H.; Porcu, P.; Duvic, M.; Musiek, A.; Rook, A.H.; Mark, L.A.; Pinter-Brown, L.; Hamid, O.; et al. Multicenter phase II trial of enzastaurin in patients with relapsed or refractory advanced cutaneous T-cell lymphoma. Leuk. Lymphoma 2011, 52, 1474–1480. [Google Scholar] [CrossRef] [PubMed]
  412. Mina, L.; Krop, I.; Zon, R.T.; Isakoff, S.J.; Schneider, C.J.; Yu, M.; Johnson, C.; Vaughn, L.G.; Wang, Y.; Hristova-Kazmierski, M.; et al. A phase II study of oral enzastaurin in patients with metastatic breast cancer previously treated with an anthracycline and a taxane containing regimen. Investig. New Drugs 2009, 27, 565–570. [Google Scholar] [CrossRef] [PubMed]
  413. Clément-Duchêne, C.; Natale, R.B.; Jahan, T.; Krupitskaya, Y.; Osarogiagbon, R.; Sanborn, R.E.; Bernstein, E.D.; Dudek, A.Z.; Latz, J.E.; Shi, P.; et al. A phase II study of enzastaurin in combination with erlotinib in patients with previously treated advanced non-small cell lung cancer. Lung Cancer 2012, 78, 57–62. [Google Scholar] [CrossRef] [PubMed]
  414. Chiappori, A.; Bepler, G.; Barlesi, F.; Soria, J.-C.; Reck, M.; Bearz, A.; Barata, F.; Scagliotti, G.; Park, K.; Wagle, A.; et al. Phase II, double-blinded, randomized study of enzastaurin plus pemetrexed as second-line therapy in patients with advanced non-small cell lung cancer. J. Thorac. Oncol. 2010, 5, 369–375. [Google Scholar] [CrossRef] [PubMed]
  415. Butowski, N.; Chang, S.M.; Lamborn, K.R.; Polley, M.-Y.; Pieper, R.; Costello, J.F.; Vandenberg, S.; Parvataneni, R.; Nicole, A.; Sneed, P.K.; et al. Phase II and pharmacogenomics study of enzastaurin plus temozolomide during and following radiation therapy in patients with newly diagnosed glioblastoma multiforme and gliosarcoma. Neuro Oncol. 2011, 13, 1331–1338. [Google Scholar] [CrossRef]
  416. Dreicer, R.; Garcia, J.; Rini, B.; Vogelzang, N.; Srinivas, S.; Somer, B.; Shi, P.; Kania, M.; Raghavan, D. A randomized, double-blind, placebo-controlled, Phase II study with and without enzastaurin in combination with docetaxel-based chemotherapy in patients with castration-resistant metastatic prostate cancer. Investig. New Drugs 2013, 31, 1044–1050. [Google Scholar] [CrossRef]
  417. Vergote, I.B.; Chekerov, R.; Amant, F.; Harter, P.; Casado, A.; Emerich, J.; Bauknecht, T.; Mansouri, K.; Myrand, S.P.; Nguyen, T.S.; et al. Randomized, phase II, placebo-controlled, double-blind study with and without enzastaurin in combination with paclitaxel and carboplatin as first-line treatment followed by maintenance treatment in advanced ovarian cancer. J. Clin. Oncol. 2013, 31, 3127–3132. [Google Scholar] [CrossRef]
  418. Wolff, R.A.; Fuchs, M.; Di Bartolomeo, M.; Hossain, A.M.; Stoffregen, C.; Nicol, S.; Heinemann, V. A double-blind, randomized, placebo-controlled, phase 2 study of maintenance enzastaurin with 5-fluorouracil/leucovorin plus bevacizumab after first-line therapy for metastatic colorectal cancer. Cancer 2012, 118, 4132–4138. [Google Scholar] [CrossRef]
  419. Richards, D.A.; Kuefler, P.R.; Becerra, C.; Wilfong, L.S.; Gersh, R.H.; Boehm, K.A.; Zhan, F.; Asmar, L.; Myrand, S.P.; Hozak, R.R.; et al. Gemcitabine plus enzastaurin or single-agent gemcitabine in locally advanced or metastatic pancreatic cancer: Results of a phase II, randomized, noncomparative study. Investig. New Drugs 2011, 29, 144–153. [Google Scholar] [CrossRef]
  420. 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] [CrossRef]
  421. Evenou, J.P.; Wagner, J.; Zenke, G.; Brinkmann, V.; Wagner, K.; Kovarik, J.; Welzenbach, K.A.; Weitz-Schmidt, G.; Guntermann, C.; Towbin, H.; et al. The potent protein kinase C-selective inhibitor AEB071 (sotrastaurin) represents a new class of immunosuppressive agents affecting early T-cell activation. J. Pharmacol. Exp. Ther. 2009, 330, 792–801. [Google Scholar] [CrossRef] [PubMed]
  422. Packer, M.; Narahara, K.A.; Elkayam, U.; Sullivan, J.M.; Pearle, D.L.; Massie, B.M.; Creager, M.A. Double-blind, placebo-controlled study of the efficacy of flosequinan in patients with chronic heart failure. J. Am. Coll. Cardiol. 1993, 22, 65–72. [Google Scholar] [CrossRef] [PubMed]
  423. 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-alpha delivered as a 21-day continuous intravenous infusion in patients with recurrent high-grade astrocytomas. Neuro Oncol. 2005, 7, 32–40. [Google Scholar] [CrossRef] [PubMed]
  424. Paz-Ares, L.; Douillard, J.-Y.; Koralewski, P.; Manegold, C.; Smit, E.F.; Reyes, J.M.; Chang, G.-C.; John, W.J.; Peterson, P.M.; Obasaju, C.K.; et al. Phase III study of gemcitabine and cisplatin with or without aprinocarsen, a protein kinase C-alpha antisense oligonucleotide, in patients with advanced-stage non-small-cell lung cancer. J. Clin. Oncol. 2006, 24, 1428–1434. [Google Scholar] [CrossRef] [PubMed]
  425. Ritch, P.; Rudin, C.M.; Bitran, J.D.; Edelman, M.J.; Makalinao, A.; Irwin, D.; Lilenbaum, R.; Peterson, P.; John, W.J. Phase II study of PKC-alpha antisense oligonucleotide aprinocarsen in combination with gemcitabine and carboplatin in patients with advanced non-small cell lung cancer. Lung Cancer 2006, 52, 173–180. [Google Scholar] [CrossRef]
  426. Advani, R.; Peethambaram, P.; Lum, B.L.; Fisher, G.A.; Hartmann, L.; Long, H.J.; Halsey, J.; Holmlund, J.T.; Dorr, A.; Sikic, B.I. A Phase II trial of aprinocarsen, an antisense oligonucleotide inhibitor of protein kinase C alpha, administered as a 21-day infusion to patients with advanced ovarian carcinoma. Cancer 2004, 100, 321–326. [Google Scholar] [CrossRef]
  427. Tolcher, A.W.; Reyno, L.; Venner, P.M.; Ernst, S.D.; Moore, M.; Geary, R.S.; Chi, K.; Hall, S.; Walsh, W.; Dorr, A.; et al. A randomized phase II and pharmacokinetic study of the antisense oligonucleotides ISIS 3521 and ISIS 5132 in patients with hormone-refractory prostate cancer. Clin. Cancer Res 2002, 8, 2530–2535. [Google Scholar]
  428. Marshall, J.L.; Eisenberg, S.G.; Johnson, M.D.; Hanfelt, J.; Dorr, F.A.; El-Ashry, D.; Oberst, M.; Fuxman, Y.; Holmlund, J.; Malik, S. A phase II trial of ISIS 3521 in patients with metastatic colorectal cancer. Clin. Color. Cancer 2004, 4, 268–274. [Google Scholar] [CrossRef]
  429. Rao, S.; Watkins, D.; Cunningham, D.; Dunlop, D.; Johnson, P.; Selby, P.; Hancock, B.W.; Fegan, C.; Culligan, D.; Schey, S.; et al. Phase II study of ISIS 3521, an antisense oligodeoxynucleotide to protein kinase C alpha, in patients with previously treated low-grade non-Hodgkin’s lymphoma. Ann. Oncol. 2004, 15, 1413–1418. [Google Scholar] [CrossRef]
  430. Propper, D.; Macaulay, V.; O’Byrne, K.; Braybrooke, J.; Wilner, S.; Ganesan, T.; Talbot, D.; Harris, A. A phase II study of bryostatin 1 in metastatic malignant melanoma. Br. J. Cancer 1998, 78, 1337–1341. [Google Scholar] [CrossRef]
  431. Gonzalez, R.; Ebbinghaus, S.; Henthorn, T.K.; Miller, D.; Kraft, A.S. Treatment of patients with metastatic melanoma with bryostatin-1--a phase II study. Melanoma Res. 1999, 9, 599–606. [Google Scholar] [CrossRef] [PubMed]
  432. Bedikian, A.Y.; Plager, C.; Stewart, J.R.; O’Brian, C.A.; Herdman, S.K.; Ross, M.; Papadopoulos, N.; Eton, O.; Ellerhorst, J.; Smith, T. Phase II evaluation of bryostatin-1 in metastatic melanoma. Melanoma Res. 2001, 11, 183–188. [Google Scholar] [CrossRef] [PubMed]
  433. Brockstein, B.; Samuels, B.; Humerickhouse, R.; Arietta, R.; Fishkin, P.; Wade, J.; Sosman, J.; Vokes, E. Phase II studies of bryostatin-1 in patients with advanced sarcoma and advanced head and neck cancer. Investig. New Drugs 2001, 19, 249–254. [Google Scholar] [CrossRef] [PubMed]
  434. Varterasian, M.L.; Pemberton, P.A.; Hulburd, K.; Rodriguez, D.H.; Murgo, A.; Al-Katib, A.M. Phase II study of bryostatin 1 in patients with relapsed multiple myeloma. Investig. New Drugs 2001, 19, 245–247. [Google Scholar] [CrossRef] [PubMed]
  435. Zonder, J.A.; Shields, A.F.; Zalupski, M.; Chaplen, R.; Heilbrun, L.K.; Arlauskas, P.; Philip, P.A. A phase II trial of bryostatin 1 in the treatment of metastatic colorectal cancer. Clin. Cancer Res. 2001, 7, 38–42. [Google Scholar] [PubMed]
  436. Pfister, D.G.; McCaffrey, J.; Zahalsky, A.J.; Schwartz, G.K.; Lis, E.; Gerald, W.; Huvos, A.; Shah, J.; Kraus, D.; Shaha, A.; et al. A phase II trial of bryostatin-1 in patients with metastatic or recurrent squamous cell carcinoma of the head and neck. Investig. New Drugs 2002, 20, 123–127. [Google Scholar] [CrossRef] [PubMed]
  437. Winegarden, J.D.; Mauer, A.M.; Gajewski, T.F.; Hoffman, P.C.; Krauss, S.A.; Rudin, C.M.; Vokes, E.E. A phase II study of bryostatin-1 and paclitaxel in patients with advanced non-small cell lung cancer. Lung Cancer 2003, 39, 191–196. [Google Scholar] [CrossRef]
  438. Nezhat, F.; Wadler, S.; Muggia, F.; Mandeli, J.; Goldberg, G.; Rahaman, J.; Runowicz, C.; Murgo, A.J.; Gardner, G.J. Phase II trial of the combination of bryostatin-1 and cisplatin in advanced or recurrent carcinoma of the cervix: A New York Gynecologic Oncology Group study. Gynecol. Oncol. 2004, 93, 144–148. [Google Scholar] [CrossRef]
  439. Lam, A.P.; Sparano, J.A.; Vinciguerra, V.; Ocean, A.J.; Christos, P.; Hochster, H.; Camacho, F.; Goel, S.; Mani, S.; Kaubisch, A. Phase II study of paclitaxel plus the protein kinase C inhibitor bryostatin-1 in advanced pancreatic carcinoma. Am. J. Clin. Oncol. 2010, 33, 121–124. [Google Scholar] [CrossRef]
  440. Haas, N.B.; Smith, M.; Lewis, N.; Littman, L.; Yeslow, G.; Joshi, I.D.; Murgo, A.; Bradley, J.; Gordon, R.; Wang, H.; et al. Weekly bryostatin-1 in metastatic renal cell carcinoma: A phase II study. Clin. Cancer Res. 2003, 9, 109–114. [Google Scholar]
  441. Madhusudan, S.; Protheroe, A.; Propper, D.; Han, C.; Corrie, P.; Earl, H.; Hancock, B.; Vasey, P.; Turner, A.; Balkwill, F.; et al. A multicentre phase II trial of bryostatin-1 in patients with advanced renal cancer. Br. J. Cancer 2003, 89, 1418–1422. [Google Scholar] [CrossRef] [PubMed]
  442. Ajani, J.A.; Jiang, Y.; Faust, J.; Chang, B.B.; Ho, L.; Yao, J.C.; Rousey, S.; Dakhil, S.; Cherny, R.C.; Craig, C.; et al. A multi-center phase II study of sequential paclitaxel and bryostatin-1 (NSC 339555) in patients with untreated, advanced gastric or gastroesophageal junction adenocarcinoma. Investig. New Drugs 2006, 24, 353–357. [Google Scholar] [CrossRef] [PubMed]
  443. Ku, G.Y.; Ilson, D.H.; Schwartz, L.H.; Capanu, M.; O’reilly, E.; Shah, M.A.; Kelsen, D.P.; Schwartz, G.K. Phase II trial of sequential paclitaxel and 1 h infusion of bryostatin-1 in patients with advanced esophageal cancer. Cancer Chemother. Pharmacol. 2008, 62, 875–880. [Google Scholar] [CrossRef] [PubMed]
  444. Cripps, M.C.; Figueredo, A.T.; Oza, A.M.; Taylor, M.J.; Fields, A.L.; Holmlund, J.T.; McIntosh, L.W.; Geary, R.S.; Eisenhauer, E.A. Phase II randomized study of ISIS 3521 and ISIS 5132 in patients with locally advanced or metastatic colorectal cancer: A National Cancer Institute of Canada clinical trials group study. Clin. Cancer Res. 2002, 8, 2188–2192. [Google Scholar]
  445. Advani, R.; Lum, B.L.; Fisher, G.A.; Halsey, J.; Geary, R.S.; Holmlund, J.T.; Kwoh, T.J.; Dorr, F.A.; Sikic, B.I. A phase I trial of aprinocarsen (ISIS 3521/LY900003), an antisense inhibitor of protein kinase C-alpha administered as a 24-hour weekly infusion schedule in patients with advanced cancer. Investig. New Drugs 2005, 23, 467–477. [Google Scholar] [CrossRef]
  446. Robertson, M.J.; Kahl, B.S.; Vose, J.M.; De Vos, S.; Laughlin, M.; Flynn, P.J.; Rowland, K.; Cruz, J.C.; Goldberg, S.L.; Musib, L.; et al. Phase II study of enzastaurin, a protein kinase C beta inhibitor, in patients with relapsed or refractory diffuse large B-cell lymphoma. J. Clin. Oncol. 2007, 25, 1741–1746. [Google Scholar] [CrossRef]
  447. Oh, Y.; Herbst, R.S.; Burris, H.; Cleverly, A.; Musib, L.; Lahn, M.; Bepler, G. Enzastaurin, an oral serine/threonine kinase inhibitor, as second- or third-line therapy of non-small-cell lung cancer. J. Clin. Oncol. 2008, 26, 1135–1141. [Google Scholar] [CrossRef]
  448. Glimelius, B.; Lahn, M.; Gawande, S.; Cleverly, A.; Darstein, C.; Musib, L.; Liu, Y.; Spindler, K.L.; Frödin, J.E.; Berglund, Å.; et al. A window of opportunity phase II study of enzastaurin in chemonaive patients with asymptomatic metastatic colorectal cancer. Ann. Oncol. 2010, 21, 1020–1026. [Google Scholar] [CrossRef]
  449. Kreisl, T.N.; Kotliarova, S.; Butman, J.A.; Albert, P.S.; Kim, L.; Musib, L.; Thornton, D.; Fine, H.A. A phase I/II trial of enzastaurin in patients with recurrent high-grade gliomas. Neuro Oncol. 2010, 12, 181–189. [Google Scholar] [CrossRef]
  450. Socinski, M.A.; Raju, R.N.; Stinchcombe, T.; Kocs, D.M.; Couch, L.S.; Barrera, D.; Rousey, S.R.; Choksi, J.K.; Jotte, R.; Patt, D.A.; et al. Randomized, phase II trial of pemetrexed and carboplatin with or without enzastaurin versus docetaxel and carboplatin as first-line treatment of patients with stage IIIB/IV non-small cell lung cancer. J. Thorac. Oncol. 2010, 5, 1963–1969. [Google Scholar] [CrossRef]
  451. Couldwell, W.T.; Hinton, D.R.; Surnock, A.A.; DeGiorgio, C.M.; Weiner, L.P.; Apuzzo, M.L.; Masri, L.; Law, R.E.; Weiss, M.H. Treatment of recurrent malignant gliomas with chronic oral high-dose tamoxifen. Clin. Cancer Res. 1996, 2, 619–622. [Google Scholar] [PubMed]
  452. Bergan, R.C.; Reed, E.; Myers, C.E.; Headlee, D.; Brawley, O.; Cho, H.K.; Figg, W.D.; Tompkins, A.; Linehan, W.M.; Kohler, D.; et al. A Phase II study of high-dose tamoxifen in patients with hormone-refractory prostate cancer. Clin. Cancer Res. 1999, 5, 2366–2373. [Google Scholar] [PubMed]
  453. Robins, H.I.; Won, M.; Seiferheld, W.F.; Schultz, C.J.; Choucair, A.K.; Brachman, D.G.; Demas, W.F.; Mehta, M.P. Phase 2 trial of radiation plus high-dose tamoxifen for glioblastoma multiforme: RTOG protocol BR-0021. Neuro Oncol. 2006, 8, 47–52. [Google Scholar] [CrossRef] [PubMed]
  454. 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] [PubMed]
  455. Rini, B.I.; Weinberg, V.; Shaw, V.; Scott, J.; Bok, R.; Park, J.W.; Small, E.J. Time to disease progression to evaluate a novel protein kinase C inhibitor, UCN-01, in renal cell carcinoma. Cancer 2004, 101, 90–95. [Google Scholar] [CrossRef] [PubMed]
  456. Welch, S.; Hirte, H.W.; Carey, M.S.; Hotte, S.J.; Tsao, M.-S.; Brown, S.; Pond, G.R.; Dancey, J.E.; Oza, A.M. UCN-01 in combination with topotecan in patients with advanced recurrent ovarian cancer: A study of the Princess Margaret Hospital Phase II consortium. Gynecol. Oncol. 2007, 106, 305–310. [Google Scholar] [CrossRef] [PubMed]
  457. Zhao, B.; Bower, M.J.; McDevitt, P.J.; Zhao, H.; Davis, S.T.; Johanson, K.O.; Green, S.M.; Concha, N.O.; Zhou, B.-B.S. Structural basis for Chk1 inhibition by UCN-01. J. Biol. Chem. 2002, 277, 46609–46615. [Google Scholar] [CrossRef]
  458. Schwartz, G.K.; Ward, D.; Saltz, L.; Casper, E.S.; Spiess, T.; Mullen, E.; Woodworth, J.; Venuti, R.; Zervos, P.; Storniolo, A.M.; et al. A pilot clinical/pharmacological study of the protein kinase C-specific inhibitor safingol alone and in combination with doxorubicin. Clin. Cancer Res. 1997, 3, 537–543. [Google Scholar]
  459. Ling, L.-U.; Tan, K.-B.; Lin, H.; Chiu, G.N.C. The role of reactive oxygen species and autophagy in safingol-induced cell death. Cell Death Dis. 2011, 2, e129. [Google Scholar] [CrossRef]
  460. Schaar, D.; Goodell, L.; Aisner, J.; Cui, X.X.; Han, Z.T.; Chang, R.; Martin, J.; Grospe, S.; Dudek, L.; Riley, J.; et al. A phase I clinical trial of 12- O-tetradecanoylphorbol-13-acetate for patients with relapsed/refractory malignancies. Cancer Chemother. Pharmacol. 2006, 57, 789–795. [Google Scholar] [CrossRef]
  461. Aiello, L.P.; Vignati, L.; Sheetz, M.J.; Zhi, X.; Girach, A.; Davis, M.D.; Wolka, A.M.; Shahri, N.; Milton, R.C. Oral protein kinase c β inhibition using ruboxistaurin: Efficacy, safety, and causes of vision loss among 813 patients (1,392 eyes) with diabetic retinopathy in the Protein Kinase C β Inhibitor-Diabetic Retinopathy Study and the Protein Kinase C β Inhibitor-Diabetic Retinopathy Study 2. Retina 2011, 31, 2084–2094. [Google Scholar] [PubMed]
  462. PKC-DRS Study Group. The effect of ruboxistaurin on visual loss in patients with moderately severe to very severe nonproliferative diabetic retinopathy: Initial results of the Protein Kinase C beta Inhibitor Diabetic Retinopathy Study (PKC-DRS) multicenter randomized clinical trial. Diabetes 2005, 54, 2188–2197. [Google Scholar]
  463. Aiello, L.P.; Davis, M.D.; Girach, A.; Kles, K.A.; Milton, R.C.; Sheetz, M.J.; Vignati, L.; Zhi, X.E. Effect of ruboxistaurin on visual loss in patients with diabetic retinopathy. Ophthalmology 2006, 113, 2221–2230. [Google Scholar] [PubMed]
  464. Sheetz, M.J.; Aiello, L.P.; Shahri, N.; Davis, M.D.; Kles, K.A.; Danis, R.P.; Mbdv Study Group. Effect of ruboxistaurin (RBX) On visual acuity decline over a 6-year period with cessation and reinstitution of therapy: Results of an open-label extension of the Protein Kinase C Diabetic Retinopathy Study 2 (PKC-DRS2). Retina 2011, 31, 1053–1059. [Google Scholar] [CrossRef] [PubMed]
  465. PKC-DRS2 Group. Effect of ruboxistaurin in patients with diabetic macular edema: Thirty-month results of the randomized PKC-DMES clinical trial. Arch. Ophthalmol. 2007, 125, 318–324. [Google Scholar] [CrossRef] [PubMed]
  466. Tuttle, K.R.; Bakris, G.L.; Toto, R.D.; McGill, J.B.; Hu, K.; Anderson, P.W. The effect of ruboxistaurin on nephropathy in type 2 diabetes. Diabetes Care 2005, 28, 2686–2690. [Google Scholar] [CrossRef] [PubMed]
  467. Tuttle, K.R.; McGill, J.B.; Haney, D.J.; Lin, T.E.; Anderson, P.W. Kidney outcomes in long-term studies of ruboxistaurin for diabetic eye disease. Clin. J. Am. Soc. Nephrol. 2007, 2, 631–636. [Google Scholar] [CrossRef] [PubMed]
  468. Vinik, A.I.; Bril, V.; Kempler, P.; Litchy, W.J.; Tesfaye, S.; Price, K.L.; Bastyr, E.J.; MBBQ Study Group. Treatment of symptomatic diabetic peripheral neuropathy with the protein kinase C beta-inhibitor ruboxistaurin mesylate during a 1-year, randomized, placebo-controlled, double-blind clinical trial. Clin. Ther. 2005, 27, 1164–1180. [Google Scholar] [CrossRef]
  469. Casellini, C.M.; Barlow, P.M.; Rice, A.L.; Casey, M.; Simmons, K.; Pittenger, G.; Bastyr, E.J.; Wolka, A.M.; Vinik, A.I. A 6-month, randomized, double-masked, placebo-controlled study evaluating the effects of the protein kinase C-beta inhibitor ruboxistaurin on skin microvascular blood flow and other measures of diabetic peripheral neuropathy. Diabetes Care 2007, 30, 896–902. [Google Scholar] [CrossRef]
  470. Bates, E.; Bode, C.; Costa, M.; Gibson, C.M.; Granger, C.; Green, C.; Grimes, K.; Harrington, R.; Huber, K.; Kleiman, N.; et al. Intracoronary KAI-9803 as an adjunct to primary percutaneous coronary intervention for acute ST-segment elevation myocardial infarction. Circulation 2008, 117, 886–896. [Google Scholar]
  471. Lincoff, M. Inhibition of delta-Protein Kinase C for Reduction of Infarct Size in Acute Myocardial Infarction—The PROTECTION AMI Trial, KAI. In Proceedings of the American College of Cardiology 60th Annual Scientific Sessions, New Orleans, LA, USA, 2–5 April 2011. [Google Scholar]
  472. Lincoff, A.M.; Roe, M.; Aylward, P.; Galla, J.; Rynkiewicz, A.; Guetta, V.; Zelizko, M.; Kleiman, N.; White, H.; McErlean, E.; et al. Inhibition of delta-protein kinase C by delcasertib as an adjunct to primary percutaneous coronary intervention for acute anterior ST-segment elevation myocardial infarction: Results of the PROTECTION AMI Randomized Controlled Trial. Eur. Heart J. 2014, 35, 2516–2523. [Google Scholar] [CrossRef] [PubMed]
  473. DeMets, D.L.; Califf, R.M. Lessons learned from recent cardiovascular clinical trials: Part II. Circulation 2002, 106, 880–886. [Google Scholar] [CrossRef] [PubMed]
  474. Julier, K.; da Silva, R.; Garcia, C.; Bestmann, L.; Frascarolo, P.; Zollinger, A.; Chassot, P.G.; Schmid, E.R.; Turina, M.I.; von Segesser, L.K.; et al. Preconditioning by sevoflurane decreases biochemical markers for myocardial and renal dysfunction in coronary artery bypass graft surgery: A double-blinded, placebo-controlled, multicenter study. Anesthesiology 2003, 98, 1315–1327. [Google Scholar] [CrossRef] [PubMed]
  475. Guarracino, F.; Landoni, G.; Tritapepe, L.; Pompei, F.; Leoni, A.; Aletti, G.; Scandroglio, A.M.; Maselli, D.; De Luca, M.; Marchetti, C.; et al. Myocardial damage prevented by volatile anesthetics: A multicenter randomized controlled study. J. Cardiothorac. Vasc. Anesth. 2006, 20, 477–483. [Google Scholar] [CrossRef] [PubMed]
  476. Lee, M.C.; Chen, C.H.; Kuo, M.C.; Kang, P.L.; Lo, A.; Liu, K. Isoflurane preconditioning-induced cardio-protection in patients undergoing coronary artery bypass grafting. Eur. J. Anaesthesiol. 2006, 23, 841–847. [Google Scholar] [CrossRef]
  477. Tritapepe, L.; Landoni, G.; Guarracino, F.; Pompei, F.; Crivellari, M.; Maselli, D.; De Luca, M.; Fochi, O.; D’Avolio, S.; Bignami, E.; et al. Cardiac protection by volatile anaesthetics: A multicentre randomized controlled study in patients undergoing coronary artery bypass grafting with cardiopulmonary bypass. Eur. J. Anaesthesiol. 2007, 24, 323–331. [Google Scholar] [CrossRef]
  478. De Hert, S.; Vlasselaers, D.; Barbé, R.; Ory, J.; Dekegel, D.; Donnadonni, R.; Demeere, J.; Mulier, J.; Wouters, P. A comparison of volatile and non volatile agents for cardioprotection during on-pump coronary surgery. Anaesthesia 2009, 64, 953–960. [Google Scholar] [CrossRef]
  479. Mentzer, R.M., Jr.; Birjiniuk, V.; Khuri, S.; Lowe, J.E.; Rahko, P.S.; Weisel, R.D.; Wellons, H.A.; Barker, M.L.; Lasley, R.D.; Adenosine Myocardial Protection Investigators. Adenosine myocardial protection: Preliminary results of a phase II clinical trial. Ann. Surg. 1999, 229, 643–650. [Google Scholar] [CrossRef]
  480. Belhomme, D.; Peynet, J.; Florens, E.; Tibourtine, O.; Kitakaze, M.; Menasché, P. Is adenosine preconditioning truly cardioprotective in coronary artery bypass surgery? Ann. Thorac. Surg. 2000, 70, 590–594. [Google Scholar] [CrossRef]
  481. Jin, Z.; Duan, W.; Chen, M.; Yu, S.; Zhang, H.; Feng, G.; Xiong, L.; Yi, D. The myocardial protective effects of adenosine pretreatment in children undergoing cardiac surgery: A randomized controlled clinical trial. Eur. J. Cardiothorac. Surg. 2011, 39, e90–e96. [Google Scholar] [CrossRef]
  482. Mangano, D.T.; Miao, Y.; Tudor, I.C.; Dietzel, C. Post-Reperfusion Myocardial Infarction: Long-Term Survival Improvement Using Adenosine Regulation with Acadesine. J. Am. Coll. Cardiol. 2006, 48, 206–214. [Google Scholar] [CrossRef] [PubMed]
  483. Newman, M.F.; Ferguson, T.B.; White, J.A.; Ambrosio, G.; Koglin, J.; Nussmeier, N.A.; Pearl, R.G.; Pitt, B.; Wechsler, A.S.; Weisel, R.D.; et al. Effect of adenosine-regulating agent acadesine on morbidity and mortality associated with coronary artery bypass grafting: The RED-CABG randomized controlled trial. JAMA 2012, 308, 157–164. [Google Scholar] [CrossRef] [PubMed]
  484. Drew, B.G.; Kingwell, B.A. Acadesine, an adenosine-regulating agent with the potential for widespread indications. Expert. Opin. Pharmacother. 2008, 9, 2137–2144. [Google Scholar] [CrossRef] [PubMed]
  485. Ahmad, A.; Sheikh, S.; Khan, M.A.; Chaturvedi, A.; Patel, P.; Patel, R.; Buch, B.C.; Anand, R.S.; Shah, T.C.; Vora, V.N.; et al. Endoxifen: A new, protein kinase C inhibitor to treat acute and mixed mania associated with bipolar I disorder. Bipolar Disord. 2021, 23, 595–603. [Google Scholar] [CrossRef] [PubMed]
  486. Cousins, M.J.; Pickthorn, K.; Huang, S.; Critchley, L.; Bell, G. The safety and efficacy of KAI-1678- an inhibitor of epsilon protein kinase C (epsilonPKC)-versus lidocaine and placebo for the treatment of postherpetic neuralgia: A crossover study design. Pain Med. 2013, 14, 533–540. [Google Scholar] [CrossRef]
  487. Moodie, J.E.; Bisley, E.J.; Huang, S.; Pickthorn, K.; Bell, G. A single-center, randomized, double-blind, active, and placebo-controlled study of KAI-1678, a novel PKC-epsilon inhibitor, in the treatment of acute postoperative orthopedic pain. Pain. Med. 2013, 14, 916–924. [Google Scholar]
  488. Shoushtari, A.N.; Khan, S.; Komatsubara, K.; Feun, L.; Acquavella, N.; Singh-Kandah, S.; Negri, T.; Nesson, A.; Abbate, K.; Cremers, S.; et al. A Phase Ib Study of Sotrastaurin, a PKC Inhibitor, and Alpelisib, a PI3Kα Inhibitor, in Patients with Metastatic Uveal Melanoma. Cancers 2021, 13, 5504. [Google Scholar] [CrossRef]
  489. Scott, D.; Mazurkiewicz, M.; Leeman, P. The long-term monitoring of ventilation rhythms of the polychaetous annelid Nereis virens sars. Comp. Biochem. Physiol. A Comp. Physiol. 1976, 53, 65–68. [Google Scholar] [CrossRef]
  490. Jacob, J.S.; McDonald, H.S. Diving bradycardia in four species of North American aquatic snakes. Comp. Biochem. Physiol. A Comp. Physiol. 1976, 53, 69–72. [Google Scholar] [CrossRef]
  491. Pascher, A.; De, P.S.; Pratschke, J.; Salamé, E.; Pirenne, J.; Isoneimi, H.; Bijarnia, M.; Krishnan, I.; Klupp, J. Protein kinase C inhibitor sotrastaurin in de novo liver transplant recipients: A randomized phase II trial. Am. J. Transplant. 2015, 15, 1283–1292. [Google Scholar] [CrossRef]
  492. Farlow, M.R.; Thompson, R.E.; Wei, L.J.; Tuchman, A.J.; Grenier, E.; Crockford, D.; Wilke, S.; Benison, J.; Alkon, D.L. A Randomized, Double-Blind, Placebo-Controlled, Phase II Study Assessing Safety, Tolerability, and Efficacy of Bryostatin in the Treatment of Moderately Severe to Severe Alzheimer’s Disease. J. Alzheimers Dis. 2019, 67, 555–570. [Google Scholar] [CrossRef] [PubMed]
Figure 1. PKC isozyme diversity. (A) The human kinome (adapted from www.cellsignal.com/reference/kinase) phylogeny is extensive, and the PKC family is one member of the AGC superfamily (bottom right), with three groups of conventional, novel, and atypical PKC isozymes. (B) PKC isozymes are homologous but contain a distinct set of structural domains responsible for their diverse functions and interactions with second messengers and other binding partners. All PKC family members are constituted by four conserved domains (C1–C4) separated by a hinge region. The pseudo-substrate site (PS) keeps the protein in its inactive form. Second messengers are indicated on the right side of the picture. BioRender.com was used to generate this figure.
Figure 1. PKC isozyme diversity. (A) The human kinome (adapted from www.cellsignal.com/reference/kinase) phylogeny is extensive, and the PKC family is one member of the AGC superfamily (bottom right), with three groups of conventional, novel, and atypical PKC isozymes. (B) PKC isozymes are homologous but contain a distinct set of structural domains responsible for their diverse functions and interactions with second messengers and other binding partners. All PKC family members are constituted by four conserved domains (C1–C4) separated by a hinge region. The pseudo-substrate site (PS) keeps the protein in its inactive form. Second messengers are indicated on the right side of the picture. BioRender.com was used to generate this figure.
Ijms 24 17600 g001
Figure 2. PKC maturation, activation, and cytosolic fate. The open PKC matures via a priming process involving three critical phosphorylation events, which is facilitated by chaperone complex proteins. PKC adopts a closed, autoinhibited, conformation in which the pseudo-substrate (PS) domain is trapped in the substrate-binding site, and the primed enzyme localizes to the cytosol. The closed mature PKC in the cytosol is then activated by second messengers and scaffold proteins that mediate the target specificity of the kinase. The cytosolic survival of the species and the fate of the protein are determined by other scaffolds and post-translational modifications that govern downregulation and degradation reactions. BioRender.com was used to generate this figure.
Figure 2. PKC maturation, activation, and cytosolic fate. The open PKC matures via a priming process involving three critical phosphorylation events, which is facilitated by chaperone complex proteins. PKC adopts a closed, autoinhibited, conformation in which the pseudo-substrate (PS) domain is trapped in the substrate-binding site, and the primed enzyme localizes to the cytosol. The closed mature PKC in the cytosol is then activated by second messengers and scaffold proteins that mediate the target specificity of the kinase. The cytosolic survival of the species and the fate of the protein are determined by other scaffolds and post-translational modifications that govern downregulation and degradation reactions. BioRender.com was used to generate this figure.
Ijms 24 17600 g002
Table 1. PKC isozymes in human disease.
Table 1. PKC isozymes in human disease.
DiseasePKC IsozymeAssociated Pathology
CancerPKCαProliferation; invasion; metastasis; RCC tumor progression; mutated in thyroid cancer; SCC tumorigenesis and progression; elevated in late-stage breast cancer; upregulated in lung cancer; activated in bladder cancer
PKCβAngiogenesis; invasion; elevated in early-stage breast cancer; downregulated in CRC; PKCβII tumor suppressor in CRC; promotes angiogenesis in NHLs and DLBCL
PKCδAngiogenesis; apoptosis in endometrial cancer; elevated in early-stage breast cancer; downregulated in CRC
PKCγIncreased in CRC
PKCεOncogene; proliferation; metastasis; chemotherapy resistance; mutated in thyroid cancer; SCC tumorigenesis, progression, and metastasis; keratinocyte cytoskeleton structure; upregulated in lung cancer; downregulated in CRC
PKCηGlioblastoma tumorigenesis; proliferation; radiation resistance; elevated in early-stage breast cancer; upregulated in lung cancer; downregulated in CRC
PKCθGastrointestinal stromal tumorigenesis
PKCι/λOncogene; upregulated in lung cancer; promotes prostate cancer invasion; conflicting roles in prostate cancer metastasis; elevation associated with poor outcomes in PDAC
PKCζElevated in late-stage breast cancer; upregulated in lung cancer; promotes RCC tumor progression
Diabetes mellitusPKCβObesity; glucose transport; insulin resistance; cholesterol and fatty acid metabolism; PKCβII late-stage vascular complications
PKCδIslet cell function
Ischemic heart diseasePKCβPathologic postischemic cardiac remodeling
PKCδROS production; apoptosis; necrosis; postischemic injury
PKCεMitochondrial function; proteasome function; ALDH2 activation
Heart failurePKCαReduced contractility; β-adrenergic receptor uncoupling
PKCβIIDecreased proteasome function; calcium dysregulation; conflicting roles in hypertrophy; conflicting roles in contractility
PKCδAbnormal activation in early HF
PKCεFibrosis; inflammation; depleted in late-stage HF
PKCζGeneration of active NADPH oxidase; Reduced basal and TGF-βI induced fibroblast proliferation; increased MMP1,3,9 release
HypertensionPKCαElevated in hypertension
PKCβElevated in hypertension
PKCδElevated in hypertension
PKCεTranslocation and increased hypertrophic cellular growth
StrokePKCδMitochondrial fission; ROS production; blood–brain barrier dysfunction
PKCεCytoprotective; cerebral perfusion
NeurodegenerationPKCδInflammation; neuronal cell death in Parkinson’s; β-amyloid formation in Alzheimer’s disease
PKCα Associated with AD, promoting synaptic loss
PKCγ Associated with SCA, promoting neuronal death
Bipolar disorderPKCαAbnormal gene expression
PKCεAbnormal neuronal transmission
NociceptionPKCαPeripheral nociception signaling
PKCγDorsal root ganglia signal transmission
PKCεSpinal cord signal transmission; peripheral nociception signaling
InflammationPKCδB-cell development; keratinocyte proliferation and dysregulated angiogenesis in psoriasis; eosinophil activation in asthma
PKCθT-cell responses; airway inflammation; joint inflammation
PKCζLung inflammation
RCC, renal cell carcinoma; HF, heart failure; SCC, squamous cell carcinoma; CRC, colorectal cancer; PDAC, pancreatic ductal adenocarcinoma; NHL, non-Hodgkin’s lymphoma; DLBCL, diffuse large B cell lymphoma; AD, Alzheimer’s disease; SCA, Spinocerebellar ataxia. Adapted from [89,101,211,382,383,384].
Table 2. Prototypical PKC inhibitors used in the clinic.
Table 2. Prototypical PKC inhibitors used in the clinic.
Compound NameStructureIndication(s)Target DomainMechanismIsozyme Selectivity
MidostaurinIjms 24 17600 i001Acute myeloid leukemiaCatalytic domainCompetitive inhibitor of ATP binding siteMultiple protein kinases
RiluzoleIjms 24 17600 i002Amyotrophic lateral sclerosisCatalytic domainCompetitive inhibitor of ATP binding sitePKCα and other kinases
CurcuminIjms 24 17600 i003Not approved; used as nutraceutical for ulcerative colitis and other inflammatory conditionsRegulatory domainC1 inhibitorPKCα
ResveratrolIjms 24 17600 i004Not approved; ongoing studies in cancer, cardiovascular disease, neurodegeneration, and agingRegulatory domainC1 inhibitorPKCα; PKCβ; PKCε
Adapted from [101,382].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Silnitsky, S.; Rubin, S.J.S.; Zerihun, M.; Qvit, N. An Update on Protein Kinases as Therapeutic Targets—Part I: Protein Kinase C Activation and Its Role in Cancer and Cardiovascular Diseases. Int. J. Mol. Sci. 2023, 24, 17600. https://doi.org/10.3390/ijms242417600

AMA Style

Silnitsky S, Rubin SJS, Zerihun M, Qvit N. An Update on Protein Kinases as Therapeutic Targets—Part I: Protein Kinase C Activation and Its Role in Cancer and Cardiovascular Diseases. International Journal of Molecular Sciences. 2023; 24(24):17600. https://doi.org/10.3390/ijms242417600

Chicago/Turabian Style

Silnitsky, Shmuel, Samuel J. S. Rubin, Mulate Zerihun, and Nir Qvit. 2023. "An Update on Protein Kinases as Therapeutic Targets—Part I: Protein Kinase C Activation and Its Role in Cancer and Cardiovascular Diseases" International Journal of Molecular Sciences 24, no. 24: 17600. https://doi.org/10.3390/ijms242417600

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