Next Article in Journal
Chemoprevention of Urothelial Cell Carcinoma Tumorigenesis by Dietary Flavokawain A in UPII-Mutant Ha-ras Transgenic Mice
Next Article in Special Issue
The Role of the Kinase Inhibitors in Thyroid Cancers
Previous Article in Journal
Escinosomes: Safe and Successful Nanovesicles to Deliver Andrographolide by a Subcutaneous Route in a Mice Model of Oxaliplatin-Induced Neuropathy
Previous Article in Special Issue
BCR-ABL1 Tyrosine Kinase Complex Signaling Transduction: Challenges to Overcome Resistance in Chronic Myeloid Leukemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Choosing Kinase Inhibitors for Androgen Deprivation Therapy-Resistant Prostate Cancer

1
Department of General Surgery, Xiangya Hospital, Central South University, Hunan 410008, China
2
Department of Molecular Medicine, The Scripps Research Institute, Jupiter, FL 33459, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2022, 14(3), 498; https://doi.org/10.3390/pharmaceutics14030498
Submission received: 1 November 2021 / Revised: 26 January 2022 / Accepted: 22 February 2022 / Published: 24 February 2022
(This article belongs to the Special Issue Cancer Therapy Resistance: Choosing Kinase Inhibitors)

Abstract

:
Androgen deprivation therapy (ADT) is a systemic therapy for advanced prostate cancer (PCa). Although most patients initially respond to ADT, almost all cancers eventually develop castration resistance. Castration-resistant PCa (CRPC) is associated with a very poor prognosis, and the treatment of which is a serious clinical challenge. Accumulating evidence suggests that abnormal expression and activation of various kinases are associated with the emergence and maintenance of CRPC. Many efforts have been made to develop small molecule inhibitors to target the key kinases in CRPC. These inhibitors are designed to suppress the kinase activity or interrupt kinase-mediated signal pathways that are associated with PCa androgen-independent (AI) growth and CRPC development. In this review, we briefly summarize the roles of the kinases that are abnormally expressed and/or activated in CRPC and the recent advances in the development of small molecule inhibitors that target kinases for the treatment of CRPC.

1. Introduction

Kinases are part of the large family of phosphotransferase, which catalyze the transfer of a phosphate moiety from a high energy molecule (such as ATP) to their substrate molecule. More than five hundred different kinases have been identified in humans, the majority of which phosphorylate proteins while some kinases act on small molecules such as lipids, carbohydrates, amino acids, and nucleotides. For example, androgen receptor (AR) is a substrate of PIM1 and activated CDC42-associated kinase-1 (ACK1), lipid is the substrate of phosphatidylinositol 4-phosphate 5-kinase type 1α (PIP5K1α), and fructose-6-phosphate is one of the substrates of phosphofructokinase (PFK). Based on their substrates, kinases are classified into protein kinases, lipid kinases, and carbohydrate kinases [1]. Protein kinases phosphorylate the amino acid serine, threonine, tyrosine, or histidine residues of the proteins. A specific protein kinase, such as IκB kinase and TANK-binding Kinase 1 (TBK1), normally has more than one substrate, and a specific protein, such as twist-related protein 1 (TWIST1), can serve as a substrate for multiple kinases. Kinases are fundamental molecules in both normal and cancer cells, and are involved in almost all cellular activities [2]. The kinases-mediated protein post-translational modifications function as key mediators or switches in many critical cellular signal pathways. For instance, NEK6 interacts with STAT3 and induces STAT3-mediated transcriptional activation through phosphorylating STAT3 [3]. Kinase abnormal activation and dysregulation have been associated with many diseases, such as autoimmune diseases and cancer.
Prostate cancer (PCa) is the most common malignancy, and the second-leading cause of cancer-related mortality in men in Western countries [4]. In tumors confined to the prostate, radical prostatectomy and radiotherapy are effective; however, for late stage disseminated disease, current therapies are merely palliative. Androgen deprivation therapy (ADT) is a systemic therapy for advanced PCa. Although a majority of patients initially respond to ADT, the responses are transient and almost all cancers eventually develop castration resistance [4,5,6,7,8,9]. Castration-resistant PCa (CRPC) is associated with a very poor prognosis, and the treatment of which is a serious clinical challenge [4,5,6,7]. Abnormal expression and activation of many kinases have been associated with PCa development, progression, as well as the emergence and maintenance of CRPC. Recently, many kinases, such as polo-like kinase (PLK1), aurora kinase A (AURKA), glycogen-synthase-kinase 3 (GSK-3), and PIP5K1α have been identified to participate in CRPC progression through regulating their substrates and/or downstream signal pathways, targeting these kinases with inhibitors can significantly interrupt the pathogenesis of CRPC (Table 1). For instance, LIM-domain kinase-2 (LIMK2) is up-regulated after ADT and promotes CRPC development through phosphorylating TWIST1 in CRPC cells. Knockdown of LIMK2 significantly suppresses CRPC development [10]. As many kinases play important roles in CRPC growth and development, kinase targeted therapies have been developed for the treatment of CRPC.
Small molecule inhibitors are attractive anti-tumor agents due to their small size, well-penetrative features, and the potential to be engineered for oral administration [11,12]. Recently, many small molecule inhibitors have been developed to specifically or mainly target the key kinases involved in CRPC emergence and maintenance, some of them have been approved by U.S. Food and Drug Administration (FDA) [11,12]. These small molecules are designed to inhibit kinase activity or interrupt kinase-mediated signal pathways that are associated with PCa androgen-independent (AI) growth, progression, and metastasis (Figure 1 and Figure 2, Table 1). Notably, some of these kinase inhibitors have been applied in clinical trials (Table 2). Here, we briefly review the roles of the key kinases in CRPC development and summarize the recent advances in the development of small molecule inhibitors that target kinases for the treatment of CRPC.
Table 1. Inhibitors target key kinases in CRPC.
Table 1. Inhibitors target key kinases in CRPC.
KinaseRole in CRPCInhibitorChemical StructureMechanism of InhibitionReference
PLK1PLK1 signaling is one of the most up-regulated pathways after androgen deprive in mouse xenograft. PLK1 activation promotes CRPC development through affecting the PI3K–AKT–mTOR pathway and androgen receptor (AR) signaling.BI2536 (combined with olaparib) Pharmaceutics 14 00498 i001Since monotherapy with olaparib leads to high expression of PLK1, PLK1 inhibitor can significantly increase the efficacy of olaparib.[13]
GSK461364A (combined with BRD4 inhibitor, JQ1) Pharmaceutics 14 00498 i002Inhibits PLK1, c-MYC expression and AR signaling[14]
WNY0824 Pharmaceutics 14 00498 i003Inhibit AR transcriptional program, down-regulate MYC and induce mitotic abnormality[15]
BRD4/BETBind to SEs that drive high expression of oncogenes; BRD4 together with HOXB13 regulated transcriptional networks to support androgen-independent cell proliferation of CRPCJQ1 (combined with PLK1 inhibitor, GSK461364A) Pharmaceutics 14 00498 i004Inhibits PLK1, c-MYC expression and AR signaling; down-regulating SE-associated genes, including genes involved in migration and invasion, especially a long non-coding RNA, MANCR[14,16]
MA4-022-1 compound Pharmaceutics 14 00498 i005Interfere BRD4-HOXB13-HOTBIN10 regulatory circuit[17]
(R)-12 (Y02234) Pharmaceutics 14 00498 i006Bind to BRD4 and block the interaction between bromodomain and acetyl lysine; suppress ERG, Myc, and AR signaling[18]
6i (Y06036) and 7m (Y06137)6i Pharmaceutics 14 00498 i007Bind to BRD4 and suppress expression of MYC and AR regulated genes[19]
7m Pharmaceutics 14 00498 i008
ABBV-075 Pharmaceutics 14 00498 i009Disrupt the recruitment of BRD4 to gene-regulatory regions co-occupied by AR, including PSA and TMPRSS2 enhancers, leading to the inhibition of transcription of AR target genes[20]
Compound 7d Pharmaceutics 14 00498 i010Disrupt the functions of AR and BET via binding to AR and suppressing transactivation of AR as well as AR mutant[21]
LIMK2LIMK2 involves in CRPC progression through increasing TWIST1 mRNA levels and stabilizing TWIST1 by phosphorylation, mediating SPOP degradationLIMK2 allosteric inhibitor (sulfonamide 2) (combined with docetaxel) Pharmaceutics 14 00498 i011Is non-ATP competitive (allosteric inhibition) and binds in the DFG-out orientation of kinase[10,22]
MEKMEK1/2 promotes CRPC progression via regulating their downstream target such as ERK1Selumetinib (AZD6244) (combined with ricolinostat (ACY1215)) Pharmaceutics 14 00498 i012Decrease AR-dependent gene (KLK2, DUSP1) mRNA levels; increase AR cytoplasmatic expression[23,24]
AURKAAURKA is overexpressed in CRPC, regulates androgen receptor variants (AR-Vs) expression, and acts as a phosphatase and functions through regulating its substrates, including YBX1, TWIST1, and LIMK2.MLN8237 (alisertib) Pharmaceutics 14 00498 i013Inhibit the growth of CRPC cells with high levels of AR[7]
S1451 (TC-S 7010) Pharmaceutics 14 00498 i014Decrease AR-V7 levels and markedly reduce AR-V-driven proliferation and survival of CRPC cells[25]
TOPKTOPK drives androgen-independent growth in prostate cancer cells (LNCaP and VCaP) via enhancing androgen receptor splice variant (ARv7)OTS-514 Pharmaceutics 14 00498 i015Repress AR transactivation, and AR stability[26]
GSK-3Promote AR-V7 transcriptionLY 2090314 Pharmaceutics 14 00498 i016Activate β-catenin signaling, which suppressed AR-V7 transcriptional activity[27]
CDK9A cofactor for AR, MYC and other oncogenic transcription factorsKB-0742 Pharmaceutics 14 00498 i017Down-regulate nascent transcription, especially AR-driven oncogenic programs and short half-life transcripts[28]
CDK4Regulate cell cycle progressionPso (3, 9-dihydroxy-2-prenylco- umestan (psoralidin)) Pharmaceutics 14 00498 i018Induce G0/G1 cell cycle arrest and cell growth inhibition in CRPC cells[29]
CDK7Regulate MED1-mediated, AR-dependent oncogenic transcriptional amplificationTHZ1 Pharmaceutics 14 00498 i019Interfere the recruitment of MED1 to chromatin and further suppress AR target gene expression[30,31]
IκB kinase (IKK)Phosphorylate IκBα, resulting in the degradation of IκBα and release of p50/p65. The p50/p65 translocates into nucleus to start the transcription of NF-κB-regulating genesUrsolic acid Pharmaceutics 14 00498 i020Inhibit IKK activation and phosphorylation of IκBα[32]
Apigenin Pharmaceutics 14 00498 i021Inhibit IKKα kinase activity, and IκBα phosphorylation and degradation[33]
α-tomatine Pharmaceutics 14 00498 i022Inhibit IκB kinase-mediated IκBα phosphorylation, IκBα degradation and p50/p65 nuclear translocation[34]
PKAPKA activation phosphorylates heat shock protein (HSP90), which binds to the unliganded AR in cytoplasm, and release AR from HSP90. The free AR binds to HSP27 and sequentially migrates into the nucleusH89 Pharmaceutics 14 00498 i023Inhibit the PKA activity and AR translocation[35]
PIM1Modulate AR stability and transcriptional activity through phosphorylating ARCompound 9 (DHPCC-9) and its pyrrolo[2,3-a]carbazole derivatives Pharmaceutics 14 00498 i024
(compound 9 R=H; compound 28 R=Br)
Inhibit the PIM1 kinase activity[36]
DHPCC-9 (1,10-dihydropyrrolo[2,3-a]carbazole-3-carbaldehyde) Pharmaceutics 14 00498 i025Inhibit angiogenesis and lymphangiogenesis as well as phosphorylation of the CXCR4 chemokine receptor[37]
Ack1Phosphorylate AR-tyrosine 267AIM-100 (4-amino-5,6-biaryl-furo[2,3-d]pyrimidine) Pharmaceutics 14 00498 i026Inhibit Ack1 tyrosine kinase activity and suppress Ack1 mediated ATM[38]
JAK2Phosphorylate Stat5a/b, regulate its dimerization, nuclear translocation, DNA binding and transcriptional activityβ-elemonic acid (β-EA) Pharmaceutics 14 00498 i027Suppress of JAK2/STAT3/MCL-1 and NF-ĸB signaling pathway[39]
EGFRPromote CRPC progression through regulating its downstream signaling; mediated docetaxel resistance in human CRPC via Akt-dependent expression of ABCB1Gefitinib Pharmaceutics 14 00498 i028Inhibit EGFR activity[40]
HER2Form EGFR/HER2 heterodimerization and regulate downstream signaling to promote CRPCdacomitinib Pharmaceutics 14 00498 i029Decrease HER2 protein stability and prevent EGFR/HER2 heterodimerization, therefore interrupting downstream signaling and increasing CRPC cell apoptosis[41]
RETIs up-regulated in NEPC, and RET expression correlated with neuroendocrine transcription factors, including POU3F2, SOX2, ONECUT2 and ASCL1AD80 Pharmaceutics 14 00498 i030block RET signaling[42]
PIP5K1αPIP5K1α produces a substrate of PI3K, PIP2, which is required for the activation of PI3K/AKT pathwayISA-2011B Pharmaceutics 14 00498 i031Inhibit AKT activity, reduce invasion, increase cancer cell apoptosis and
inhibit tumor growth in CRPC model
[43]
SK1High levels of SK1 have been identified, and SK1 level and activity are associated with prostate cancer progression, recurrence and chemoresistanceFTY720 (fingolimod) Pharmaceutics 14 00498 i032Activate caspase-3 and induce apoptosis[44,45,46]
SK2Regulate expression of the oncogene c-Myc and promote PCa progressionABC294640 Pharmaceutics 14 00498 i033Reduce CRPC cell growth and the expression of c-Myc and AR; inhibit dihydroceramide desaturase (DEGS), resulting in the increase of dihydroceramides[47]
HK2Increased HK2 expression and activity is associated with CRPC development, especially PTEN- and TP53-deficiency-driven CRPC2-deoxyglucose (2-DG) Pharmaceutics 14 00498 i034Cause AMPK phosphorylation, leading to inhibit mTORC1-S6K1 translation signaling and sequentially block anti-apoptotic protein MCL-l synthesis[48]
Table 2. Kinase inhibitors for CRPC in clinical trials.
Table 2. Kinase inhibitors for CRPC in clinical trials.
InhibitorChemical StructureTargetClinical Trial PhaseModel System/Patient CharacteristicsTreatmentResultsReference
Alisertib Pharmaceutics 14 00498 i035AURKAPhase II (NCT01799278)Patients with castration-resistant and neuroendocrine prostate cancerAlisertibSome but not all patients with NEPC with Aurora-A and N-Myc activation do benefit from alisertib.[49]
AlisertibAURKAI/II (NCT01848067)Patients with metastatic castration-resistant prostate cancerAlisertib in combination with abiraterone and prednisoneNo clear signal indicates that alisertib might be beneficial for patients with mCRPC progressing on abiraterone[50]
AlisertibAURKAI (NCT01094288)Patients with solid tumors, including CRPCAlisertib combines with docetaxel1 complete response in a patient with bladder cancer, six partial responses in patients with castration-resistant prostate cancer, and 1 partial response in a patient with angiosarcoma.[51]
AZD5363
(Capivasertib)
Pharmaceutics 14 00498 i036AKTI (NCT02121639)Patients with metastatic castration-resistant prostate cancerAZD5363 combines with docetaxel and prednisolone chemotherapyPSA reduction to <50% at 12 weeks occurred in seven patients[52]
AZD5363
(Capivasertib)
AKTI (NCT02525068)mCRPC patients who previously failed abiraterone and/or enzalutamideAZD5363 combines with enzalutamideDecrease plasma exposure of capivasertib; Three patients show the criteria for response[53]
BKM120
(Buparlisib)
Pharmaceutics 14 00498 i037PI3KII (NCT02035124)Patients with metastatic castration-resistant prostate cancercabazitaxel plus BKM120Withdrawn due to slow accrual and no patients being enrolledhttps://clinicaltrials.gov/ct2/show/NCT02035124?term=PI3K+inhibitor&cond=CRPC&draw=2&rank=1 (30 October 2021)
AZD8186 Pharmaceutics 14 00498 i038PI3KbetaI (NCT03218826)Patients with advanced solid tumors, including advanced prostate carcinomaAZD8186 plus docetaxelExpected completion: 1 April 2022. Outcome measure: Maximum tolerated dose (MTD) and recommended phase 2 dose (RP2D); incidence of adverse eventshttps://clinicaltrials.gov/ct2/show/NCT03218826?term=PI3K+inhibitor&cond=CRPC&draw=2&rank=3 (30 October 2021)
LY3023414
(Samotolisib)
Pharmaceutics 14 00498 i039PI3K/mTORII (NCT02407054)Patients with prostate cancer, including metastatic castration-resistant prostate cancerEnzalutamide plus LY3023414Outcome measure: progression free survival (PFS); time to disease progression (TTP)https://clinicaltrials.gov/ct2/show/NCT02407054?term=PI3K+inhibitor&cond=CRPC&draw=2&rank=4 (30 October 2021)
ZEN003694This compound is under clinical trial and its structure is unpublished.BET BromodomainII (NCT04471974)Patients with metastatic castration-resistant prostate cancerZEN003694 plus enzalutamide and pembrolizumabExpected completion: 31 December 2025. Outcome measure: composite response rate; progression-free survival (PFS); PSA50 response proportionhttps://clinicaltrials.gov/ct2/show/NCT04471974?term=BET+inhibitor&cond=CRPC&draw=2&rank=1 (30 October 2021)
ZEN003694BET BromodomainIb/IIa (NCT02711956)Patients with metastatic castration-resistant prostate cancerZEN003694 in combination with enzalutamideOutcome measure: dose escalation and dose confirmation; PSA response ratehttps://clinicaltrials.gov/ct2/show/NCT02711956?term=BET+inhibitor&cond=CRPC&draw=2&rank=3 (30 October 2021)
Birabresib (MK-8628/OTX015) Pharmaceutics 14 00498 i040BET BromodomainIb (NCT02259114)Patients with advanced solid tumors, including castration-resistant prostate carcinomaBirabresibHas dose-proportional exposure and a favorable safety profile; recommended phase II with a dose of 80 mg once daily with continuous dosing[54]
Palbociclib Pharmaceutics 14 00498 i041CDK4/6II (NCT02905318)Patients with metastatic castration-resistant prostate cancerPalbociclibExpected completion: 21 May 2022. Outcome measure: clinical benefit rate; PSA decline ≥ 50%https://clinicaltrials.gov/ct2/show/NCT02905318?term=CDK+inhibitor&cond=CRPC&draw=2&rank=3 (30 October 2021)
Trametinib Pharmaceutics 14 00498 i042MEK1/2II (NCT02881242)Patients with metastatic castration-resistant prostate cancerTrametinibExpected completion: 31 January 2023. Outcome measure: PSA response rate; response rate assessed by RECIST criteria; overall survival.https://clinicaltrials.gov/ct2/show/NCT02881242 (30 October 2021)
CEP-11981 (ESK981) Pharmaceutics 14 00498 i043Tyrosine kinaseII (NCT04159896)Patients with metastatic castration-resistant prostate cancerESK981 plus nivolumabExpected completion: 1 March 2022. Outcome measure: the prostate specific antigen (PSA) >= 50% response rate; the safety and tolerability of ESK981 plus nivolumabhttps://clinicaltrials.gov/ct2/show/NCT04159896?term=Tyrosine+kinase+inhibitor&cond=CRPC&draw=2&rank=1 (30 October 2021)
CEP-11981 (ESK981)Tyrosine kinaseII (NCT03456804)Patients with metastatic castration-resistant prostate cancerESK981Expected completion: 31 October 2022. Outcome measure: PSA decline of >=50% (PSA50); PSA progression free survival (PFS); Time to PSA responsehttps://clinicaltrials.gov/ct2/show/NCT03456804?term=Tyrosine+kinase+inhibitor&cond=CRPC&draw=2&rank=2 (30 October 2021)
masitinib Pharmaceutics 14 00498 i044Tyrosine kinaseIII (NCT03761225)Patients with metastatic castration-resistant prostate cancerMasitinib plus docetaxelOutcome measure: progression free survival (PFS); overall survival (OS)https://clinicaltrials.gov/ct2/show/NCT03761225?term=Tyrosine+kinase+inhibitor&cond=CRPC&draw=2&rank=3 (30 October 2021)
Cabozantinib (XL184) Pharmaceutics 14 00498 i045Tyrosine kinaseI (NCT01683994)Patients with metastatic castration-resistant prostate cancerCabozantinib plus docetaxel and prednisoneThe median time to progression and overall survival time were 13.6 and 16.3 months; cabozantinib is safely added to docetaxel/ prednisone with possible enhanced efficacy[55]
lapatinib Pharmaceutics 14 00498 i046HER2II (NCT00246753)Patients with castration-resistant prostate cancerlapatinibShow single agent activity in CRPC patients as evaluated by PSA[56]

2. Serine/Threonine Kinase and Their Therapeutic Implications

2.1. IκB Kinases (IKKs)

NF-κB inflammatory signaling plays an important role in immune responses and survival of both normal and malignant cells. There are two traditional NF-κB activation pathways: the classical NF-κB activation pathway and the alternative NF-κB activation pathway. The classical pathway is triggered by the activation of the tripartite IKK complex composed of IKKα, IKKβ and IKKγ/NEMO, while the alternative pathway is mediated by IKKα-induced p100 processing [57,58,59]. Accumulating evidence suggests that NF-κB signaling in both PCa cells and immune/inflammatory cells plays an important role in the emergence and maintenance of CRPC. In PCa cells, NF-κB inflammatory signaling is constitutively activated to facilitate PCa cell AI growth and survival by up-regulating the expression of survival genes, such as cytokines (CXCL13, the receptor activator of nuclear factor-kappaB ligand (RANKL), etc.) and stem cell transcription factors (TWIST2, SOX2, OCT4 and NANOG, etc.). Meanwhile, abnormal activation of NF-κB inflammatory signaling in tumor-associated stromal cells, such as myeloid-derived suppressor cells (MDSCs) and tumor-associated macrophages (TAMs), up-regulates the production of pro-inflammatory cytokines, chemokines and growth factors, which facilitates the formation of immunosuppressive TME and promotes the growth and development of CRPC [59,60]. It was reported that ursolic acid, a pentacyclic triterpenoid, inhibits IκB kinase activation and IκBα phosphorylation in PCa cells. Treatment with ursolic acid inhibits TNFα-induced NF-κB activation and the STAT3 signal pathway, resulting in decreased PCa cell proliferation in vitro and xenograft tumor growth in animal models [32]. Apigenin, a common plant flavonoid, can inhibit IKKα activation, IκBα phosphorylation and degradation. Treatment with apigenin leads to down-regulation of NF-κB-regulating genes, such as Bcl-2, cyclooxygenase-2 (COX-2) and matrix metalloproteinase 9 (MMP-9), and sensitizes hormone-refractory PCa cells to TNFα-induced apoptosis [33]. It was also reported that α-tomatine suppresses NF-κB signaling in PCa cells, resulting in down-regulation of NF-κB-dependent anti-apoptotic genes, such as Bcl-2, Bcl-xL and survivin (Figure 1 and Figure 2, Table 1) [34].

2.2. TANK-Binding Kinase 1 (TBK1)

The TBK1, a non-canonical IKK family member, plays a pivotal role in antiviral innate immunity. TBK1 is a critical component of antiviral responses through phosphorylation and activation of interferon regulatory factor 3 (IRF3), IRF7, and STAT1. Lipopolysaccharide (LPS) treatment induces TBK1/IKKi-dependent phosphorylation and activation of IRF3 in macrophages [61,62]. It was reported that TBK1 regulates PCa dormancy and chemotherapeutic resistance in PC3 and C4-2B models. TBK1 can interact with mTOR and reduce its activity; knockdown of TBK1 decreases drug resistance of PCa cells [63,64]. Currently, there are several TBK1 inhibitors available for laboratory use and preclinical studies. TBK1 inhibitor BX795 inhibits tumor growth through inducing mitotic phase arrest and cell apoptosis in a pre-clinical model of oral squamous cell carcinoma (OSCC) [65]. The combined treatment of BX795 and MEK inhibitor AZD6244 enhances apoptosis in AZD6244-resistant and NRAS mutated melanoma [66]. We reported that TBK1 and IKKi, another non-canonical IKK family member, can functionally compensate for each other. Individual knockdown of either IKKi or TBK1 have minor effects on cell survival while simultaneous knockdown of both TBK1 and IKKi significantly inhibited cell proliferation, suggesting that simultaneously targeting both TBK1 and IKKi is necessary for the efficient shutdown of cancer cell growth [62]. We developed TBK1/IKKi dual inhibitors, SR8185, 200A, and 200B that show potent anti-tumor activity in human breast, prostate, and oral cancer cells. These inhibitors function partially through suppression of TBK1/IKKi-mediated AKT phosphorylation and vascular endothelial-derived growth factor (VEGF) expression [62].

2.3. Protein Kinase A (PKA)

The PKA, composed of two regulatory subunits (R) and two catalytic subunits (C), is an enzyme whose activity is dependent on cellular cyclic AMP (cAMP) levels. The PKA is inactive when the two R subunits dimerize and bind to the C subunits. However, when cAMP binds to the R subunits, PKA undergoes a conformation change that leads to the release and free of active C subunit, and the phosphorylation of its substrates [35,67]. In PCa, PKA can synergize low levels of androgen during ADT to enhance androgen signaling and to promote the emergence of CRPC. PKA phosphorylates heat shock protein (HSP90), which binds to the unliganded AR in cytoplasm, leading to the release of AR from HSP90. The freed AR binds to HSP27 and translocates to the nucleus, resulting in up-regulation of AR targeting genes in androgen depleted environment. Treatment with PKA inhibitor H89 potently inhibits AR nuclear translocation [35]. It was also reported that PKA activity is regulated by PKA inhibitor proteins (PKIs) in PCa. The PKI family member PKIα binds to PKA that blocks PKA activity and disrupts PKA-mediated signaling events. PKIα acts as a molecular switch, diverting GPCR-Gαs-cAMP signaling toward EPAC-RAP1 and MAPK activation that promotes PCa growth, progression, and metastasis (Figure 1 and Figure 2, Table 1) [68].

2.4. PIM Kinase

PIM1 kinase, a serine/threonine kinase, belongs to the PIM family. PIM1 is a proto-oncogene and is associated with the development and progression of various human cancers, including PCa [69]. Two PIM1 kinase isoforms, PIM-1S and PIM-1L, are up-regulated in PCa, which promote PCa progression via modulating AR stability and transcriptional activity. PIM-1S phosphorylates AR at Ser213 that leads to recruitment of ubiquitin E3 ligase mdm2 and the degradation of AR, while PIM-1L phosphorylates AR at Thr-850 that results in recruitment of ubiquitin E3 ligase RING finger protein 6 (RNF6) and stabilization of AR [70]. Treatment with PIM1 inhibitors compound 9 and its pyrrolo[2,3-a]carbazole derivatives suppresses the proliferation of PCa cells [36]. It was observed that PIM kinases, PIM1 and PIM3, promote PCa cell migration and growth through phosphorylating CXCR4 chemokine receptor. Treatment with PIM kinases inhibitor DHPCC-9 suppresses PCa growth and metastasis (Figure 1 and Figure 2, Table 1) [37].

2.5. Tumor Progression Locus 2 (Tpl2)

Tpl2 is a serine/threonine protein kinase and acts as an essential modulator of immune responses. Tpl2 has been associated with tumor progression, metastasis, and therapy resistance [71]. The expression of Tpl2 in metastatic PCa is much higher than that in local primary PCa, and Tpl2 is associated with the emergence and maintenance of CRPC [6,72]. Tpl2 activates CXCL12/CXCR4 and focal adhesion kinase (FAK)/AKT signal pathway to promote PCa AI survival, metastasis, and chemoresistance. Tpl2 could be a promising therapeutic target for CRPC and metastatic CRPC (mCRPC) [6]. Several Tpl2 inhibitors, such as 1,7-naphthyridine-3-carbonitriles and quinoline-3-carbonitriles, were developed for the treatment of autoimmune diseases [73]. Interferon-α (IFNα) is identified as a natural inhibitor of Tpl2 in bladder cancer. IFN-α inhibits Tpl2 phosphorylation and suppresses COX-2 expression [74].

2.6. NEK6

NEK6, a mitotic-related serine/threonine kinase, is activated in cell mitosis and is required for mitotic spindle formation and cytokinesis [75]. Abnormal expression of NEK6 has been reported in human cancers, including hepatic cancer, breast cancer and PCa [3,76]. NEK6 interacts with and phosphorylates STAT3, leading to activation of STAT3-mediated transcription and transformation of JB6 Cl41 mouse epidermal cells [3]. In PCa, NEK6 phosphorylates transcription factor FOXJ2 and induces PCa AI growth. Knockdown of NEK6 restores the sensitivity of CRPC cells to ADT in PCa xenograft models [76], suggesting that combined NEK6 inhibition with ADT would be an effective therapeutic strategy for CRPC.

2.7. Glycogen-Synthase-Kinase 3β (GSK-3β)

GSK-3 is a serine/threonine kinase, which includes isoform GSK3α and GSK3β. GSK-3 participates in many important cellular events through phosphorylating its substrates, such as CCAAT/enhancer-binding protein (C/EBPα) and beta-catenin. GSK-3 is associated with human cancer development and progression [77]. Inhibition of GSK3β induces the AR nuclear export, resulting in growth inhibition of AR-positive CRPC cells [78]. There is a functional link between GSK-3, β-catenin and AR-V7 in CRPC cells. Inhibition of GSK-3 by inhibitor LY-2090314 activates β-catenin signaling, which sequentially suppresses AR-V7 transcription activity. Therefore, AR-V7-positive PCa cells are more vulnerable to LY-2090314 treatment than AR-V7-negative PCa cells [41]. In addition, GSK3β activity can be regulated by aspartyl (asparaginyl) β hydrolase (ASPH). Knockdown of ASPH increases the phosphorylation of GSK3β, suggesting that ASPH may interfere with the interaction of GSK3β with its upstream kinases (Figure 1 and Figure 2, Table 1) [79].

2.8. T-LAK Cell-Originated Protein Kinase (TOPK)

TOPK is characterized as a mitogen-activated serine/threonine protein kinase. TOPK is not expressed or is expressed at a very low level in most normal adult tissues; however, it is highly expressed in various cancers, including PCa. High expression of TOPK in PCa is associated with advanced stages, aggressiveness, and metastasis [26,80]. TOPK drives PCa cell AI growth via up-regulation of AR splice variant (AR-V7). Treatment with TOPK inhibitor OTS-514 suppresses AR transactivation and stability [26].

2.9. Cyclin-Dependent Kinases (CDKs)

The CDKs family are a group of different kinases that are involved in regulation of cell cycle, transcription, metabolism, and many other cellular events. To be active, CDKs need bind to a cyclin protein, and different combinations of specific CDKs and cyclins mark different parts of the cell cycle [81]. CDK mutations and abnormal activation have been found in many cancers, including PCa. CDK4/6 together with cyclin D1 regulates the phosphorylation of retinoblastoma tumor suppressor (Rb), which determines S phase entrance. Treatment with 3, 9-dihydroxy-2-prenylcoumestan (pso), a furanocoumarin and a CDK4 inhibitor, induces G0/G1 cell cycle arrest and growth inhibition of CRPC cells [29]. CDK12 plays an important role in transcriptional regulation, RNA splicing, and DNA integration. CDK12 mutations/loss are associated with genomic instability and have been observed in many human cancers, including PCa [82,83]. PCa with CDK12 mutations progresses rapidly to metastasis and castration resistance [83]. Abnormal CDK7 activation has been found in many cancers, including PCa, and CDK7 inhibitors have potent antitumor activity [30,84]. CDK7 can phosphorylate AR coactivator MED1 in CRPC cells. CDK7 inhibition, either through genetic silence or its inhibitor such as THZ1, interferes with the recruitment of MED1 to chromatin that suppresses the expression of AR regulated genes, resulting in inhibition of CRPC development [30,31]. CDK9 is a key component of P-TEFb transcription elongation complex. CDK9 has also been identified as a cofactor for AR, MYC and other oncogenic transcription factors [28,85]. A selective CDK9 inhibitor, KB-0742, shows potent anti-tumor activity in CRPC models. Treatment with KB-0742 rapidly down-regulates nascent transcription and AR-driven oncogenic programs in CRPC cells (Figure 1 and Figure 2, Table 1) [28].

2.10. Bromodomain (BRD)-Containing Kinases (BETs)

BETs are master regulators of global transcription elongation. BETs have been identified as attractive therapeutic targets for anti-viral infection, inflammation, and cancer [86]. BETs promote CRPC development by regulating AR-, ERG-, and c-Myc-mediated transcription [87]. BRD4, a bromodomain and extra terminal domain (BET) protein, interacts with N-terminal domain of AR to regulate AR signaling [88]. BRD4 also binds to super-enhancers (SEs), which drives high expression of oncogenes in tumor cells. BETs inhibitor JQ1 shows potent anticancer activity in PCa cells through down-regulation of SE-associated genes [16]. MZ1, a BRD4-selective degrader, inhibits metastasis of CRPC cells [89]. In addition, BRD4 is a H3K27ac reader and involves in CHPT1 super enhancer (SE) activity and CHPT1 expression in enzalutamide (Enz)-resistant CRPC cells [90]. BRD4 is also reported as a vital component of BRD4-KDM5C-PTEN oncogenic pathway in CRPC development. BRD4 up-regulates lysine-specific histone demethylase 5C (KDM5C) expression, which represses phosphatase and tensin homolog (PTEN) gene [91]. BRD4 is also a targeted gene of microRNA-200a (miR-200a), miR-200a suppresses CRPC development via interfering with the BRD4-mediated AR signaling [92]. Treatment with BRD4 inhibitors (MA4-022-1 compound and analogues) inhibits PCa cell proliferation and migration [17]. Two benzo[d]isoxazole derivatives, 6i (Y06036) and 7m (Y06137), function as potent BRD4 inhibitors, which bind to BRD4 and suppress the expression of MYC and AR regulated genes [19]. (R)-12, a benzoxazinone-containing 3,5-dimethylisoxazole derivative, is a newly developed BET inhibitor, which binds to BRD4 and blocks the interaction between bromodomain and acetyl lysine. Treatment with (R)-12 suppresses ERG, Myc and AR signaling, and inhibits CRPC cell proliferation, colony formation, and tumor growth [18]. Treatment with BET inhibitor ABBV-075 disrupts BRD4 recruitment and down-regulates AR target genes [20]. 7d, one of the several diphenylamine derivatives developed for the inhibition of BET activity in CRPC, exhibits favorable metabolic stability and might be a promising agent for the treatment of clinical CRPC (Figure 2 and Table 1) [21].

2.11. Aurora Kinase A (AURKA)

AURKA is a cell cycle-regulated kinase that is involved in microtubule formation and/or stabilization at the spindle pole during chromosome segregation. AURKA is regulated by androgen, and is highly expressed in CRPC samples with amplification and/or high expression of AR. Treatment with AURKA specific inhibitor alisertib (MLN8237) significantly inhibits the growth of AR-overexpressing CRPC cells [7]. AURKA regulates androgen receptor variants (AR-Vs) expression, knockdown of AURKA leads to decreased synthesis of AR-V transcripts, such as AR-V7, but not the full-length AR. Treatment with AURKA inhibitor S1451 significantly decreases AR-V7 expression levels and markedly reduces AR-V-driven proliferation and survival of CRPC cells [25]. It has also been reported that AURKA phosphorylates Y-box binding protein-1 (YBX1), resulting in stabilization and nuclear translocation of YBX1 in CRPC cells. In turn, YBX1 also stabilizes AURKA, and therefore forming an AURKA-YBX1 synergistic loop that promotes PCa epithelial to mesenchymal transition (EMT), progression and chemoresistance [93]. SPOP, an adapter protein for E3 ubiquitin ligase, is a direct substrate of AURKA. AURKA induces SPOP degradation via phosphorylation of SPOP, leading to stabilization of AR, AR-V7, and c-Myc. Treatment with AURKA inhibitor alisertib (MLN8237) sensitizes CRPC cells to enzalutamide (Figure 1 and Figure 2, Table 1) [94].

2.12. AMP-Activated Protein Kinase (AMPK)

AMPK is a conserved serine/threonine protein kinase and is regarded as a metabolic sensor in mammalian cells. Forkhead box protein M1 (FOXM1) drives CRPC docetaxel resistance via induction of AMPK/mTOR-mediated autophagy [95]. Prostate Leucine Zipper (PrLZ), an important gene for CRPC development, promotes PCa docetaxel resistance through suppression of LKB1/AMPK-mediated autophagy [96].

2.13. Mitogen-Activated Protein Kinase (MAPK) and Jun Kinase (JNK)

Abnormal MAPK activation promotes CRPC development, resulting in early relapse and short disease-free survival [97]. A p38 MAPK and heat shock protein 27 (Hsp27) driven signaling axis has been identified as an important regulator of AR activity in hormone-sensitive PCa cells. Inhibition of p38 MAPK suppresses Hsp27 and the hypoxia-mediated AR activation and CRPC development in PCa mouse models [98]. ADT induces a compensatory activation of MAPK or JNK signaling, combination of JNK inhibitor AS602801 with enzalutamide synergistically inhibits PCa cell proliferation, invasion, and migration in vitro and xenograft tumor development in mouse models [99].

2.14. Phosphatidylinositol-3-Kinase (PI3K)/AKT

PI3K/AKT pathway plays an important role in cell metabolism, growth, proliferation, and survival. Abnormal activation of PI3K/AKT pathway is closely associated with CRPC development [100]. The combined treatment of enzalutamide with AKT inhibitor (AZD5363) significantly slows down enzalutamide-resistant PCa progression in mouse models [101]. Targeting both the PI3K/AKT pathway and androgen receptor axis significantly delays CRPC development in animal models [102]. The combined treatment of GRT (an aspalathin-rich green rooibos extract), Bcl-2 inhibitor ABT-737, PI3K inhibitor LY294002, and AKT inhibitor GSK 690693 exhibits potent inhibitory effect on CRPC development [103]. Treatment with PI3K specific inhibitor wortmannin and AKT inhibitor MK2206 blocks PI3K/AKT signaling and inhibits PCa cell invasion and colony formation [104]. In addition, the combination of Bcl-2 inhibitor ABT-737 with AKT inhibitor erufosine also shows a very potent anticancer effect on CRPC cells (Figure 1 and Figure 2, Table 1) [105].

2.15. Polo-Like Kinase 1 (PLK1)

PLK1 belongs to polo-like kinases (PLKs) family, which contains four serine/threonine protein kinases that are critical regulators in cell cycle progression, mitosis, cytokinesis, and DNA damage responses [106]. PLK1 is overexpressed in PCa and is associated with higher grade tumors. PLK1 signaling is one of the most activated pathways after ADT. PLK1 activation promotes CRPC development through PI3K–AKT–mTOR pathway as well as AR and Wnt/β-catenin signaling. PLK1 inhibition sensitizes PCa cells to androgen signaling inhibitors [107,108]. Targeting PLK1 with specific inhibitor BI2536 enhances the effect of PARP inhibitor on CRPC cells [13]. As PLK1 inhibition leads to β-catenin signaling activation c-MYC overexpression, PLK1 inhibitor GSK461364A shows a strong synergistic inhibitory effect on CRPC cells when combined with BRD4 inhibitor JQ1, which inhibits c-MYC expression and AR signaling [14]. Treatment with BETs and PLK1 dual inhibitor WNY0824 significantly suppresses AR-positive CRPC cell proliferation in vitro and xenograft tumor growth in animal models (Figure 1 and Figure 2, Table 1) [15].

2.16. LIM-Domain Kinase-2 (LIMK2)

LIMK2, a serine/threonine and tyrosine kinase, contains two LIM motifs, a PDZ and kinase domain. The expression of LIMK2 is very high in PCa, and is further increased after ADT. LIMK2 promotes CRPC development through phosphorylation and stabilization of TWIST1 [10]. In addition, LIMK2 phosphorylates and degrades SPOP, which results in stabilization of AR, AR-V7 and c-MYC in CRPC cells [109]. LIMK2 also phosphorylates and degrades PTEN [110]. LIMK2 allosteric inhibitor (LI) shows high synergistic effect with docetaxel on CRPC cells [10,109].

3. Tyrosine Kinase and Their Therapeutic Implications

Tyrosine kinases phosphorylate tyrosine amino acid residues, which function primarily as growth factor receptors or/and in downstream signaling of growth factors. Tyrosine kinases include receptor tyrosine kinases (RTKs) and non-receptor tyrosine kinases (NRTKs). Abnormal activation of tyrosine kinases is associated with the development of various cancers, including PCa [111,112].

3.1. The Receptor Tyrosine Kinases (RTKs)

The epidermal growth factor receptor (EGFR, also known as HER1) is a receptor tyrosine kinase and is a member of ErbB family [112]. Abnormal EGFR signaling is associated with PCa pathogenesis, progression, and CRPC development. It was reported that the nuclear localization of EGFR can be regulated by Annexin A1, and the increased nuclear EGFR promotes CRPC development. Cyclosporin H can interrupt nuclear EGFR and its downstream signaling, which consequently suppresses PCa growth [113]. EGFR induces CRPC docetaxel resistance via regulating Akt-ABCB1 signaling. Treatment with EGFR inhibitor gefitinib significantly suppresses proliferation of docetaxel-resistant PCa cells in vitro and xenograft tumor growth in animal models (Figure 2 and Table 1) [40].
Human epidermal growth factor receptor 2 (HER2) is another member of ErbB family, which is well known to be associated with breast cancer. HER2 is also involved in AR activation and PCa progression [41,56,114]. Lapatinib, a dual EGFR and HER2 inhibitor, has been applied in a phase II trial and shown potent initial antitumor activity in CRPC patients [56]; however, development of lapatinib resistance limits its clinical application. The FDA-approved ErbB inhibitor dacomitinib, at equimolar concentrations as lapatinib, can decrease HER2 protein stability and prevent EGFR/HER2 heterodimerization that lead to interrupting downstream signaling and increasing CRPC cell apoptosis [41]. It has been reported that both MET and VEGFR signal pathways play an important role in CRPC [115]. Dual inhibition of MET and VEGFR by cabozantinib (formerly XL184, a multi-targeted tyrosine kinase inhibitor) or the combined axitinib (VEGFR inhibitor) with crizotinib (MET inhibitor) shows strong synergistic inhibitory effect on CRPC (Figure 2 and Table 1) [115,116].

3.2. The Non-Receptor Tyrosine Kinases (NRTKs)

The rearranged during transfection (RET) is a tyrosine kinase and plays an important role in cell growth and differentiation. Abnormal RET expression is reported in various cancers, including PCa. RET expression is increased in neuroendocrine PCa (NEPC), and is associated with neuroendocrine transcription factors, such as POU3F2, SOX2, ONECUT2, and ASCL1. Treatment with RET inhibitor AD80 induces cell death and suppresses tumor development in NEPC xenograft models [42]. The Lck/yes-related protein tyrosine kinase (Lyn) belongs to the Src kinase family (SFK) and plays a pivotal role in the pathogenesis of various tumors. The expression of Lyn tyrosine kinase is increased in human CRPC as compared with hormone naive PCa or normal prostate tissue. Overexpression of Lyn enhances AR transcriptional activity and accelerates CRPC progression. Targeting Lyn kinase induces AR dissociation from its molecular chaperone Hsp90, leading to AR ubiquitination and degradation [117]. Ack1 tyrosine kinase phosphorylates AR at Tyr-267, treatment with Ack1 inhibitor AIM-100 suppresses AR-Tyr-267 phosphorylation and Ack1 mediated ATM (ataxia telangiectasia mutated) expression as well as the growth of radioresistant CRPC (Figure 2 and Table 1) [118].
The Janus family kinases (Jaks), Jak1, Jak2, Jak3, and Tyk2, form one subgroup of the NRTKs. Jaks are involved in cell growth, survival, development, and differentiation. The activation of Jaks by IFNs induces phosphorylation of transcription factor Stat1 at Y701, leading to Stat1 dimerization and translocation to nucleus. Another protein tyrosine kinase Pyk2 also plays a critical role in the Jak-mediated MAPK and Stat1 activation [119]. JAK2 can mediate both tyrosine phosphorylation and serine phosphorylation [120]. JAK2 is overexpressed in PCa cells. Jak2 inhibitor AZD1480 inhibits Stat5a/b phosphorylation, dimerization, nuclear translocation, DNA binding and transcriptional activity in PCa cells. Treatment of Jak2 inhibitor AZD1480 suppresses PCa xenograft tumor development, and prolongs survival time of tumor-bearing mice as compared with vehicle or docetaxel-treated groups [121]. In addition, it was reported that β-elemonic acid (β-EA), a natural triterpene, inhibits growth and triggers apoptosis of human CRPC cells through the suppression of JAK2/STAT3/MCL-1 and NF-ĸB signaling pathways (Figure 2 and Table 1) [39].

4. Lipid Kinases and Their Therapeutic Implications

Lipid kinases phosphorylate lipids in the cell, both on the plasma membrane and on the membranes of the organelles. The addition of phosphate groups can change the reactivity and localization of the lipid, leading to signal transmission.

4.1. Phosphatidylinositol 4-Phosphate 5-Kinase Type 1α (PIP5K1α)

PIP5K1α belongs to the family of PI-4-phosphate 5 kinases (PIP5Ks) and contacts the substrate PI4P via two major regions of an arginine/lysine-rich region (the β8-α4ϲ loop) and the activation loop (Protein Data Bank (PDB: 4TZ7) residues 378–415, a structure of PIP5K1α) [122]. PIP5K1α is upstream of PI3K since PIP5K1α produces a substrate of PI3K, PtdIns-4,5-P2 (PIP2), which is required for the activation of PI3K/AKT pathway. The activation of PI3K/AKT signaling is observed in various cancers, including PCa [43]. PIP5K1α is highly expressed in advanced and metastatic PCa as compared with primary tumors, and PIP5K1α expression is positively correlated with AR expression. Overexpression of PIP5K1α increases AKT activity and the invasiveness of PCa cells. Knockdown of PIP5K1α or treatment with its inhibitor ISA-2011B significantly inhibits AKT activity, increases cancer cell apoptosis and suppresses CRPC growth in animal models (Figure 2 and Table 1) [43].

4.2. Sphingosine Kinases (SKs)

SKs are lipid kinases that catalyze the conversion of sphingosine to sphingosine-1-phosphate (S1P). On activation, SKs migrate from cytosol to plasma membrane where they transfer a γ phosphate from ATP or GTP to sphingosine. There are two SKs in mammalian cells, SK1 and SK2. SK1 is highly expressed in certain types of cancers, including PCa [123]. SK1 expression level and activity are associated with PCa progression, recurrence, and chemotherapy resistance [44,123,124]. AI PCa PC3 cells have much higher SK1 activity than AD PCa LNCaP cells, and SK1 involves in AI survival of PCa cells [123,125]. Treatment with SK1 inhibitor FTY720 (fingolimod) activates caspase-3 and induces apoptosis in DU145 cells. FTY720 treatment also sensitizes CRPC cells to radiotherapy. Furthermore, SK1 inhibition overcomes enzalutamide resistance and enhances enzalutamide efficacy in CRPC cells [44,45,46]. It was also reported that treatment with SK2 inhibitor ABC294640 reduces CRPC cell growth as well as the expression of c-Myc and AR. ABC294640 treatment also inhibits dihydroceramide desaturase (DEGS), resulting in the increase of dihydroceramides (Figure 2 and Table 1) [47].

5. Carbohydrate Kinases and Their Therapeutic Implications

5.1. Hexokinase

Hexokinase is the most common enzyme in glucose metabolism, which converts D-glucose to glucose-6-phosphate by transferring the gamma phosphate of an ATP to the C6 position. Elevated expression of hexokinase 2 (HK2) has been identified in several cancers [126]. In PCa, increased HK2 expression and activity are associated with CRPC development, especially PTEN- and TP53-deficiency-driven CRPC [48,126,127]. Under ADT, HK2 stability is regulated by IL13Rα1 in PCa cells. IL13Rα1 recruits ubiquitin protein ligase E3C to mediate ubiquitination and degradation of HK2, leading to glycolytic inhibition and PCa cell apoptosis [128]. Targeting HK2 with its inhibitor 2-deoxyglucose (2-DG) suppresses CRPC growth. When 2-DG is combined with chloroquine (CQ), which targets ULK1-dependent autophagy, the treatment leads to a near-complete tumor suppression and remarkably extend survival in PTEN- and TP53-deficiency-driven CRPC animal models (Figure 2 and Table 1) [48].

5.2. Phosphofructokinase (PFK)

PFK catalyzes the conversion of fructose-6-phosphate to fructose-1,6-bisphosphate and is a key enzyme in the regulation of glycolysis. PFK acts as an important mediator of metabolism in cancer cells to support growth and survival. The expression level of PFK is elevated in several cancers, such as breast cancer, PCa, and lung cancer; the expression of phosphofructokinase platelet (PFKP), an isoform of PFK, is associated with poor prognosis of breast cancer [129]. In CRPC, PFK is associated with PCa cell proliferation and its expression is regulated by KDM4B [130].

6. Other Kinases and Their Therapeutic Implications

In addition to the kinases that phosphorylate proteins, lipids, and carbohydrates, there are some kinases acting on nucleotides, such as nucleoside-phosphate kinases and nucleoside-diphosphate kinases. Other small molecules such as creatine, phosphoglycerate, riboflavin, dihydroxyacetone, shikimate can also be substrates of kinases.

6.1. Riboflavin Kinases (RFK)

RFK catalyzes the phosphorylation of riboflavin to create flavin mononucleotide (FMN), which is an important cofactor and a precursor to flavin adenine dinucleotide (FAD) as well as a redox cofactor used by many enzymes in metabolism. It has been reported that RFK involved in cisplatin resistance and cellular protection from oxidative stress in PC3 cells. Knockdown of RFK enhances the sensitivity of PC3 cells to cisplatin [131].

6.2. Thymidine Kinase

Thymidine kinase is one of the nucleoside kinases that is responsible for nucleoside phosphorylation. It phosphorylates thymidine to create thymidine monophosphate (dTMP). Thymidine kinase activity is closely correlated with cell cycle and has been used as a tumor marker in clinical chemistry [132]. Osteocalcin-thymidine kinase (OC-TK) is highly expressed in AI PCa cells as compared with AD PCa cells [133]. In addition, the concentrations of serum thymidine kinase-1 (TK-1) are correlated with PCa stages. Together with PSA, TK-1 can be used as a diagnostic marker for PCa [134].

7. Conclusions and Perspectives

ADT was first developed in 1941 by Dr. Charles Huggins, who won the Nobel Prize in Physiology and Medicine in 1966, and still remains the principal and backbone treatment for patients with advanced PCa, because the majority of patients respond to ADT with a mean remission time of 2–3 years, which is a big advantage for those patients with advanced PCa. Currently there is no better systemic therapy that could replace ADT for the treatment of advanced PCa. However, once it becomes castration-resistant, the PCa is lethal and incurable. Many kinase-mediated signaling pathways are abnormally activated and involved in the pathogenesis of CRPC. And, therefore, kinases have been regarded as one of the most important and promising targets for the treatment of CRPC. Small molecule inhibitors exhibit good penetration and accessibility as compared with other large molecules; many efforts have been made to develop small molecule inhibitors that target kinases mutated or abnormally activated in CRPC (Figure 1 and Figure 2). Many of these small molecule inhibitors are still at the early stages in preclinical studies, some are in clinical trials (Table 1 and Table 2).
Multiple kinases or pathways can be activated simultaneously in one single tumor cell, and different tumor cells in the same tumor have different kinase or pathway activation. Additionally, one kinase or pathway can be compensated for by other kinases or pathways once being targeted. Although we appreciate that most of these available kinase inhibitors are designed to target one specific kinase, which can avoid side (off-target) effects caused by nonspecific inhibition of other targets important for normal cells, targeting one single kinase or pathway may not be efficient or no effect at all for the suppression of tumor development. For instance, both MET and VEGFR signaling play an important role in CRPC development [115]; the inhibition of MET and VEGFR by a multi-targeted tyrosine kinase inhibitor cabozantinib shows much stronger anti-CRPC activity than either VEGFR inhibitor axitinib or MET inhibitor crizotinib used alone [115,116]. We reported that TBK1 and IKKi can functionally compensate for each other. Knockdown of either IKKi or TBK1 has a minor effect on cell survival while simultaneous knockdown of both TBK1 and IKKi significantly inhibits cell proliferation [62]. TBK1/IKKi dual inhibitors show potent anti-tumor activity in human breast, prostate, and oral cancer cells [62].
Currently, most kinase inhibitors are developed to target the kinases in cancer cells. tumor is composed of not only cancer cells but also many types of stromal cells, such as tumor-associated macrophages (TAMs), myeloid-derived suppressor cells (MDSCs), T cells, B cells, and NK cells [12,135]. These stromal cells, particularly infiltrated immune/inflammatory cells, either contribute to or are affected by the immunosuppressive tumor microenvironment (TME) [136]. The abnormal signaling mediated by kinase activation/suppression in immune/inflammatory cells in TME would be potential good targets. Therefore, development of therapeutic strategies, particularly small molecule inhibitors/activators that could not only inhibit abnormal kinase activation in tumor cells to suppress tumor cell proliferation, but also modify the kinases activity in immune/inflammatory cells to tilt the balance of TME from immune suppression to effective immune surveillance, or/and to restore the tumor-killing function of “exhausted” CD8+ T cells, would be highly appreciated.

Author Contributions

Conceptualization, S.Z., S.P., Z.C. (Zihua Chen), Z.C. (Zhikang Chen) and J.-L.L.; writing—original draft preparation, S.Z., S.P. and J.-L.L.; writing—review and editing, S.Z., S.P., Z.C. (Zihua Chen), Z.C. (Zhikang Chen) and J.-L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Frenchman’s Creek Women for Cancer Research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Manning, G.; Whyte, D.B.; Martinez, R.; Hunter, T.; Sudarsanam, S. The Protein Kinase Complement of the Human Genome. Science 2002, 298, 1912–1934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Gross, S.; Rahal, R.; Stransky, N.; Lengauer, C.; Hoeflich, K.P. Targeting cancer with kinase inhibitors. J. Clin. Investig. 2015, 125, 1780–1789. [Google Scholar] [CrossRef] [PubMed]
  3. Jeon, Y.J.; Lee, K.Y.; Cho, Y.-Y.; Pugliese, A.; Kim, H.G.; Jeong, C.-H.; Bode, A.M.; Dong, Z. Role of NEK6 in Tumor Promoter-induced Transformation in JB6 C141 Mouse Skin Epidermal Cells. J. Biol. Chem. 2010, 285, 28126–28133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: Globocan Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  5. Karantanos, T.; Corn, P.G.; Thompson, T.C. Prostate cancer progression after androgen deprivation therapy: Mechanisms of castrate resistance and novel therapeutic approaches. Oncogene 2013, 32, 5501–5511. [Google Scholar] [CrossRef]
  6. Lee, H.W.; Cho, H.J.; Lee, S.J.; Song, H.J.; Cho, H.J.; Park, M.C.; Seol, H.J.; Lee, J.-I.; Kim, S.; Lee, H.M.; et al. Tpl2 induces castration resistant prostate cancer progression and metastasis. Int. J. Cancer 2015, 136, 2065–2077. [Google Scholar] [CrossRef]
  7. Kivinummi, K.; Urbanucci, A.; Leinonen, K.; Tammela, T.L.J.; Annala, M.; Isaacs, W.B.; Bova, G.S.; Nykter, M.; Visakorpi, T. The expression of AURKA is androgen regulated in castration-resistant prostate cancer. Sci. Rep. 2017, 7, 17978. [Google Scholar] [CrossRef] [Green Version]
  8. Zhong, S.; Jeong, J.-H.; Huang, C.; Chen, X.; Dickinson, S.I.; Dhillon, J.; Yang, L.; Luo, J.-L. Targeting INMT and interrupting its methylation pathway for the treatment of castration resistant prostate cancer. J. Exp. Clin. Cancer Res. 2021, 40, 307. [Google Scholar] [CrossRef]
  9. Zhong, S.; Huang, C.; Chen, Z.; Luo, J.-L. Targeting Inflammatory Signaling in Prostate Cancer Castration Resistance. J. Clin. Med. 2021, 10, 5000. [Google Scholar] [CrossRef]
  10. Nikhil, K.; Chang, L.; Viccaro, K.; Jacobsen, M.; McGuire, C.; Satapathy, S.R.; Tandiary, M.A.; Broman, M.M.; Cresswell, G.; He, Y.J.; et al. Identification of LIMK2 as a therapeutic target in castration resistant prostate cancer. Cancer Lett. 2019, 448, 182–196. [Google Scholar] [CrossRef] [Green Version]
  11. Imai, K.; Takaoka, A. Comparing antibody and small-molecule therapies for cancer. Nat. Cancer 2006, 6, 714–727. [Google Scholar] [CrossRef] [PubMed]
  12. Zhong, S.; Jeong, J.-H.; Chen, Z.; Chen, Z.; Luo, J.-L. Targeting Tumor Microenvironment by Small-Molecule Inhibitors. Transl. Oncol. 2020, 13, 57–69. [Google Scholar] [CrossRef] [PubMed]
  13. Li, J.; Wang, R.; Kong, Y.; Broman, M.M.; Carlock, C.; Chen, L.; Li, Z.; Farah, E.; Ratliff, T.L.; Liu, X. Targeting Plk1 to Enhance Efficacy of Olaparib in Castration-Resistant Prostate Cancer. Mol. Cancer Ther. 2017, 16, 469–479. [Google Scholar] [CrossRef] [Green Version]
  14. Mao, F.; Li, J.; Luo, Q.; Wang, R.; Kong, Y.; Carlock, C.; Liu, Z.; Elzey, B.D.; Liu, X. Plk1 Inhibition Enhances the Efficacy of BET Epigenetic Reader Blockade in Castration-Resistant Prostate Cancer. Mol. Cancer Ther. 2018, 17, 1554–1565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Xu, Y.; Wang, Q.; Xiao, K.; Liu, Z.; Zhao, L.; Song, X.; Hu, X.; Feng, Z.; Gao, T.; Zuo, W.; et al. Novel dual BET and PLK1 inhibitor WNY0824 exerts potent anti-tumor effects in CRPC by inhibiting transcription factor function and inducing mitotic abnormality. Mol. Cancer Ther. 2020, 19, 1221–1231. [Google Scholar] [CrossRef] [Green Version]
  16. Nagasawa, M.; Tomimatsu, K.; Terada, K.; Kondo, K.; Miyazaki, K.; Miyazaki, M.; Motooka, D.; Okuzaki, D.; Yoshida, T.; Kageyama, S.; et al. Long non-coding RNA MANCR is a target of BET bromodomain protein BRD4 and plays a critical role in cellular migration and invasion abilities of prostate cancer. Biochem. Biophys. Res. Commun. 2020, 526, 128–134. [Google Scholar] [CrossRef]
  17. Nerlakanti, N.; Yao, J.; Nguyen, D.T.; Patel, A.K.; Eroshkin, A.M.; Lawrence, H.R.; Ayaz, M.; Kuenzi, B.; Agarwal, N.; Chen, Y.; et al. Targeting the BRD4-HOXB13 Coregulated Transcriptional Networks with Bromodomain-Kinase Inhibitors to Suppress Metastatic Castration-Resistant Prostate Cancer. Mol. Cancer Ther. 2018, 17, 2796–2810. [Google Scholar] [CrossRef] [Green Version]
  18. Xue, X.; Zhang, Y.; Wang, C.; Zhang, M.; Xiang, Q.; Wang, J.; Wang, A.; Li, C.; Zhang, C.; Zou, L.; et al. Benzoxazinone-containing 3,5-dimethylisoxazole derivatives as BET bromodomain inhibitors for treatment of castration-resistant prostate cancer. Eur. J. Med. Chem. 2018, 152, 542–559. [Google Scholar] [CrossRef]
  19. Zhang, M.; Zhang, Y.; Song, M.; Xue, X.; Wang, J.; Wang, C.; Zhang, C.; Li, C.; Xiang, Q.; Zou, L.; et al. Structure-Based Discovery and Optimization of Benzo[d]isoxazole Derivatives as Potent and Selective BET Inhibitors for Potential Treatment of Castration-Resistant Prostate Cancer (CRPC). J. Med. Chem. 2018, 61, 3037–3058. [Google Scholar] [CrossRef]
  20. Faivre, E.J.; Wilcox, D.; Lin, X.; Hessler, P.; Torrent, M.; He, W.; Uziel, T.; Albert, D.H.; McDaniel, K.; Kati, W.; et al. Exploitation of Castration-Resistant Prostate Cancer Transcription Factor Dependencies by the Novel BET Inhibitor ABBV-075. Mol. Cancer Res. 2016, 15, 35–44. [Google Scholar] [CrossRef] [Green Version]
  21. Yu, J.; Zhou, P.; Du, W.; Xu, R.; Yan, G.; Deng, Y.; Li, X.; Chen, Y. Metabolically stable diphenylamine derivatives suppress androgen receptor and BET protein in prostate cancer. Biochem. Pharmacol. 2020, 177, 113946. [Google Scholar] [CrossRef] [PubMed]
  22. Goodwin, N.C.; Cianchetta, G.; Burgoon, H.A.; Healy, J.; Mabon, R.; Strobel, E.D.; Allen, J.; Wang, S.; Hamman, B.D.; Rawlins, D.B. Discovery of a Type III Inhibitor of LIM Kinase 2 That Binds in a DFG-Out Conformation. ACS Med. Chem. Lett. 2014, 6, 53–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Corno, C.; Arrighetti, N.; Ciusani, E.; Corna, E.; Carenini, N.; Zaffaroni, N.; Gatti, L.; Perego, P. Synergistic Interaction of Histone Deacetylase 6- and MEK-Inhibitors in Castration-Resistant Prostate Cancer Cells. Front. Cell Dev. Biol. 2020, 8, 610. [Google Scholar] [CrossRef] [PubMed]
  24. Nickols, N.G.; Nazarian, R.; Zhao, S.G.; Tan, V.; Uzunangelov, V.; Xia, Z.; Baertsch, R.; Neeman, E.; Gao, A.C.; Thomas, G.; et al. MEK-ERK signaling is a therapeutic target in metastatic castration resistant prostate cancer. Prostate Cancer Prostatic Dis. 2019, 22, 531–538. [Google Scholar] [CrossRef]
  25. Jones, D.; Noble, M.; Wedge, S.R.; Robson, C.N.; Gaughan, L. Aurora A regulates expression of AR-V7 in models of castrate resistant prostate cancer. Sci. Rep. 2017, 7, srep40957. [Google Scholar] [CrossRef] [Green Version]
  26. Alhawas, L.; Amin, K.S.; Salla, B.; Banerjee, P.P. T-LAK cell-originated protein kinase (TOPK) enhances androgen receptor splice variant (ARv7) and drives androgen-independent growth in prostate cancer. Carcinogenesis 2021, 42, 423–435. [Google Scholar] [CrossRef]
  27. Nakata, D.; Koyama, R.; Nakayama, K.; Kitazawa, S.; Watanabe, T.; Hara, T. Glycogen synthase kinase-3 inhibitors suppress the AR-V7-mediated transcription and selectively inhibit cell growth in AR-V7-positive prostate cancer cells. Prostate 2017, 77, 955–961. [Google Scholar] [CrossRef]
  28. Richters, A.; Doyle, S.K.; Freeman, D.B.; Lee, C.; Leifer, B.S.; Jagannathan, S.; Kabinger, F.; Koren, J.V.; Struntz, N.B.; Urgiles, J.; et al. Modulating Androgen Receptor-Driven Transcription in Prostate Cancer with Selective CDK9 Inhibitors. Cell Chem. Biol. 2021, 28, 134–147.e14. [Google Scholar] [CrossRef]
  29. Gulappa, T.; Reddy, R.S.; Suman, S.; Nyakeriga, A.M.; Damodaran, C. Molecular interplay between cdk4 and p21 dictates G0/G1 cell cycle arrest in prostate cancer cells. Cancer Lett. 2013, 337, 177–183. [Google Scholar] [CrossRef] [Green Version]
  30. Asangani, I.; Blair, I.A.; Van Duyne, G.; Hilser, V.J.; Moiseenkova-Bell, V.; Plymate, S.; Sprenger, C.; Wand, A.J.; Penning, T.M. Using biochemistry and biophysics to extinguish androgen receptor signaling in prostate cancer. J. Biol. Chem. 2021, 296, 100240. [Google Scholar] [CrossRef]
  31. Rasool, R.U.; Natesan, R.; Deng, Q.; Aras, S.; Lal, P.; Effron, S.S.; Mitchell-Velasquez, E.; Posimo, J.M.; Carskadon, S.; Baca, S.C.; et al. CDK7 Inhibition Suppresses Castration-Resistant Prostate Cancer through MED1 Inactivation. Cancer Discov. 2019, 9, 1538–1555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Shanmugam, M.K.; Rajendran, P.; Li, F.; Nema, T.; Vali, S.; Abbasi, T.; Kapoor, S.; Sharma, A.; Kumar, A.P.; Ho, P.C.; et al. Ursolic acid inhibits multiple cell survival pathways leading to suppression of growth of prostate cancer xenograft in nude mice. Klin. Wochenschr. 2011, 89, 713–727. [Google Scholar] [CrossRef] [PubMed]
  33. Shukla, S.; Gupta, S. Suppression of Constitutive and Tumor Necrosis Factor α-Induced Nuclear Factor (NF)-κB Activation and Induction of Apoptosis by Apigenin in Human Prostate Carcinoma PC-3 Cells: Correlation with Down-Regulation of NF-κB-Responsive Genes. Clin. Cancer Res. 2004, 10, 3169–3178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Lee, S.-T.; Wong, P.-F.; He, H.; Hooper, J.D.; Mustafa, M.R. Alpha-Tomatine Attenuation of In Vivo Growth of Subcutaneous and Orthotopic Xenograft Tumors of Human Prostate Carcinoma PC-3 Cells Is Accompanied by Inactivation of Nuclear Factor-Kappa B Signaling. PLoS ONE 2013, 8, e57708. [Google Scholar] [CrossRef] [Green Version]
  35. Dagar, M.; Singh, J.P.; Dagar, G.; Tyagi, R.K.; Bagchi, G. Phosphorylation of HSP90 by protein kinase A is essential for the nuclear translocation of androgen receptor. J. Biol. Chem. 2019, 294, 8699–8710. [Google Scholar] [CrossRef] [Green Version]
  36. Akué-Gédu, R.; Rossignol, E.; Azzaro, S.; Knapp, S.; Filippakopoulos, P.; Bullock, A.N.; Bain, J.; Cohen, P.; Prudhomme, M.; Anizon, F.; et al. Synthesis, Kinase Inhibitory Potencies, and in Vitro Antiproliferative Evaluation of New Pim Kinase Inhibitors. J. Med. Chem. 2009, 52, 6369–6381. [Google Scholar] [CrossRef]
  37. Santio, N.M.; Eerola, S.K.; Paatero, I.; Yli-Kauhaluoma, J.; Anizon, F.; Moreau, P.; Tuomela, J.; Härkönen, P.; Koskinen, P.J. Pim Kinases Promote Migration and Metastatic Growth of Prostate Cancer Xenografts. PLoS ONE 2015, 10, e0130340. [Google Scholar] [CrossRef] [Green Version]
  38. Mahajan, K.; Challa, S.; Coppola, D.; Lawrence, H.; Luo, Y.; Gevariya, H.; Zhu, W.; Chen, Y.; Lawrence, N.J.; Mahajan, N.P. Effect of Ack1 tyrosine kinase inhibitor on ligand-independent androgen receptor activity. Prostate 2010, 70, 1274–1285. [Google Scholar] [CrossRef] [Green Version]
  39. Bao, X.; Zhu, J.; Ren, C.; Zhao, A.; Zhang, M.; Zhu, Z.; Lu, X.; Zhang, Y.; Li, X.; Sima, X.; et al. β-elemonic acid inhibits growth and triggers apoptosis in human castration-resistant prostate cancer cells through the suppression of JAK2/STAT3/MCL-1 and NF-ĸB signal pathways. Chem. Interactions 2021, 342, 109477. [Google Scholar] [CrossRef]
  40. Hour, T.-C.; Chung, S.-D.; Kang, W.-Y.; Lin, Y.-C.; Huang, C.-Y.; Huang, A.-M.; Wu, W.J.; Huang, S.-P.; Pu, Y.-S. EGFR mediates docetaxel resistance in human castration-resistant prostate cancer through the Akt-dependent expression of ABCB1 (MDR1). Arch. Toxicol. 2015, 89, 591–605. [Google Scholar] [CrossRef]
  41. Jathal, M.K.; Steele, T.M.; Siddiqui, S.; Mooso, B.A.; D’abronzo, L.S.; Drake, C.M.; Whang, Y.E.; Ghosh, P.M. Dacomitinib, but not lapatinib, suppressed progression in castration-resistant prostate cancer models by preventing HER2 increase. Br. J. Cancer 2019, 121, 237–248. [Google Scholar] [CrossRef] [PubMed]
  42. VanDeusen, H.R.; Ramroop, J.R.; Morel, K.L.; Bae, S.Y.; Sheahan, A.V.; Sychev, Z.; Lau, N.A.; Cheng, L.C.; Tan, V.M.; Li, Z.; et al. Targeting RET Kinase in Neuroendocrine Prostate Cancer. Mol. Cancer Res. 2020, 18, 1176–1188. [Google Scholar] [CrossRef]
  43. Semenas, J.; Hedblom, A.; Miftakhova, R.; Sarwar, M.; Larsson, R.; Shcherbina, L.; Johansson, M.E.; Härkönen, P.; Sterner, O.; Persson, J.L. The role of PI3K/AKT-related PIP5K1 and the discovery of its selective inhibitor for treatment of advanced prostate cancer. Proc. Natl. Acad. Sci. USA 2014, 111, E3689–E3698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lin, H.-M.; Mak, B.; Yeung, N.; Huynh, K.; Meikle, T.G.; Mellett, N.A.; Kwan, E.M.; Fettke, H.; Tran, B.; Davis, I.D.; et al. Overcoming enzalutamide resistance in metastatic prostate cancer by targeting sphingosine kinase. EBioMedicine 2021, 72, 103625. [Google Scholar] [CrossRef]
  45. Wang, J.D. Early induction of apoptosis in androgen-independent prostate cancer cell line by FTY720 requires caspase-3 activation. Prostate 1999, 40, 50–55. [Google Scholar] [CrossRef]
  46. Pchejetski, D.; Bohler, T.; Brizuela, L.; Sauer, L.; Doumerc, N.; Golzio, M.; Salunkhe, V.; Teissié, J.; Malavaud, B.; Waxman, J.; et al. FTY720 (Fingolimod) Sensitizes Prostate Cancer Cells to Radiotherapy by Inhibition of Sphingosine Kinase-1. Cancer Res. 2010, 70, 8651–8661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Venant, H.; Rahmaniyan, M.; Jones, E.E.; Lu, P.; Lilly, M.B.; Garrett-Mayer, E.; Drake, R.R.; Kraveka, J.M.; Smith, C.D.; Voelkel-Johnson, C. The Sphingosine Kinase 2 Inhibitor ABC294640 Reduces the Growth of Prostate Cancer Cells and Results in Accumulation of Dihydroceramides In Vitro and In Vivo. Mol. Cancer Ther. 2015, 14, 2744–2752. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, L.; Wang, J.; Xiong, H.; Wu, F.; Lan, T.; Zhang, Y.; Guo, X.; Wang, H.; Saleem, M.; Jiang, C.; et al. Co-targeting hexokinase 2-mediated Warburg effect and ULK1-dependent autophagy suppresses tumor growth of PTEN- and TP53- deficiency-driven castration-resistant prostate cancer. EBioMedicine 2016, 7, 50–61. [Google Scholar] [CrossRef] [Green Version]
  49. Beltran, H.; Oromendia, C.; Danila, D.C.; Montgomery, B.; Hoimes, C.; Szmulewitz, R.Z.; Vaishampayan, U.; Armstrong, A.J.; Stein, M.; Pinski, J.; et al. A Phase II Trial of the Aurora Kinase A Inhibitor Alisertib for Patients with Castration-resistant and Neuroendocrine Prostate Cancer: Efficacy and Biomarkers. Clin. Cancer Res. 2019, 25, 43–51. [Google Scholar] [CrossRef] [Green Version]
  50. Lin, J.; Patel, S.A.; Sama, A.R.; Hoffman-Censits, J.H.; Kennedy, B.; Kilpatrick, D.; Ye, Z.; Yang, H.; Mu, Z.; Leiby, B.; et al. A Phase I/II Study of the Investigational Drug Alisertib in Combination With Abiraterone and Prednisone for Patients With Metastatic Castration-Resistant Prostate Cancer Progressing on Abiraterone. Oncologist 2016, 21, 1296–1297. [Google Scholar] [CrossRef] [Green Version]
  51. Graff, J.N.; Higano, C.S.; Hahn, N.M.; Taylor, M.H.; Zhang, B.; Zhou, X.; Venkatakrishnan, K.; Leonard, E.J.; Sarantopoulos, J. Open-label, multicenter, phase 1 study of alisertib (MLN8237), an aurora A kinase inhibitor, with docetaxel in patients with solid tumors. Cancer 2016, 122, 2524–2533. [Google Scholar] [CrossRef]
  52. Crabb, S.J.; Birtle, A.J.; Martin, K.; Downs, N.; Ratcliffe, I.; Maishman, T.; Ellis, M.; Griffiths, G.; Thompson, S.; Ksiazek, L.; et al. ProCAID: A phase I clinical trial to combine the AKT inhibitor AZD5363 with docetaxel and prednisolone chemotherapy for metastatic castration resistant prostate cancer. Investig. New Drugs 2017, 35, 599–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kolinsky, M.; Rescigno, P.; Bianchini, D.; Zafeiriou, Z.; Mehra, N.; Mateo, J.; Michalarea, V.; Riisnaes, R.; Crespo, M.; Figueiredo, I.; et al. A phase I dose-escalation study of enzalutamide in combination with the AKT inhibitor AZD5363 (capivasertib) in patients with metastatic castration-resistant prostate cancer. Ann. Oncol. 2020, 31, 619–625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lewin, J.; Soria, J.-C.; Stathis, A.; Delord, J.-P.; Peters, S.; Awada, A.; Aftimos, P.G.; Bekradda, M.; Rezai, K.; Zeng, Z.; et al. Phase Ib Trial With Birabresib, a Small-Molecule Inhibitor of Bromodomain and Extraterminal Proteins, in Patients With Selected Advanced Solid Tumors. J. Clin. Oncol. 2018, 36, 3007–3014. [Google Scholar] [CrossRef]
  55. Madan, R.A.; Karzai, F.H.; Al Harthy, M.; Petrylak, D.P.; Kim, J.W.; Arlen, P.M.; Rosner, I.; Theoret, M.R.; Cordes, L.; Bilusic, M.; et al. Cabozantinib plus docetaxel and prednisone in metastatic castration-resistant prostate cancer. Br. J. Urol. 2021, 127, 435–444. [Google Scholar] [CrossRef] [PubMed]
  56. Whang, Y.E.; Armstrong, A.J.; Rathmell, W.K.; Godley, P.A.; Kim, W.Y.; Pruthi, R.S.; Wallen, E.M.; Crane, J.M.; Moore, D.T.; Grigson, G.; et al. A phase II study of lapatinib, a dual EGFR and HER-2 tyrosine kinase inhibitor, in patients with castration-resistant prostate cancer. Urol. Oncol. Semin. Orig. Investig. 2013, 31, 82–86. [Google Scholar] [CrossRef] [PubMed]
  57. Luo, J.-L.; Kamata, H.; Karin, M. The Anti-Death Machinery in IKK/NF-κB Signaling. J. Clin. Immunol. 2005, 25, 541–550. [Google Scholar] [CrossRef]
  58. Luo, J.-L.; Kamata, H.; Karin, M. IKK/NF- B signaling: Balancing life and death-a new approach to cancer therapy. J. Clin. Investig. 2005, 115, 2625–2632. [Google Scholar] [CrossRef] [Green Version]
  59. Taniguchi, K.; Karin, M. NF-κB, inflammation, immunity and cancer: Coming of age. Nat. Rev. Immunol. 2018, 18, 309–324. [Google Scholar] [CrossRef]
  60. Thomas-Jardin, S.; Dahl, H.; Nawas, A.F.; Bautista, M.; Delk, N.A. NF-κB signaling promotes castration-resistant prostate cancer initiation and progression. Pharmacol. Ther. 2020, 211, 107538. [Google Scholar] [CrossRef]
  61. Solis, M.; Romieu-Mourez, R.; Goubau, D.; Grandvaux, N.; Mesplede, T.; Julkunen, I.; Nardin, A.; Salcedo, M.; Hiscott, J. Involvement of TBK1 and IKKε in lipopolysaccharide-induced activation of the interferon response in primary human macrophages. Eur. J. Immunol. 2007, 37, 528–539. [Google Scholar] [CrossRef] [PubMed]
  62. Li, J.; Huang, J.; Jeong, J.-H.; Park, S.-J.; Wei, R.; Peng, J.; Luo, Z.; Chen, Y.T.; Feng, Y.; Luo, J.-L. Selective TBK1/IKKi dual inhibitors with anticancer potency. Int. J. Cancer 2014, 134, 1972–1980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kim, J.K.; Jung, Y.; Wang, J.; Joseph, J.; Mishra, A.; Hill, E.; Krebsbach, P.H.; Pienta, K.J.; Shiozawa, Y.; Taichman, R.S. TBK1 Regulates Prostate Cancer Dormancy through mTOR Inhibition. Neoplasia 2013, 15, 1064–1074. [Google Scholar] [CrossRef] [Green Version]
  64. Alam, M.; Hasan, G.M.; Hassan, I. A review on the role of TANK-binding kinase 1 signaling in cancer. Int. J. Biol. Macromol. 2021, 183, 2364–2375. [Google Scholar] [CrossRef] [PubMed]
  65. Bai, L.-Y.; Chiu, C.-F.; Kapuriya, N.P.; Shieh, T.-M.; Tsai, Y.-C.; Wu, C.-Y.; Sargeant, A.M.; Weng, J.-R. BX795, a TBK1 inhibitor, exhibits antitumor activity in human oral squamous cell carcinoma through apoptosis induction and mitotic phase arrest. Eur. J. Pharmacol. 2015, 769, 287–296. [Google Scholar] [CrossRef] [PubMed]
  66. Vu, H.L.; Aplin, A.E. Targeting TBK1 Inhibits Migration and Resistance to MEK Inhibitors in Mutant NRAS Melanoma. Mol. Cancer Res. 2014, 12, 1509–1519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Chin, K.-V.; Yang, W.-L.; Ravatn, R.; Kita, T.; Reitman, E.; Vettori, D.; Cvijic, M.E.; Shin, M.; Iacono, L. Reinventing the Wheel of Cyclic AMP. Ann. N. Y. Acad. Sci. 2002, 968, 49–64. [Google Scholar] [CrossRef]
  68. Hoy, J.J.; Parra, N.S.; Park, J.; Kuhn, S.; Iglesias-Bartolome, R. Protein kinase A inhibitor proteins (PKIs) divert GPCR-Gαs-cAMP signaling toward EPAC and ERK activation and are involved in tumor growth. FASEB J. 2020, 34, 13900–13917. [Google Scholar] [CrossRef]
  69. Holder, S.; Abdulkadir, S. PIM1 Kinase as a Target in Prostate Cancer: Roles in Tumorigenesis, Castration Resistance, and Docetaxel Resistance. Curr. Cancer Drug Targets 2014, 14, 105–114. [Google Scholar] [CrossRef]
  70. Linn, D.E.; Yang, X.; Xie, Y.; Alfano, A.; Deshmukh, D.; Wang, X.; Shimelis, H.; Chen, H.; Li, W.; Xu, K.; et al. Differential Regulation of Androgen Receptor by PIM-1 Kinases via Phosphorylation-dependent Recruitment of Distinct Ubiquitin E3 Ligases. J. Biol. Chem. 2012, 287, 22959–22968. [Google Scholar] [CrossRef] [Green Version]
  71. Njunge, L.W.; Estania, A.P.; Guo, Y.; Liu, W.; Yang, L. Tumor progression locus 2 (TPL2) in tumor-promoting Inflammation, Tumorigenesis and Tumor Immunity. Theranostics 2020, 10, 8343–8364. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, H.W.; Joo, K.M.; Lim, J.E.; Cho, H.J.; Cho, H.J.; Park, M.C.; Seol, H.J.; Seo, S.I.; Lee, J.-I.; Kim, S.; et al. Tpl2 Kinase Impacts Tumor Growth and Metastasis of Clear Cell Renal Cell Carcinoma. Mol. Cancer Res. 2013, 11, 1375–1386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. George, D.; Salmeron, A. Cot/Tpl-2 protein kinase as a target for the treatment of inflammatory disease. Curr. Top. Med. Chem. 2009, 9, 611–622. [Google Scholar] [CrossRef] [PubMed]
  74. Qiang, Z.; Zhou, Z.-Y.; Peng, T.; Jiang, P.-Z.; Shi, N.; Njoya, E.M.; Azimova, B.; Liu, W.-L.; Chen, W.-H.; Zhang, G.-L.; et al. Inhibition of TPL2 by interferon-α suppresses bladder cancer through activation of PDE4D. J. Exp. Clin. Cancer Res. 2018, 37, 288. [Google Scholar] [CrossRef] [Green Version]
  75. O’Regan, L.; Fry, A.M. The Nek6 and Nek7 Protein Kinases Are Required for Robust Mitotic Spindle Formation and Cytokinesis. Mol. Cell. Biol. 2009, 29, 3975–3990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Choudhury, A.; Schinzel, A.C.; Cotter, M.B.; Lis, R.T.; Labella, K.; Lock, Y.J.; Izzo, F.; Guney, I.; Bowden, M.; Li, Y.Y.; et al. Castration Resistance in Prostate Cancer Is Mediated by the Kinase NEK6. Cancer Res. 2016, 77, 753–765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zhu, Q.; Yang, J.; Han, S.; Liu, J.; Holzbeierlein, J.; Thrasher, J.B.; Li, B. Suppression of glycogen synthase kinase 3 activity reduces tumor growth of prostate cancer in vivo. Prostate 2011, 71, 835–845. [Google Scholar] [CrossRef]
  78. Schütz, S.V.; Schrader, A.J.; Zengerling, F.; Genze, F.; Cronauer, M.V.; Schrader, M. Inhibition of Glycogen Synthase Kinase-3β Counteracts Ligand-Independent Activity of the Androgen Receptor in Castration Resistant Prostate Cancer. PLoS ONE 2011, 6, e25341. [Google Scholar] [CrossRef] [Green Version]
  79. Barboro, P.; Benelli, R.; Tosetti, F.; Costa, D.; Capaia, M.; Astigiano, S.; Venè, R.; Poggi, A.; Ferrari, N. Aspartate β-hydroxylase targeting in castration-resistant prostate cancer modulates the NOTCH/HIF1α/GSK3β crosstalk. Carcinogenesis 2020, 41, 1246–1252. [Google Scholar] [CrossRef]
  80. Sun, H.; Zhang, L.; Shi, C.; Hu, P.; Yan, W.; Wang, Z.; Duan, Q.; Lu, F.; Qin, L.; Lu, T.; et al. TOPK is highly expressed in circulating tumor cells, enabling metastasis of prostate cancer. Oncotarget 2015, 6, 12392–12404. [Google Scholar] [CrossRef]
  81. Malumbres, M.; Barbacid, M. Mammalian cyclin-dependent kinases. Trends Biochem. Sci. 2005, 30, 630–641. [Google Scholar] [CrossRef] [PubMed]
  82. Chilà, R.; Guffanti, F.; Damia, G. Role and therapeutic potential of CDK12 in human cancers. Cancer Treat. Rev. 2016, 50, 83–88. [Google Scholar] [CrossRef] [PubMed]
  83. Reimers, M.A.; Yip, S.M.; Zhang, L.; Cieslik, M.; Dhawan, M.; Montgomery, B.; Wyatt, A.W.; Chi, K.N.; Small, E.J.; Chinnaiyan, A.M.; et al. Clinical Outcomes in Cyclin-dependent Kinase 12 Mutant Advanced Prostate Cancer. Eur. Urol. 2020, 77, 333–341. [Google Scholar] [CrossRef] [PubMed]
  84. Sava, G.; Fan, H.; Coombes, R.C.; Buluwela, L.; Ali, S. CDK7 inhibitors as anticancer drugs. Cancer Metastasis Rev. 2020, 39, 805–823. [Google Scholar] [CrossRef]
  85. Olson, C.M.; Jiang, B.; Erb, M.A.; Liang, Y.; Doctor, Z.M.; Zhang, Z.; Zhang, T.; Kwiatkowski, N.; Boukhali, M.; Green, J.L.; et al. Pharmacological perturbation of CDK9 using selective CDK9 inhibition or degradation. Nat. Chem. Biol. 2018, 14, 163–170. [Google Scholar] [CrossRef]
  86. Winter, G.E.; Mayer, A.; Buckley, D.L.; Erb, M.; Roderick, J.E.; Vittori, S.; Reyes, J.; di Iulio, J.; Souza, A.; Ott, C.J.; et al. BET Bromodomain Proteins Function as Master Transcription Elongation Factors Independent of CDK9 Recruitment. Mol. Cell 2017, 67, 5–18.e19. [Google Scholar] [CrossRef] [Green Version]
  87. Fernandez-Salas, E.; Wang, S.; Chinnaiyan, A.M. Role of BET proteins in castration-resistant prostate cancer. Drug Discov. Today: Technol. 2016, 19, 29–38. [Google Scholar] [CrossRef]
  88. Asangani, I.; Dommeti, V.L.; Wang, X.; Malik, R.; Cieslik, M.; Yang, R.; Escara-Wilke, J.; Wilder-Romans, K.; Dhanireddy, S.; Engelke, C.; et al. Therapeutic targeting of BET bromodomain proteins in castration-resistant prostate cancer. Nature 2014, 510, 278–282. [Google Scholar] [CrossRef]
  89. Shafran, J.S.; Andrieu, G.; Györffy, B.; Denis, G.V. BRD4 Regulates Metastatic Potential of Castration-Resistant Prostate Cancer through AHNAK. Mol. Cancer Res. 2019, 17, 1627–1638. [Google Scholar] [CrossRef] [Green Version]
  90. Wen, S.; He, Y.; Wang, L.; Zhang, J.; Quan, C.; Niu, Y.; Huang, H. Aberrant activation of super enhancer and choline metabolism drive antiandrogen therapy resistance in prostate cancer. Oncogene 2020, 39, 6556–6571. [Google Scholar] [CrossRef]
  91. Hong, Z.; Wu, G.; Xiang, Z.-D.; Xu, C.-D.; Huang, S.-S.; Li, C.; Shi, L.; Wu, D.-L. KDM5C is transcriptionally regulated by BRD4 and promotes castration-resistance prostate cancer cell proliferation by repressing PTEN. Biomed. Pharmacother. 2019, 114, 108793. [Google Scholar] [CrossRef] [PubMed]
  92. Guan, H.; You, Z.; Wang, C.; Fang, F.; Peng, R.; Mao, L.; Xu, B.; Chen, M. MicroRNA-200a suppresses prostate cancer progression through BRD4/AR signaling pathway. Cancer Med. 2019, 8, 1474–1485. [Google Scholar] [CrossRef]
  93. Nikhil, K.; Raza, A.; Haymour, H.S.; Flueckiger, B.; Chu, J.; Shah, K. Aurora Kinase A-YBX1 Synergy Fuels Aggressive Oncogenic Phenotypes and Chemoresistance in Castration-Resistant Prostate Cancer. Cancers 2020, 12, 660. [Google Scholar] [CrossRef] [Green Version]
  94. Nikhil, K.; Kamra, M.; Raza, A.; Haymour, H.; Shah, K. Molecular Interplay between AURKA and SPOP Dictates CRPC Pathogenesis via Androgen Receptor. Cancers 2020, 12, 3247. [Google Scholar] [CrossRef] [PubMed]
  95. Lin, J.-Z.; Wang, W.-W.; Hu, T.-T.; Zhu, G.-Y.; Li, L.-N.; Zhang, C.-Y.; Xu, Z.; Yu, H.-B.; Wu, H.-F.; Zhu, J.-G. FOXM1 contributes to docetaxel resistance in castration-resistant prostate cancer by inducing AMPK/mTOR-mediated autophagy. Cancer Lett. 2020, 469, 481–489. [Google Scholar] [CrossRef] [PubMed]
  96. Zeng, J.; Liu, W.; Fan, Y.-Z.; He, D.-L.; Li, L. PrLZ increases prostate cancer docetaxel resistance by inhibiting LKB1/AMPK-mediated autophagy. Theranostics 2018, 8, 109–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Mukherjee, R.; McGuinness, D.H.; McCall, P.; Underwood, M.; Seywright, M.; Orange, C.; Edwards, J. Upregulation of MAPK pathway is associated with survival in castrate-resistant prostate cancer. Br. J. Cancer 2011, 104, 1920–1928. [Google Scholar] [CrossRef]
  98. Cheung, S.; Jain, P.; So, J.; Shahidi, S.; Chung, S.; Koritzinsky, M. p38 MAPK Inhibition Mitigates Hypoxia-Induced AR Signaling in Castration-Resistant Prostate Cancer. Cancers 2021, 13, 831. [Google Scholar] [CrossRef]
  99. Li, Z.; Sun, C.; Tao, S.; Osunkoya, A.O.; Arnold, R.S.; Petros, J.A.; Zu, X.; Moreno, C.S. The JNK inhibitor AS602801 Synergizes with Enzalutamide to Kill Prostate Cancer Cells In Vitro and In Vivo and Inhibit Androgen Receptor Expression. Transl. Oncol. 2020, 13, 100751. [Google Scholar] [CrossRef]
  100. Taylor, B.S.; Schultz, N.; Hieronymus, H.; Gopalan, A.; Xiao, Y.; Carver, B.S.; Arora, V.K.; Kaushik, P.; Cerami, E.; Reva, B.; et al. Integrative Genomic Profiling of Human Prostate Cancer. Cancer Cell 2010, 18, 11–22. [Google Scholar] [CrossRef] [Green Version]
  101. Toren, P.; Kim, S.; Cordonnier, T.; Crafter, C.; Davies, B.R.; Fazli, L.; Gleave, M.E.; Zoubeidi, A. Combination AZD5363 with Enzalutamide Significantly Delays Enzalutamide-resistant Prostate Cancer in Preclinical Models. Eur. Urol. 2015, 67, 986–990. [Google Scholar] [CrossRef] [PubMed]
  102. Thomas, C.; Lamoureux, F.; Crafter, C.; Davies, B.R.; Beraldi, E.; Fazli, L.; Kim, S.; Thaper, D.; Gleave, M.; Zoubeidi, A. Synergistic Targeting of PI3K/AKT Pathway and Androgen Receptor Axis Significantly Delays Castration-Resistant Prostate Cancer Progression In Vivo. Mol. Cancer Ther. 2013, 12, 2342–2355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Huang, S.-H.; Tseng, J.-C.; Lin, C.-Y.; Kuo, Y.-Y.; Wang, B.-J.; Kao, Y.-H.; Muller, C.J.; Joubert, E.; Chuu, C.-P. Rooibos suppresses proliferation of castration-resistant prostate cancer cells via inhibition of Akt signaling. Phytomedicine 2019, 64, 153068. [Google Scholar] [CrossRef] [PubMed]
  104. Lombardi, A.; Cavalheiro, R.; Porto, C.; Vicente, C. Estrogen Receptor Signaling Pathways Involved in Invasion and Colony Formation of Androgen-Independent Prostate Cancer Cells PC-3. Int. J. Mol. Sci. 2021, 22, 1153. [Google Scholar] [CrossRef] [PubMed]
  105. Abdik, E.A.; Kaleagasioglu, F.; Abdik, H.; Sahin, F.; Berger, M.R. ABT-737 and erufosine combination against castration-resistant prostate cancer. Anti-Cancer Drugs 2019, 30, 383–393. [Google Scholar] [CrossRef] [PubMed]
  106. Strebhardt, K. Multifaceted polo-like kinases: Drug targets and antitargets for cancer therapy. Nat. Rev. Drug Discov. 2010, 9, 643–660. [Google Scholar] [CrossRef]
  107. Zhang, Z.; Hou, X.; Shao, C.; Li, J.; Cheng, J.-X.; Kuang, S.; Ahmad, N.; Ratliff, T.; Liu, X. Plk1 Inhibition Enhances the Efficacy of Androgen Signaling Blockade in Castration-Resistant Prostate Cancer. Cancer Res. 2014, 74, 6635–6647. [Google Scholar] [CrossRef] [Green Version]
  108. Li, J.; Karki, A.; Hodges, K.B.; Ahmad, N.; Zoubeidi, A.; Strebhardt, K.; Ratliff, T.L.; Konieczny, S.F.; Liu, X. Cotargeting Polo-Like Kinase 1 and the Wnt/β-Catenin Signaling Pathway in Castration-Resistant Prostate Cancer. Mol. Cell. Biol. 2015, 35, 4185–4198. [Google Scholar] [CrossRef] [Green Version]
  109. Nikhil, K.; Haymour, H.S.; Kamra, M.; Shah, K. Phosphorylation-dependent regulation of SPOP by LIMK2 promotes castration-resistant prostate cancer. Br. J. Cancer 2020, 124, 995–1008. [Google Scholar] [CrossRef]
  110. Nikhil, K.; Kamra, M.; Raza, A.; Shah, K. Negative cross talk between LIMK2 and PTEN promotes castration resistant prostate cancer pathogenesis in cells and in vivo. Cancer Lett. 2020, 498, 1–18. [Google Scholar] [CrossRef]
  111. Jiao, Q.; Bi, L.; Ren, Y.; Song, S.; Wang, Q.; Wang, Y.-S. Advances in studies of tyrosine kinase inhibitors and their acquired resistance. Mol. Cancer 2018, 17, 36. [Google Scholar] [CrossRef] [PubMed]
  112. Huang, L.; Jiang, S.; Shi, Y. Tyrosine kinase inhibitors for solid tumors in the past 20 years (2001–2020). J. Hematol. Oncol. 2020, 13, 1–23. [Google Scholar] [CrossRef]
  113. Mota, S.T.S.; Vecchi, L.; Alves, D.A.; Cordeiro, A.O.; Guimarães, G.S.; Campos-Fernández, E.; Maia, Y.C.P.; Dornelas, B.D.C.; Bezerra, S.M.; de Andrade, V.P.; et al. Annexin A1 promotes the nuclear localization of the epidermal growth factor receptor in castration-resistant prostate cancer. Int. J. Biochem. Cell Biol. 2020, 127, 105838. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, M.; Hu, Y.; Yu, T.; Ma, X.; Wei, X.; Wei, Y. Pan-HER-targeted approach for cancer therapy: Mechanisms, recent advances and clinical prospect. Cancer Lett. 2018, 439, 113–130. [Google Scholar] [CrossRef] [PubMed]
  115. Lee, R.J.; Smith, M.R. Targeting MET and Vascular Endothelial Growth Factor Receptor Signaling in Castration-Resistant Prostate Cancer. Cancer J. 2013, 19, 90–98. [Google Scholar] [CrossRef] [Green Version]
  116. Modena, A.; Massari, F.; Ciccarese, C.; Brunelli, M.; Santoni, M.; Montironi, R.; Martignoni, G.; Tortora, G. Targeting Met and VEGFR Axis in Metastatic Castration-Resistant Prostate Cancer: ‘Game Over’? Target. Oncol. 2016, 11, 431–446. [Google Scholar] [CrossRef]
  117. Zardan, A.; Nip, K.M.; Thaper, D.; Toren, P.; Vahid, S.; Beraldi, E.; Fazli, L.; Lamoureux, F.; Gust, K.M.; Cox, M.; et al. Lyn tyrosine kinase regulates androgen receptor expression and activity in castrate-resistant prostate cancer. Oncogenesis 2014, 3, e115. [Google Scholar] [CrossRef] [Green Version]
  118. Mahajan, K.; Coppola, D.; Rawal, B.; Chen, Y.; Lawrence, H.R.; Engelman, R.W.; Lawrence, N.J.; Mahajan, N.P. Ack1-mediated Androgen Receptor Phosphorylation Modulates Radiation Resistance in Castration-resistant Prostate Cancer. J. Biol. Chem. 2012, 287, 22112–22122. [Google Scholar] [CrossRef] [Green Version]
  119. Takaoka, A.; Tanaka, N.; Mitani, Y.; Miyazaki, T.; Fujii, H.; Sato, M.; Kovarik, P.; Decker, T.; Schlessinger, J.; Taniguchi, T. Protein tyrosine kinase Pyk2 mediates the Jak-dependent activation of MAPK and Stat1 in IFN-gamma, but not IFN-alpha, signaling. EMBO J. 1999, 18, 2480–2488. [Google Scholar] [CrossRef] [Green Version]
  120. Zhu, X.; Wen, Z.; Xu, L.Z.; Darnell, J. Stat1 serine phosphorylation occurs independently of tyrosine phosphorylation and requires an activated Jak2 kinase. Mol. Cell. Biol. 1997, 17, 6618–6623. [Google Scholar] [CrossRef] [Green Version]
  121. Gu, L.; Liao, Z.; Hoang, D.T.; Dagvadorj, A.; Gupta, S.; Blackmon, S.; Ellsworth, E.; Talati, P.; Leiby, B.; Zinda, M.; et al. Pharmacologic Inhibition of Jak2–Stat5 Signaling By Jak2 Inhibitor AZD1480 Potently Suppresses Growth of Both Primary and Castrate-Resistant Prostate Cancer. Clin. Cancer Res. 2013, 19, 5658–5674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Amos, S.-B.T.; Kalli, A.C.; Shi, J.; Sansom, M.S. Membrane Recognition and Binding by the Phosphatidylinositol Phosphate Kinase PIP5K1A: A Multiscale Simulation Study. Structure 2019, 27, 1336–1346.e2. [Google Scholar] [CrossRef] [Green Version]
  123. Pchejetski, D.; Böhler, T.; Stebbing, J.; Waxman, J. Therapeutic potential of targeting sphingosine kinase 1 in prostate cancer. Nat. Rev. Urol. 2011, 8, 569–578. [Google Scholar] [CrossRef] [PubMed]
  124. Malavaud, B.; Pchejetski, D.; Mazerolles, C.; de Paiva, G.R.; Calvet, C.; Doumerc, N.; Pitson, S.; Rischmann, P.; Cuvillier, O. Sphingosine kinase-1 activity and expression in human prostate cancer resection specimens. Eur. J. Cancer 2010, 46, 3417–3424. [Google Scholar] [CrossRef] [PubMed]
  125. Pchejetski, D.; Golzio, M.; Bonhoure, E.; Calvet, C.; Doumerc, N.; Garcia, V.; Mazerolles, C.; Rischmann, P.; Teissié, J.; Malavaud, B.; et al. Sphingosine Kinase-1 as a Chemotherapy Sensor in Prostate Adenocarcinoma Cell and Mouse Models. Cancer Res. 2005, 65, 11667–11675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Deng, Y.; Lu, J. Targeting hexokinase 2 in castration-resistant prostate cancer. Mol. Cell. Oncol. 2015, 2, e974465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Martin, P.L.; Yin, J.-J.; Seng, V.; Casey, O.; Corey, E.; Morrissey, C.; Simpson, R.M.; Kelly, K. Androgen deprivation leads to increased carbohydrate metabolism and hexokinase 2-mediated survival in Pten/Tp53-deficient prostate cancer. Oncogene 2016, 36, 525–533. [Google Scholar] [CrossRef]
  128. Feng, E.A.T.; Wang, J.; Cheng, K.; Lu, Q.; Zhao, R.; Wang, S.; Zhang, Q.; Ge, L.; Pan, J.; Song, G.; et al. IL13Rα1 prevents a castration resistant phenotype of prostate cancer by targeting hexokinase 2 for ubiquitin-mediated degradation. Cancer Biol. Med. 2021, 18, 1–23. [Google Scholar] [CrossRef]
  129. Lang, L.; Chemmalakuzhy, R.; Shay, C.; Teng, Y. PFKP Signaling at a Glance: An Emerging Mediator of Cancer Cell Metabolism. Adv. Exp. Med. Biol. 2019, 1134, 243–258. [Google Scholar] [CrossRef]
  130. Wu, M.-J.; Chen, C.-J.; Lin, T.-Y.; Liu, Y.-Y.; Tseng, L.-L.; Cheng, M.-L.; Chuu, C.-P.; Tsai, H.-K.; Kuo, W.-L.; Kung, H.-J.; et al. Targeting KDM4B that coactivates c-Myc-regulated metabolism to suppress tumor growth in castration-resistant prostate cancer. Theranostics 2021, 11, 7779–7796. [Google Scholar] [CrossRef]
  131. Hirano, G.; Izumi, H.; Yasuniwa, Y.; Shimajiri, S.; Ke-Yong, W.; Sasagiri, Y.; Kusaba, H.; Matsumoto, K.; Hasegawa, T. Involvement of riboflavin kinase expression in cellular sensitivity against cisplatin. Int. J. Oncol. 2011, 38, 893–902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Bitter, E.E.; Townsend, M.H.; Erickson, R.; Allen, C.; O’neill, K.L. Thymidine kinase 1 through the ages: A comprehensive review. Cell Biosci. 2020, 10, 1–16. [Google Scholar] [CrossRef] [PubMed]
  133. Chung, L.W.K.; Kao, C.; Sikes, R.A.; Zhau, H.E. Human prostate cancer progression models and therapeutic intervention. Mol. Biol. Prostate Cancer 1997, 43, 815–820. [Google Scholar] [CrossRef] [Green Version]
  134. Hanousková, L.; Řezáč, J.; Průša, R.; Kotaška, K. Thymidine Kinase-1 as Additional Diagnostic Marker of Prostate Cancer. Clin. Lab. 2020, 66. [Google Scholar] [CrossRef]
  135. Wang, M.; Zhao, J.; Zhang, L.; Wei, F.; Lian, Y.; Wu, Y.; Gong, Z.; Zhang, S.; Zhou, J.; Cao, K.; et al. Role of tumor microenvironment in tumorigenesis. J. Cancer 2017, 8, 761–773. [Google Scholar] [CrossRef]
  136. Li, K.; Tian, H. Development of small-molecule immune checkpoint inhibitors of PD-1/PD-L1 as a new therapeutic strategy for tumour immunotherapy. J. Drug Target. 2019, 27, 244–256. [Google Scholar] [CrossRef]
Figure 1. Diagram of prostate cancer progression and the activated/up-regulated kinases in primary prostate cancer (PPC), castration-resistant prostate cancer (CRPC), and metastatic CRPC (mCRPC).
Figure 1. Diagram of prostate cancer progression and the activated/up-regulated kinases in primary prostate cancer (PPC), castration-resistant prostate cancer (CRPC), and metastatic CRPC (mCRPC).
Pharmaceutics 14 00498 g001
Figure 2. Therapeutic strategies that target key kinases in CRPC. Note: 1: combined GSK461364A with JQ1; 2: WNY0824; 3: ursolic acid/apigenin/α-tomatine; 4: H89; 5: compound 9; 6: DHPCC-9; 7: OTS-514; 8: LY-2090314; 9: S1451/alisertib(MLN8237); 10: AS602801; 11: combined wortmannin with MK2206; 12: pso; 13: THZ1; 14: KB-0742; 15: JQ1; 16: MZ1; 17: MA4-022-1; 18: 6i/7m/(R)-12/ABBV-075/compound 7d; 19: AIM-100; 20: β-elemonic acid; 21: Gefitinib; 22: dacomitinib; 23: AD80; 24: ISA-2011B; 25: FTY720; 26: ABC294640; 27: 2-deoxyglucose.
Figure 2. Therapeutic strategies that target key kinases in CRPC. Note: 1: combined GSK461364A with JQ1; 2: WNY0824; 3: ursolic acid/apigenin/α-tomatine; 4: H89; 5: compound 9; 6: DHPCC-9; 7: OTS-514; 8: LY-2090314; 9: S1451/alisertib(MLN8237); 10: AS602801; 11: combined wortmannin with MK2206; 12: pso; 13: THZ1; 14: KB-0742; 15: JQ1; 16: MZ1; 17: MA4-022-1; 18: 6i/7m/(R)-12/ABBV-075/compound 7d; 19: AIM-100; 20: β-elemonic acid; 21: Gefitinib; 22: dacomitinib; 23: AD80; 24: ISA-2011B; 25: FTY720; 26: ABC294640; 27: 2-deoxyglucose.
Pharmaceutics 14 00498 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhong, S.; Peng, S.; Chen, Z.; Chen, Z.; Luo, J.-L. Choosing Kinase Inhibitors for Androgen Deprivation Therapy-Resistant Prostate Cancer. Pharmaceutics 2022, 14, 498. https://doi.org/10.3390/pharmaceutics14030498

AMA Style

Zhong S, Peng S, Chen Z, Chen Z, Luo J-L. Choosing Kinase Inhibitors for Androgen Deprivation Therapy-Resistant Prostate Cancer. Pharmaceutics. 2022; 14(3):498. https://doi.org/10.3390/pharmaceutics14030498

Chicago/Turabian Style

Zhong, Shangwei, Shoujiao Peng, Zihua Chen, Zhikang Chen, and Jun-Li Luo. 2022. "Choosing Kinase Inhibitors for Androgen Deprivation Therapy-Resistant Prostate Cancer" Pharmaceutics 14, no. 3: 498. https://doi.org/10.3390/pharmaceutics14030498

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