The protein kinase C-related kinase (PRK) family of serine/threonine kinases, PRK1, PRK2 and PRK3, are effectors for the Rho family small G proteins. An array of studies have linked these kinases to multiple signalling pathways and physiological roles, but while PRK1 is relatively well-characterized, the entire PRK family remains understudied. Here, we provide a holistic overview of the structure and function of PRKs and describe the molecular events that govern activation and autoregulation of catalytic activity, including phosphorylation, protein interactions and lipid binding. We begin with a structural description of the regulatory and catalytic domains, which facilitates the understanding of their regulation in molecular detail. We then examine their diverse physiological roles in cytoskeletal reorganization, cell adhesion, chromatin remodelling, androgen receptor signalling, cell cycle regulation, the immune response, glucose metabolism and development, highlighting isoform redundancy but also isoform specificity. Finally, we consider the involvement of PRKs in pathologies, including cancer, heart disease and bacterial infections. The abundance of PRK-driven pathologies suggests that these enzymes will be good therapeutic targets and we briefly report some of the progress to date.

The PRK serine/threonine kinase family comprises three members, PRK1 (also PKN/PKN1/PKNα), PRK2 (PKN2/PKNγ) and PRK3 (PKN3/PKNβ). PRK1 was identified from a human hippocampus library [1] and shown to have a catalytic domain related to protein kinase C. PRK2 and PRK3 were later identified, with 83% and 71% similarity to PRK1, respectively [2–4]. PRK1 and PRK2 are ubiquitously expressed [5–8], whereas PRK3, initially only detectable in metastatic cancers [4], was subsequently found at low levels in all tissues [8].

The PRKs are downstream effectors of Rho family small G proteins. These GTPases behave as molecular switches for signalling pathways and the Rho family are best known as orchestrators of actin cytoskeletal rearrangements, with diverse roles in cell migration [9,10], the cell cycle [11] and control of vesicular trafficking [12].

The PRKs have been detected in the cytosol [13–15], the nucleus [4,14,16,17] and at the plasma membrane [4,18]. PRK1 also localizes to endosomes in a RhoB-dependent manner [13,19], while PRK2 has been detected on membranes in lamellipodia [14] and in apical junctions [20], and PRK3 was found in the perinuclear Golgi [4,17].

Despite their importance in regulating key cellular processes, the PRKs are understudied kinases and our structural and functional understanding of them remains incomplete.

All PRKs have broadly the same domain structure (Figure 1).

PRK domain arrangement.

Figure 1.
PRK domain arrangement.

PRKs have three N-terminal HR1 domains (HR1a-c), a C2-like domain and a serine/threonine kinase domain. PRK2 and PRK3 have one and two proline-rich regions, respectively. The activation loop Thr on each PRK is labelled. The PRK1 HR1a pseudosubstrate, the PRK2 dimerization site and its PRK1/PRK3 equivalents, and the site of autoinhibition that is relieved by arachidonic acid, are indicated. The limits of the various domains are indicated above each sequence.

Figure 1.
PRK domain arrangement.

PRKs have three N-terminal HR1 domains (HR1a-c), a C2-like domain and a serine/threonine kinase domain. PRK2 and PRK3 have one and two proline-rich regions, respectively. The activation loop Thr on each PRK is labelled. The PRK1 HR1a pseudosubstrate, the PRK2 dimerization site and its PRK1/PRK3 equivalents, and the site of autoinhibition that is relieved by arachidonic acid, are indicated. The limits of the various domains are indicated above each sequence.

Close modal

HR1 domains

PRKs contain three N-terminal Homology Region 1 (HR1) domains, HR1a, HR1b and HR1c. RhoA is an interacting partner of the PRK1 HR1a domain [21–23] and all three members of the Rho subfamily, RhoA, RhoB and RhoC, bind the HR1a domains of PRK1, PRK2 and PRK3 in vitro[24]. RhoA binds weakly to PRK1 HR1b [23,25,26] and this appears to be independent of HR1a binding [25]. In contrast, RhoB may bind cooperatively to HR1a and HR1b, implying subtle differences in the Rho subfamily interactions [24].

Rac1 interacts with PRK2 [7] and with both PRK1 HR1a and HR1b [25], but there is no evidence for cooperative binding. The interaction of Rac1 with PRK3 has not been investigated but key interacting residues are conserved in the PRK1/PRK3 HR1 domains, suggesting an interaction is likely.

The HR1c domain does not appear to bind to RhoA [23] or Rac1 (unpublished observations) and no other binding partners have been reported. Currently, its function is a puzzle, although it is as well conserved as the other HR1 domains.

HR1a is an antiparallel coiled coil that binds RhoA [27]. Although there were two potential contact sites identified in the structure (Figure 2A), the mutational analysis suggested that contact site II is used in the solution [28]. The structure of Rac1 in complex with HR1b [29] showed that HR1b interacts with Rac1 at a site equivalent to the contact II site of RhoA (Figure 2B). The Rac1 C-terminal polybasic region folds back and interacts with both the G domain of Rac1 and with HR1b (Figure 2C). There is no structure of the PRK1 HR1c domain, but our NMR studies indicate that it also forms a coiled coil with an extra short, C-terminal α-helix [30].

Structures of HR1 domain complexes.

Figure 2.
Structures of HR1 domain complexes.

The Mg2+ on GTPases is shown as a magenta sphere and the nucleotide in a stick representation. The GTPase switch regions are coloured grey. (A) Crystal structure of RhoA interacting with the HR1a domain of PRK1, which forms an antiparallel coiled coil [27] (PDB:1CXZ). Two different contacts, I and II, were identified in the crystals. (B) Solution structure of Rac1 interacting with the HR1b domain of PRK1 [29] (PDB:2RMK). The polybasic region in the C-terminal tail of Rac1 is coloured blue. (C) Close-up of the Rac1 polybasic region in complex with PRK1 HR1b. The polyproline helix (179-PPP-181) provides the C-terminal tail with an extended conformation, allowing the polybasic region (182-VKKRKRK-188) to contact the HR1b domain. (D) Crystal structure of PRK1 HR1b interacting with leucine-rich repeats of the Salmonella protein SspH1 [134] (PDB:4NKG).

Figure 2.
Structures of HR1 domain complexes.

The Mg2+ on GTPases is shown as a magenta sphere and the nucleotide in a stick representation. The GTPase switch regions are coloured grey. (A) Crystal structure of RhoA interacting with the HR1a domain of PRK1, which forms an antiparallel coiled coil [27] (PDB:1CXZ). Two different contacts, I and II, were identified in the crystals. (B) Solution structure of Rac1 interacting with the HR1b domain of PRK1 [29] (PDB:2RMK). The polybasic region in the C-terminal tail of Rac1 is coloured blue. (C) Close-up of the Rac1 polybasic region in complex with PRK1 HR1b. The polyproline helix (179-PPP-181) provides the C-terminal tail with an extended conformation, allowing the polybasic region (182-VKKRKRK-188) to contact the HR1b domain. (D) Crystal structure of PRK1 HR1b interacting with leucine-rich repeats of the Salmonella protein SspH1 [134] (PDB:4NKG).

Close modal

PRK2 and PRK3 have similar HR1 sequences so they likely form similar structures [3,4]. Whether they interact with the same set of Rho family proteins in vivo is not clear, although the functional differences between the PRKs suggest that at least some interactions are unique.

It is unknown whether two Rho family proteins can bind simultaneously to PRK1. This would be structurally feasible, since there are relatively long linkers between the HR1 domains. Furthermore, structures of constructs containing two or more HR1 domains are not available, leaving potential inter-domain interactions unexplored.

The C2-like domain

PRKs include a C2-like domain with weak homology to the C2 domains of PKC proteins. These comprise 8 antiparallel β-sheets folded into a β-sandwich, which in classical PKCs bind to membrane phospholipids in a Ca2+-dependent manner [31]. In contrast, C2 domains of novel and atypical PKCs are calcium-insensitive and bind acidic phospholipids. They are, therefore, critical for the recruitment of proteins to the plasma membrane [31]. No structural data is available for the C2-like domains of PRKs, however, they are likely to be insensitive to Ca2+ but bind lipids to activate the PRKs [32–34].

Proline-rich regions

PRK2 has a proline-rich region with a class II motif (PPPAPPR) between the C2 and kinase domains, which interacts with SH3 domains in the NCK/NCK2 adaptor proteins and more weakly with the SH3 domain of phospholipase Cγ [6,35]. PRK3 has a similar motif (PPPKPPR) and also a class I motif (RRGPSPP). These are involved in protein–protein interactions with SH3 domains of the RhoGAP Graf and Graf2 proteins [4,36].

Kinase domain

The PRKs are members of the AGC kinase superfamily and the PRK1 kinase domain structure has been solved alone and with inhibitors [37] (Figure 3A). A Thr in the activation loop is phosphorylated by PDK1 (3-phosphoinositide dependent kinase) [38,39] to activate the kinase. The kinase domain has a C-terminal extension that encircles the catalytic domain and contains three conserved features found in AGC kinases, which govern the allosteric regulation of their catalytic domain (Figure 3A, Table 1):

  • (1) An NFD motif, whose phenylalanine contacts the adenine ring of the substrate ATP.

  • (2) A turn motif (TM) C-terminal to the NFD [40,41], which is phosphorylated to enhance intramolecular interactions involving the hydrophobic motif (HM) [37,42].

  • (3) HM at the extreme C-terminus. Phosphorylated HM interacts with a region of the N-lobe known as the PDK1-interacting fragment (PIF) pocket, which transmits conformational changes to the active site, allosterically activating the kinase [39]. The PRK HM includes a phosphomimic Asp residue, allowing constitutive interaction with the PIF pocket.

Table 1.
Conserved motifs in the kinase domain and C-terminal extension of the PRKs
ProteinActivation loopNFDTMHM
PRK1 T774 903-NFD-905 S916 932-FLDFDF-937 
PRK2 T816 945-NFD-947 T958 974-FRDFDY-979 
PRK3 T718 847-YFE-849 T860 876-FRDFDF-881 
ProteinActivation loopNFDTMHM
PRK1 T774 903-NFD-905 S916 932-FLDFDF-937 
PRK2 T816 945-NFD-947 T958 974-FRDFDY-979 
PRK3 T718 847-YFE-849 T860 876-FRDFDF-881 

Phosphorylation of the activation loop threonine by PDK1 is aided by interactions between the PRK HM motif and the PDK1 PIF pocket, which activates PDK1 and brings the two proteins into close proximity. However, this intermolecular interaction prevents the immediate allosteric activation of the PRK kinase. This requires phosphorylation of the PRK TM, leading to the release of PDK1, PRK intramolecular HM/PIF pocket interaction and kinase activation [17,43–45]. The mammalian target of rapamycin (mTOR) 2, a protein complex with diverse signalling roles, phosphorylates the PRK1 TM [43], as does cyclin-dependent kinase (CDK) 1 [45], suggesting multiple pathways feed into PRK activation.

The PRK kinase domain and its activation.

Figure 3.
The PRK kinase domain and its activation.

(A) Crystal structure of the PRK1 kinase domain (apoprotein) [37] (PDB:4OTD). Key residues within the activation loop (cyan), hydrophobic motif (red), turn motif (purple) and NFD motif (green) are shown as sticks on the structure. The active site, ATP-binding pocket and the allosteric regulation site, the PIF pocket, are labelled. (B) PRK1 activation events. The events have been placed in a speculative temporal order partly based on work on PRK2 dimerization studies [51] (1–7). (1) Dimerization keeps PRK1 inactive by allosterically preventing the intramolecular HM/PIF pocket interaction. (2) Recruitment to the cell membrane via the C2-like domain may allow lipids such as arachidonic acid to disrupt this dimer. (3) HM phosphomimic can bind to the PDK1 PIF pocket. (4) PDK1 phosphorylates the activation loop. (5) TM phosphorylation by mTOR2/CDK1 releases PDK1 by promoting the intramolecular HM/PIF pocket interaction. (6) RhoA binds to the HR1a domain and relieves pseudosubstrate autoinhibition. (7) PRK1 autophosphorylates leading to additional activation. Phosphorylated residues [57] are denoted with a star.

Figure 3.
The PRK kinase domain and its activation.

(A) Crystal structure of the PRK1 kinase domain (apoprotein) [37] (PDB:4OTD). Key residues within the activation loop (cyan), hydrophobic motif (red), turn motif (purple) and NFD motif (green) are shown as sticks on the structure. The active site, ATP-binding pocket and the allosteric regulation site, the PIF pocket, are labelled. (B) PRK1 activation events. The events have been placed in a speculative temporal order partly based on work on PRK2 dimerization studies [51] (1–7). (1) Dimerization keeps PRK1 inactive by allosterically preventing the intramolecular HM/PIF pocket interaction. (2) Recruitment to the cell membrane via the C2-like domain may allow lipids such as arachidonic acid to disrupt this dimer. (3) HM phosphomimic can bind to the PDK1 PIF pocket. (4) PDK1 phosphorylates the activation loop. (5) TM phosphorylation by mTOR2/CDK1 releases PDK1 by promoting the intramolecular HM/PIF pocket interaction. (6) RhoA binds to the HR1a domain and relieves pseudosubstrate autoinhibition. (7) PRK1 autophosphorylates leading to additional activation. Phosphorylated residues [57] are denoted with a star.

Close modal

Kitagawa et al. [46] first suggested autoregulation by the N-terminus when they showed that a peptide corresponding to amino acids 39–53 of PRK1 HR1a, with Ile46 mutated to serine, was phosphorylated by PRK1. They suggested that HR1a contains a pseudosubstrate motif that autoinhibits the kinase, which is released when RhoA binds to HR1a. The PRK1 consensus is not well-defined, however, and the pseudosubstrate hypothesis has not been confirmed [47,48]. RhoA may enhance PRK activation in other ways, e.g. when PRK1 binds to RhoA, activation loop phosphorylation by PDK1 increases [49]. The Drosophila orthologue Pkn is activated by RhoA and Rac1 [50], while RhoC has also been linked to increased PRK3 activity [17], but exactly how these Rho GTPases achieve this is not understood.

The PRK2 N-terminus inhibits the catalytic domain by preventing its interaction with PDK1 [51]; this may be relieved by RhoA binding. In addition, PRK2 activity appears to be inhibited in trans by intermolecular dimer formation, mediated by residues 464–500 (Figure 1). Peptides based on this region inhibit the catalytic activity of all three PRKs [48,51,52]. It is thought that when the intramolecular HM/PIF pocket interaction is stabilized, the ensuing conformational changes displace the dimerization motif, leading to PRK monomerization and activation. The intramolecular HM/PIF pocket and dimerization interactions are, therefore, allosterically mutually exclusive, with opposing effects on catalytic activity. Intermolecular interactions involving other regions in the N-terminus have been suggested [51], implying that PRK dimerization may be mediated by multiple, cooperative events.

PRK1 and PRK2 can also be activated by proteolysis, e.g. by caspase-3 during apoptosis [53–55]. The constitutively active PRK1 fragment does not seem to play a direct role in apoptotic events but may be involved in morphological changes observed in microglia [56].

PRK1 is activated by phosphatidylinositol 4,5-bisphosphate (PIP2), phosphatidylinositol 3,4,5-triphosphate (PIP3) [32] and arachidonic acid [53]. Arachidonic acid is thought to remove the autoinhibition exhibited by a region of PRK1 and PRK2 that overlaps with or is adjacent to the C2-like domain [57], highlighting the need for the PRKs to be recruited to a membrane to be successfully activated. Arachidonic acid also activates PRK3, although its activation of PRK2/PRK3 is less potent than that of PRK1 [4,52,58]. PRKs are activated by other lipids, including linoleic acid, oleic acid and cardiolipin [5,58–61], and the isoforms show distinct lipid activation profiles and possibly activation mechanisms [52]. Phospholipids also bind directly to PRK1 HR1a, suggesting that HR1 domains may be involved in membrane binding [62].

Taken together, it is clear that PRK regulation is complex, involving a combination of Rho protein binding to at least two HR1 domains, competition between inter- and intramolecular interactions, oligomerization, autophosphorylation and lipid binding. This complexity allows PRKs to be activated in response to many different signalling cues, or in multiple locations, and opens the possibility of PRKs having multiple levels of activation, rather than being a binary ‘on-off' switch. Figure 3B postulates a temporal order of activation.

Many physiological roles of PRKs have been elucidated by identifying direct substrates for their enzyme activity [47,63] and Table 2 lists some of these. PRK1 is the best-characterized isoform and its roles are summarized in Figure 4.

Physiological roles of PRK1.

Figure 4.
Physiological roles of PRK1.

The key activatory pathways are highlighted. The various processes in which PRK1 is implicated are highlighted in different colours.

Figure 4.
Physiological roles of PRK1.

The key activatory pathways are highlighted. The various processes in which PRK1 is implicated are highlighted in different colours.

Close modal
Table 2.
Key PRK substrates. The PRK isoform, substrate and corresponding cellular process/function are listed
PRKSubstrateCell process/functionReference
PRK1 α-actinin Cytoskeletal regulation [64
PRK1 vimentin Cytoskeletal regulation [71
PRK2 Cortactin Cytoskeletal regulation [66
PRK3 p130cas Cytoskeletal regulation [123
PRK1, PRK2 HDAC5, HDAC7, HDAC9 Chromatin remodelling [88
PRK1 H3 Thr11 Androgen receptor signalling [85
PRK1, PRK2 Cdc25 Cell cycle [15,156
PRK1, PRK2 Pyrin Immune response [98
PRK1 RHP3A Vesicular trafficking [157
PRK1 GSK-3β Glucose metabolism [104
PRK1, PRK2 PI3KC2-β Nutrient signalling [106
PRK2 DUSP6 Inhibition of transcription [120
PRK1 Tau Microtubule disruption [73,74,104
PRK1 α Crystallin B Molecular chaperone [125
PRK1 HPV E6 oncoprotein Cell immortalization [158
PRK2 HCV RNA Pol HCV replication [136,138
PRKSubstrateCell process/functionReference
PRK1 α-actinin Cytoskeletal regulation [64
PRK1 vimentin Cytoskeletal regulation [71
PRK2 Cortactin Cytoskeletal regulation [66
PRK3 p130cas Cytoskeletal regulation [123
PRK1, PRK2 HDAC5, HDAC7, HDAC9 Chromatin remodelling [88
PRK1 H3 Thr11 Androgen receptor signalling [85
PRK1, PRK2 Cdc25 Cell cycle [15,156
PRK1, PRK2 Pyrin Immune response [98
PRK1 RHP3A Vesicular trafficking [157
PRK1 GSK-3β Glucose metabolism [104
PRK1, PRK2 PI3KC2-β Nutrient signalling [106
PRK2 DUSP6 Inhibition of transcription [120
PRK1 Tau Microtubule disruption [73,74,104
PRK1 α Crystallin B Molecular chaperone [125
PRK1 HPV E6 oncoprotein Cell immortalization [158
PRK2 HCV RNA Pol HCV replication [136,138

Cytoskeletal regulation and cell adhesion

PRK1 interacts with the actin cross-linking protein α-actinin [64] and its overexpression in fibroblasts resulted in actin reorganization [65]. Cortactin, which cross-links F-actin, is inhibited by Rac1-activated PRK2, leading to key cytoskeletal rearrangements in astrocyte migration [66]. The PRKs also regulate the formation of stress fibres, actin-rich bundles essential for cell migration. Exogenous expression of PDK1 and PRK1 in mammalian cells led to insulin-induced actin cytoskeletal reorganization and a reduction in the number of stress fibres as well as membrane ruffling [67]. Expression of active PRK2 caused the formation of fewer, larger stress fibres but its kinase-dead mutant led to their almost complete loss [7]. Finally, RNAi-mediated knockdown of PRK3 in human umbilical endothelial cells (HUVECs) compromised their migration [68] and their ability to form stress fibres downstream of TNF-α [69]. More recently, a link between membrane tension, RhoA and ROCK1/PRK2 provided a mechanism for rear retraction and therefore directional migration [70]. Together, these studies suggest that PRKs are involved in the modulation of actin stress fibres and the cellular reorganization of F-actin, although their precise roles may differ.

As well as regulating the actin cytoskeleton, PRK1 phosphorylates the intermediate filament protein vimentin, leading to vimentin filament disruption in vitro[71], while PRK3 knockdown also led to vimentin disruption [17]. PRK1 also disrupts neurofilament assembly and axonal transport [72]. In neurons affected with Alzheimer's disease, PRK1 has been linked to neurofibrillary tangles and their key component, the protein tau. PRK1 phosphorylates the tau protein leading to disruption of the microtubule assembly [73,74].

In osteoclasts, PRK3 is activated downstream of the Wnt5–RhoA signalling axis and binds the non-receptor tyrosine kinases c-Src and Pyk2 (Figure 5A). This activates c-Src, which is essential for actin ring formation, a requirement for osteoclast bone resorption [75,76]. Therefore, PRK3 has a role in bone resorption in vivo and could be a therapeutic target for bone-loss diseases.

Detailed insight into selected PRK pathways.

Figure 5.
Detailed insight into selected PRK pathways.

(A) The role of PRK3 in bone resorption by osteoclasts. PRK3 is activated downstream of the Wnt5/RhoA signalling axis and it forms a complex via its proline-rich region with the non-receptor tyrosine kinases c-Src and Pyk2. This leads to c-Src activation, which is essential for actin ring formation, a requirement for bone resorption in osteoclasts. (B) The role of PRK1 in AR signalling and chromatin remodelling. The steroid hormone testosterone (T) is converted into dihydrotestosterone (DHT) which binds to the androgen receptor (AR). The AR interacts with PRK1 and the complex translocates to the nucleus where it binds to androgen-responsive elements (AREs). PRK1 phosphorylates Thr11 of Histone3, recruiting the demethylase JMJD2C and leading to increased transcription of target genes. The thromboxane receptors TPα and TPβ, activated by thromboxane A2 (TXA2), induce Histone3 Thr11 phosphorylation by PRK1 and enhance recruitment of the AR, which also leads to Histone3 Thr11 phosphorylation. (C) PRK2 phosphorylates and activates the Cdc25 phosphatase, which in turn activates CDK1/cyclin B to enter mitosis. CDK1/cyclin B then phosphorylates PRK2 which is also important during cytokinesis. D) PRK1-dependent down-regulation of the pyrin inflammasome. Pyrin is part of the pyrin inflammasome which promotes IL-1β release, leading to inflammation during the innate immune response. PRK1 phosphorylates pyrin causing it to be sequestered by 14-3-3 proteins.

Figure 5.
Detailed insight into selected PRK pathways.

(A) The role of PRK3 in bone resorption by osteoclasts. PRK3 is activated downstream of the Wnt5/RhoA signalling axis and it forms a complex via its proline-rich region with the non-receptor tyrosine kinases c-Src and Pyk2. This leads to c-Src activation, which is essential for actin ring formation, a requirement for bone resorption in osteoclasts. (B) The role of PRK1 in AR signalling and chromatin remodelling. The steroid hormone testosterone (T) is converted into dihydrotestosterone (DHT) which binds to the androgen receptor (AR). The AR interacts with PRK1 and the complex translocates to the nucleus where it binds to androgen-responsive elements (AREs). PRK1 phosphorylates Thr11 of Histone3, recruiting the demethylase JMJD2C and leading to increased transcription of target genes. The thromboxane receptors TPα and TPβ, activated by thromboxane A2 (TXA2), induce Histone3 Thr11 phosphorylation by PRK1 and enhance recruitment of the AR, which also leads to Histone3 Thr11 phosphorylation. (C) PRK2 phosphorylates and activates the Cdc25 phosphatase, which in turn activates CDK1/cyclin B to enter mitosis. CDK1/cyclin B then phosphorylates PRK2 which is also important during cytokinesis. D) PRK1-dependent down-regulation of the pyrin inflammasome. Pyrin is part of the pyrin inflammasome which promotes IL-1β release, leading to inflammation during the innate immune response. PRK1 phosphorylates pyrin causing it to be sequestered by 14-3-3 proteins.

Close modal

Given their involvement in cytoskeletal regulation, it is logical that the PRKs have also been linked to cell adhesion. PRK3-depleted HUVECs showed the irregular distribution of VE-cadherin/β-catenin, which link adherens junctions to the actin cytoskeleton [69], and a reduction in ICAM-1 and in Pyk2 phosphorylation, both of which are involved in cell adhesion. A study using PRK3 knockout mice implicated PRK3 in the regulation of glycosylation of ICAM-1 and integrins, suggesting that PRK3 may directly affect cell adhesion [8]. In addition, the PRK3 knockout mice showed decreased micro-vessel sprouting, while MEFs displayed reduced migration which was not rescued by growth factors. Taken together, these results suggest that PRK3 has a unique physiological role in angiogenesis that is linked to cell adhesion. In this setting, PRK3 bound preferentially to RhoC, rather than RhoA/B suggesting that RhoC plays a role in angiogenesis. The PRK3 knockout mice showed reduced metastasis but no change in tumour angiogenesis, so its role in metastasis is not through increased angiogenesis.

PRK1 and PRK2 also play a part in cell adhesion. ROCK/PRK1 disrupt adhesion of endothelial cells downstream of thrombin [77]. PRK1 regulates both tight junctions [78] and adherens junctions; in the latter, PRK1 kinase activity was necessary for cell-surface expression of N-cadherin and integrins in fibroblasts [79]. In contrast, PRK2 regulates apical junction formation in bronchial endothelial cells [20], while RhoA-activated PRK2 induces Fyn/Src-dependent phosphorylation of β/γ-catenin and p120ctn, promoting keratinocyte cell–cell adhesion [80]. The role of PRK2 in cell adhesion has been further demonstrated in myoblasts [81], where it interacts directly with Cdo, a cell-surface protein, Akt and the adaptor protein APPL1, leading to Akt activation and differentiation. This is in contrast with PRK1, which is a negative regulator of Akt [82–84].

Androgen receptor signalling, chromatin remodelling and transcriptional regulation

Metzger et al.[16] implicated PRK1 and PRK2 in androgen receptor (AR) signalling (Figure 5B), showing that PRK1 binds to the transactivation domain of the AR in vitro and that the PRK1/AR complex binds androgen response elements of AR-regulated genes. A subsequent study [85] showed that PRK1 phosphorylates Thr11 of histone H3. H3T11 phosphorylation recruits JMJD2C, which demethylates H3 Lys9, activating AR-dependent gene expression. All three PRKs are activated downstream of the thromboxane receptors TPα and TPβ [86]. PRK1/PRK2 activation in thromboxane signalling is crucial for chromatin remodelling through phosphorylation of H3T11 (Figure 5B) and enhanced recruitment of the AR [87]. The histone deacetylases HDAC5/7/9 are phosphorylated by PRK1/PRK2 within their nuclear localization signal [88]. This is thought to reduce nuclear import by allowing the binding of 14-3-3 scaffolding proteins to HDACs, thereby preventing the HDACs accessing the nucleus to repress transcription. Overall, the PRKs directly affect chromatin structure leading to transcriptional up-regulation.

PRK1 has been implicated in regulating transcription factors and is localized to the nucleus under stress conditions [89], where it interacts with a neurone-specific helix-loop-helix transcription factor, NDRF2, to amplify transcription [90]. PRK1 also stimulates expression of atrial natriuretic factor (ANF) in cardiomyocytes [91], binds to cyclin T2a to enhance myoD-dependent transcription [92] and up-regulates transcription of focal adhesion proteins such as paxillin [93].

Cell cycle regulation

PRK1 and PRK2 move to the cleavage furrow of HeLa cells during telophase and the midbody during cytokinesis [15]. Furthermore, siRNA-depletion of PRK2 led to failed abscission, suggesting a direct role in cytokinesis. PRK2 depletion delays the G2/M transition, suggesting a role in mitotic entry (Figure 5C) and the pro-mitotic Cdc25 phosphatase is activated by PRK2 phosphorylation [15]. PRK1 and PRK2 are, therefore, involved in both mitotic entry and exit in human cells.

The G1 phase cyclinD1–CDK6 complex phosphorylates PRK1 and modulates vascular smooth muscle cell division and migration [94,95]. RhoA was shown to transcriptionally regulate cyclinD1 expression via PRK1 in human embryonic stem cells leading to cell proliferation [96]. This suggests that a feedback loop exists between PRK1 and cyclinD1.

The immune response

Pyrin is involved in the formation of a protein complex known as the pyrin inflammasome, part of the innate immune system that forms in response to bacterial infection and induces interleukin (IL) release from macrophages to combat infection. Some toxins, such as Clostridium botulinum C3 and Clostridium difficile TcdB inactivate RhoA, activating the inflammasome and therefore the innate immune response [97]. Depletion of PRK1 and PRK2 in macrophages leads to inflammasome-dependent IL-1β release, suggesting that these PRKs inhibit pyrin inflammasome activation [98]. The PRKs phosphorylate pyrin, leading to recruitment of 14-3-3 proteins and inhibition of inflammasome activation (Figure 5D). Mutation in the pyrin gene in patients with the autoinflammatory disease Familial Mediterranean Fever result in a pyrin variant with reduced binding to PRKs and increased inflammasome activation. Another autoinflammatory disease, hyperimmunoglobulinemia D syndrome (HIDS), is caused by mutations in the gene for mevalonate kinase, which is necessary for the correct lipid modification and therefore localization of RhoA, underlining the importance of the RhoA-PRK signalling axis in inflammasome regulation.

PRK1 has been implicated in lymphocyte trafficking by studies with a PRK1 kinase-inactive knock-in mouse [99], whose lymphocytes displayed reduced migration from secondary lymphoid organs to the blood and lymph. The same mouse model indicated that PRK1 also regulates trafficking of all leukocytes [100].

Glucose metabolism and nutrient signalling

Actin cytoskeletal reorganization and membrane ruffling was observed in adipocytes and fibroblasts expressing the insulin receptor upon insulin treatment or when PRK1 or PDK1 were expressed [67]. Non-phosphorylatable PRK1 or kinase-dead PDK1 had no effect, suggesting that activation of PRK1 via PDK1 is vital for signal transduction from the insulin receptor to the actin cytoskeleton.

The effect of PRK1 on insulin-stimulated glucose transport was dependent on both active RhoA and phosphoinositide-3-kinase (PI3K) [101], and expression of PRK1 and active RhoA increased the levels of GLUT4 glucose transporters translocated to the membrane (Figure 6A). Both PRK1 and PRK3 interact with phospholipase D1 [102,103], which could provide the link between PRK1 and GLUT4 vesicle translocation. PRK1 also phosphorylates and inhibits glycogen synthase kinase, GSK-3β, leading to activation of glycogen synthase [104].

The role of PRK1 in glucose metabolism and nutrient signalling.

Figure 6.
The role of PRK1 in glucose metabolism and nutrient signalling.

(A) The role of PRK1 in glucose metabolism. (1) Insulin binds to its receptor initiating a signal cascade (2) involving the insulin receptor substrate 1 (IRS1)-dependent activation of phosphoinositide 3-kinase (PI3K). This activates PDK1 which in turn activates PRK1. PRK1 interacts with phospholipase D1 (PLD1) to enhance the translocation (3) of GLUT4 receptors to the membrane permitting glucose entry (4). PRK1 also inhibits GSK-3β, an inhibitor of glycogen synthase (GS), to promote glycogen synthesis (5). B) The role of PRK1 in nutrient signalling. PRK1 and PRK2, which are activated by mTOR2, phosphorylate PI3KC2-β, a class II PI3K. This enzyme catalyzes PIP2 formation causing the sequestration of PI3KC2-β by 14-3-3 proteins and subsequent mTOR1 activation when nutrients are abundant. When nutrients are scarce, PI3KC2-β activity leads to mTOR1 inhibition.

Figure 6.
The role of PRK1 in glucose metabolism and nutrient signalling.

(A) The role of PRK1 in glucose metabolism. (1) Insulin binds to its receptor initiating a signal cascade (2) involving the insulin receptor substrate 1 (IRS1)-dependent activation of phosphoinositide 3-kinase (PI3K). This activates PDK1 which in turn activates PRK1. PRK1 interacts with phospholipase D1 (PLD1) to enhance the translocation (3) of GLUT4 receptors to the membrane permitting glucose entry (4). PRK1 also inhibits GSK-3β, an inhibitor of glycogen synthase (GS), to promote glycogen synthesis (5). B) The role of PRK1 in nutrient signalling. PRK1 and PRK2, which are activated by mTOR2, phosphorylate PI3KC2-β, a class II PI3K. This enzyme catalyzes PIP2 formation causing the sequestration of PI3KC2-β by 14-3-3 proteins and subsequent mTOR1 activation when nutrients are abundant. When nutrients are scarce, PI3KC2-β activity leads to mTOR1 inhibition.

Close modal

In skeletal muscle cells, PRK2 responds to insulin stimulation by promoting glucose uptake, glycogen synthesis and glucose oxidation [105]. Here, PRK2 knockdown led to reduced mTOR phosphorylation and protein synthesis and increased AMP-activated protein kinase (AMPK) signalling, suggesting that PRK2 inhibits AMPK and promotes energy storage and cell growth.

PRK1 and PRK2 both phosphorylate PI3KC2-β, a class II PI3K which catalyzes PIP2 formation, leading to mTOR1 inactivation [106]. Phosphorylated PI3KC2-β is sequestered by 14-3-3, allowing mTOR1 activation when nutrients and growth factors are abundant. Therefore, mTOR2, which phosphorylates the TM of PRK1 and PRK2 activating them, promotes mTOR1-dependent nutrient signalling in response to growth factors and nutrients (Figure 6B). Since PRK2 also activates Akt [81], an activator of mTOR1, it appears PRK2 has a synergistic effect on cell growth. All the above may be under the control of class I PI3Ks, which can activate PRK1 and PRK2 via mTOR2-dependent TM phosphorylation and PI3K/PDK1-dependent activation loop phosphorylation [43].

Development

PRK2, but not PRK1/PRK3, is essential during mouse embryogenesis. PRK2 knockout embryos showed severe cardiovascular abnormalities, a collapsed mesenchyme, insufficient axial turning, a neural crest migration defect [107] and neural tube closure defects [108], effects which were unobserved with PRK1 and PRK3 knockouts.

Recently, PRK1 has been identified as a key player in mouse cerebellar development [83], indicated by PRK1 knockout mice having problems with axonal outgrowth and with synapse formation in Purkinje cells. Additionally, the Drosophila melanogaster orthologue dPkn affects the spatiotemporal regulation of myosin, leading to its relocalization, an event thought to promote the asymmetric division of neural stem cells [109]. A mutation in the dPkn gene has also been linked to wing morphogenesis in the context of the Jun-terminal kinase pathway [110], while dPkn activity has also been associated with negative regulation of actin-myosin contraction in other developmental processes [111] and with dorsal closure [50].

Cancer

PRK1 is overexpressed in human prostate cancer [16] and promotes transcription of AR-regulated genes by phosphorylating histone H3T11, an epigenetic marker of prostate cancer [85]. The importance of PRK1 in prostate cancer is further emphasized as its inhibition abrogates the proliferation of AR-induced tumour cells. In addition, PRK1 and PRK2, activated by thromboxane receptors, mimic and enhance androgen-dependent chromatin remodelling and promote the migration and proliferation of prostate cancer cells [87]. PRK1 is a key driver of metastasis of androgen-independent prostate cancer cells in vivo, based on both knockdown of PRK1 or use of a PRK1 inhibitor [93]. Additionally, PRK3, downstream of PI3K, is necessary for the migration of PC-3 cells, while PRK3 inhibition abrogated lymph node metastasis in an orthotopic mouse prostate tumour model [112]. Contrastingly PRK1 and PRK2, but not PRK3, were important for migration in PC-3 cells when the PRKs are activated downstream of the TPα and TPβ receptors [86,87].

PRK1 mRNA is abundant in several other malignant tissues, particularly ovarian cancer, suggesting that PRK1 is implicated in several cancers [113]. A PRK1 mutation (E216K) in the HR1c domain has been described in rhabdomyosarcoma and linked to inhibition of myogenic differentiation [114], while a nuclear factor 1 X-type (NFIX)-PRK1 fusion has been described in a patient with secretory skin carcinoma [115]. Other PRK1 fusions have been described in melanocytoma and lung squamous cell and hepatocellular carcinomas [116,117], which lack the N-terminal domains and are, therefore, presumed to be constitutively activated.

PRK2, and to a lesser extent PRK1, is vital for cell migration and invasion of a bladder tumour cell line [118]. Consistent with this, silencing of PRK2 led to decreased cell proliferation, migration and colony formation in smoke-exposed keratinocytes, while in unsilenced cells PRK2 was overexpressed and hyperphosphorylated at the activation loop threonine [119]. In contrast, in colon cancer cells PRK2 activated the DUSP6 phosphatase, reducing phosphorylation of Erk1/2 and activation of the CREB/Elk1 transcription factors [120]. This inhibited the transcription of IL4/10, potentially explaining the reduced levels of tumour-associated macrophages in human colon cancer. This is, however, the only anti-tumourigenic role of PRK2 described so far.

PRK2 is highly expressed in triple-negative breast cancer (TNBC) and its depletion in mouse TNBC cells impaired cell proliferation. PRK2 and PRK1 localize to the transition zone of cilia, in agreement with their ability to interact with several cilia proteins and contribute to cilia size and signalling, while PRK2 depletion in polarized epithelial cells impaired planer polarity and anchorage-independent growth [121]. In contrast, CDK10/Cyclin M-dependent phosphorylation of two threonines in the loop between the PRK2 HR1a and HR1b domains may stabilize the interaction of PRK2 with RhoA, leading to actin polymerization and repression of ciliogenesis during cell cycle exit [122].

The role of PRK3 in cancer, particularly metastasis, is being increasingly characterized. Its importance in primary tumour growth and lymph node metastasis was confirmed in an orthotopic mouse prostate cancer model [17], where RhoC was identified its preferred binding partner. PRK3 and RhoC are overexpressed in breast cancer cell lines and PRK3 is critical for their migration in a RhoC-dependent manner. Additionally, PRK3 is vital for melanoma metastasis in mice and has been implicated in the glycosylation of ICAM-1, which is involved in signalling between tumour cells during metastasis [8]. PRK3 also co-localizes with the actin cytoskeleton remodelling adaptor protein p130cas in pro-invasive cell structures and both are overexpressed in breast and prostate cancer. PRK3 phosphorylates p130cas and this is required for PRK3-dependent malignant growth and invasiveness of MEFs expressing constitutively active c-Src [123]. PRK3, therefore, drives metastasis, a hallmark of cancer.

Heart disease

PRK1 activation has been linked to cardiac myocyte survival, increased production and phosphorylation of α crystallin B (a molecular chaperone) and increased cardiac proteasome activity, all of which would protect the heart during ischaemia [124,125]. Hypotonic swelling of cardiac myocytes activates PRK1 and may involve signalling between Src, RhoA, PRK1 and ERK leading to survival [126]. During ischaemia/reperfusion (I/R), PRK1 localizes to the sarcoplasmic reticulum and associates with Ca2+-calmodulin-dependent kinase 2 delta (CamKIIδ) [127], preventing its phosphorylation of phospholamban, which may explain PRK1-dependent cardioprotection in vivo. PRK1 knockouts exhibit slight systolic and diastolic dysfunction and show greater I/R injury [127], consistent with a role in cardioprotection. The involvement of PRK1 in cardioprotection is also suggested by its involvement in the transcriptional up-regulation of ANF, which is secreted by the heart to offset high pressure [91,128]. In humans with chromosome 19p13.12 microdeletion syndrome, PRK1 is lost, leading to several congenital cardiac defects and further implicating a role for PRK1 in normal heart physiology [129].

PRK1 is part of a signalling complex assembled by AKAP-Lbc, which includes PRK1, p38α MAPK, MLTK and MKK3: RhoA/PRK1 signalling is needed to activate p38α in cardiomyocytes leading to compensatory cardiac hypertrophy and preservation of cardiac function following pressure overload [130–132]. PRK1 and PRK2 phosphorylate Myocardin-related transcription factor A (MRTFA), which can then up-regulate cardiac hypertrophic gene expression [133]. However, knockout of PRK1/PRK2 in cardiomyocytes led to greater resistance to cardiac dysfunction, also implicating the PRKs in pathological cardiac hypertrophy and fibrosis regulation [133].

Pathogen infection

Salmonella inject cells with an E3 ubiquitin ligase, SspH1, which selectively removes host proteins. The leucine-rich repeat region of SspH1 binds to PRK1 HR1b (Figure 2D), relieving its own autoinhibition [134]. This leads to SspH1-mediated proteasomal degradation of PRK1, which, given the role of PRK1 in cellular immunity, may be beneficial to the infecting pathogen.

The CagA toxin was responsible for the recruitment of PRK2 from the cytosol to the plasma membrane in Helicobacter pylori-infected gastric adenocarcinoma cells [135]. Surprisingly, while CagA inhibited PRK2 activity, cells infected with CagA showed increased cell motility and invasion, and it was suggested that this may be due to disrupted PRK2-dependent cell adhesion, leading to loss of cellular polarity.

Some pathogens, rather than inhibiting PRK2, depend on it for virulence, e.g. the hepatitis C virus (HCV), where PRK2 phosphorylates its RNA-dependent RNA polymerase allowing viral replication [136–138]. Another example is the interaction of PRK2 with the leucine-rich repeat domains of the YopM proteins of Yersinia pseudotuberculosis and Yersinia enterocolitica; this interaction is necessary for full virulence [139,140] and leads to PRK-dependent pyrin inflammasome inactivation [141,142].

Here, we have summarized the current knowledge about the regulation of the PRKs and the multiple signalling pathways in which they operate.

Several attempts to target the PRK kinase domain have been made with inhibitors that target the active site [37,52,143–148] but, as expected, the well-documented issues with specificity around small molecule inhibitors of kinases apply equally to PRK inhibitors [149]. The complex regulation that we have described above suggests that engineering more specific inhibitors, e.g. to block the PIF pocket, is a potential alternative approach [150]. Extensive understanding of the kinase domain structure, dynamics and allosteric regulation will be necessary for the rational design of such inhibitors.

Attempts have been made to target PRK2 in HCV infection in mice, with siRNA-PRK2 treatment showing potent anti-HCV efficacy, although this needs to be tested in more animal models [151,152]. The low expression of PRK3 in normal tissues and its high expression in cancer suggest that it will have a good therapeutic window. The siRNA/lipoplex complex known as Atu027, designed to inactivate PRK3, showed efficacy in mouse prostate and pancreatic cancer models [68] and very low toxicity in patients [153]. This is currently in Phase II clinical trials for pancreatic cancer [154] and a preliminary Phase Ib/IIa study has demonstrated promising results [155].

As more is discovered about the roles of PRKs in several pathologies, specific inhibitors that could represent useful therapeutics will be eagerly sought.

  • The PRKs are understudied kinases and there is immediate need for further deciphering of their structure and function.

  • More studies are needed to identify novel activation mechanisms, substrates and interacting partners.

  • Inhibitors that are specific for each PRK isoform with few off target effects are required to dissect the role of the PRKs in signalling pathways and as lead therapeutics in several pathologies.

The authors declare that there are no competing interests associated with the manuscript.

All authors were involved in writing the manuscript. GS produced all the figures.

Open access for this article was enabled by the participation of University of Cambridge in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.

G.S. was funded by the A.G. Leventis Foundation. Figures were produced with BioRender.

AMPK

AMP-activated protein kinase

ANF

atrial natriuretic factor

AR

androgen receptor

CDK

cyclin-dependent kinase

GPCR

G protein-coupled receptor

GSK-3β

glycogen synthase kinase 3β

HCV

hepatitis C virus

HDAC

histone deacetylase

HIDS

hyperimmunoglobulinemia D syndrome

HM

hydrophobic motif

HPV

human papilloma virus

HR1

homology region 1

HUVECs

human umbilical endothelial cells

I/R

ischaemia/reperfusion

MAPK

mitogen-activated protein kinase

mTOR

mammalian target of rapamycin

MTRFA

Myocardin-related transcription factor A

PDK1

3-phosphoinositide dependent kinase

PI3K

phosphoinositide-3-kinase

PIP2

phosphatidylinositol 4,5-bisphosphate

PIP3

phosphatidylinositol 3,4,5-triphosphate

PKB

protein kinase B

PKC

protein kinase C

PP2A

protein phosphatase 2A

PRK

protein kinase C-related kinase

PTEN

phosphatase and tensin homologue

ROCK

Rho-associated coiled coil kinase

TM

turn motif

TP

thromboxane receptor

TXA2

thromboxane A2

1
Mukai
,
H.
and
Ono
,
Y.
(
1994
)
A novel protein kinase with leucine zipper-like sequences: its catalytic domain is highly homologous to that of protein kinase C
.
Biochem. Biophys. Res. Commun.
199
,
897
904
2
Palmer
,
R.H.
,
Ridden
,
J.
and
Parker
,
P.J.
(
1994
)
Identification of multiple, novel, protein kinase C-related gene products
.
FEBS Lett.
356
,
5
8
3
Palmer
,
R.H.
,
Ridden
,
J.
and
Parker
,
P.J.
(
1995
)
Cloning and expression patterns of two members of a novel protein-kinase-C-related kinase family
.
Eur. J. Biochem.
227
,
344
351
4
Oishi
,
K.
,
Mukai
,
H.
,
Shibata
,
H.
,
Takahashi
,
M.
and
Ona
,
Y.
(
1999
)
Identification and characterization of PKNβ, a novel isoform of protein kinase PKN: expression and arachidonic acid dependency are different from those of PKNα
.
Biochem. Biophys. Res. Commun.
261
,
808
814
5
Kitagawa
,
M.
,
Mukai
,
H.
,
Shibata
,
H.
and
Ono
,
Y.
(
1995
)
Purification and characterization of a fatty acid-activated protein kinase (PKN) from rat testis
.
Biochem. J.
310
,
657
664
6
Quilliam
,
L.A.
,
Lambert
,
Q.T.
,
Mickelson-Young
,
L.A.
,
Westwick
,
J.K.
,
Sparks
,
A.B.
,
Kay
,
B.K.
et al. (
1996
)
Isolation of a NCK-associated kinase, PRK2, an SH3-binding protein and potential effector of Rho protein signaling
.
J. Biol. Chem.
271
,
28772
6
7
Vincent
,
S.
and
Settleman
,
J.
(
1997
)
The PRK2 kinase is a potential effector target of both Rho and Rac GTPases and regulates actin cytoskeletal organization
.
Mol. Cell. Biol.
17
,
2247
2256
8
Mukai
,
H.
,
Muramatsu
,
A.
,
Mashud
,
R.
,
Kubouchi
,
K.
,
Tsujimoto
,
S.
,
Hongu
,
T.
et al. (
2016
)
PKN3 is the major regulator of angiogenesis and tumor metastasis in mice
.
Sci. Rep.
6
,
18979
9
Zegers
,
M.M.
and
Friedl
,
P.
(
2014
)
Rho GTPases in collective cell migration
.
Small GTPases
5
,
e983869
10
Ridley
,
A.J.
(
2015
)
Rho GTPase signalling in cell migration
.
Curr. Opin. Cell Biol.
36
,
103
112
11
Olson
,
M.F.
,
Ashworth
,
A.
and
Hall
,
A.
(
1995
)
An essential role for Rho, Rac, and Cdc42 GTPases in cell cycle progression through G1
.
Science
269
,
1270
1272
12
Symons
,
M.
and
Rusk
,
N.
(
2003
)
Control of vesicular trafficking by Rho GTPases
.
Curr. Biol.
13
,
R409
R418
13
Mellor
,
H.
,
Flynn
,
P.
,
Nobes
,
C.D.
,
Hall
,
A.
and
Parker
,
P.J.
(
1998
)
PRK1 is targeted to endosomes by the small GTPase, RhoB
.
J. Biol. Chem.
273
,
4811
4814
14
Gross
,
C.
,
Heumann
,
R.
and
Erdmann
,
K.S.
(
2001
)
The protein kinase C-related kinase PRK2 interacts with the protein tyrosine phosphatase PTP-BL via a novel PDZ domain binding motif
.
FEBS Lett.
496
,
101
104
15
Schmidt
,
A.
,
Durgan
,
J.
,
Magalhaes
,
A.
and
Hall
,
A.
(
2007
)
Rho GTPases regulate PRK2/PKN2 to control entry into mitosis and exit from cytokinesis
.
EMBO J.
26
,
1624
1636
16
Metzger
,
E.
,
Müller
,
J.M.
,
Ferrari
,
S.
,
Buettner
,
R.
and
Schüle
,
R.
(
2003
)
A novel inducible transactivation domain in the androgen receptor: implications for PRK in prostate cancer
.
EMBO J.
22
,
270
280
17
Unsal-Kacmaz
,
K.
,
Ragunathan
,
S.
,
Rosfjord
,
E.
,
Dann
,
S.
,
Upeslacis
,
E.
,
Grillo
,
M.
et al. (
2012
)
The interaction of PKN3 with RhoC promotes malignant growth
.
Mol. Oncol.
6
,
284
298
18
Zhu
,
Y.
,
Stolz
,
D.B.
,
Guo
,
F.
,
Ross
,
M.A.
,
Watkins
,
S.C.
,
Tan
,
B.J.
et al. (
2004
)
Signaling via a novel integral plasma membrane pool of a serine/threonine protein kinase PRK1 in mammalian cells
.
FASEB J.
18
,
1722
1724
19
Gampel
,
A.
,
Parker
,
P.J.
and
Mellor
,
H.
(
1999
)
Regulation of epidermal growth factor receptor traffic by the small GTPase RhoB
.
Curr. Biol.
9
,
955
958
20
Wallace
,
S.W.
,
Magalhaes
,
A.
and
Hall
,
A.
(
2011
)
The Rho target PRK2 regulates apical junction formation in human bronchial epithelial cells
.
Mol. Cell Biol.
31
,
81
91
21
Watanabe
,
G.
,
Saito
,
Y.
,
Madaule
,
P.
,
Ishizaki
,
T.
,
Fujisawa
,
K.
,
Morii
,
N.
et al. (
1996
)
Protein kinase N (PKN) and PKN-related protein rhophilin as targets of small GTPase Rho
.
Science
271
,
645
648
22
Amano
,
M.
,
Mukai
,
H.
,
Ono
,
Y.
,
Chihara
,
K.
,
Matsui
,
T.
,
Hamajima
,
Y.
et al. (
1996
)
Identification of a putative target for Rho as the serine-threonine kinase protein kinase N
.
Science
271
,
648
650
23
Flynn
,
P.
,
Mellor
,
H.
,
Palmer
,
R.
,
Panayotou
,
G.
and
Parker
,
P.J.
(
1998
)
Multiple interactions of PRK1 with rhoA: FUNCTIONAL ASSIGNMENT OF THE HR1 REPEAT MOTIF
.
J. Biol. Chem.
273
,
2698
2705
24
Hutchinson
,
C.L.
,
Lowe
,
P.N.
,
McLaughlin
,
S.H.
,
Mott
,
H.R.
and
Owen
,
D.
(
2013
)
Differential binding of RhoA, RhoB, and RhoC to protein kinase C-related kinase (PRK) isoforms PRK1, PRK2, and PRK3: PRKs have the highest affinity for RhoB
.
Biochemistry
52
,
7999
8011
25
Owen
,
D.
,
Lowe
,
P.N.
,
Nietlispach
,
D.
,
Brosnan
,
C.E.
,
Chirgadze
,
D.Y.
,
Parker
,
P.J.
et al. (
2003
)
Molecular dissection of the interaction between the small G proteins Rac1 and rhoA and protein kinase C-related kinase 1 (PRK1)
.
J. Biol. Chem.
278
,
50578
50587
26
Blumenstein
,
L.
and
Ahmadian
,
M.R.
(
2004
)
Models of the cooperative mechanism for Rho effector recognition
.
J. Biol. Chem.
279
,
53419
53426
27
Maesaki
,
R.
,
Ihara
,
K.
,
Shimizu
,
T.
,
Kuroda
,
S.
,
Kaibuchi
,
K.
and
Hakoshima
,
T.
(
1999
)
The structural basis of Rho effector recognition revealed by the crystal structure of human RhoA complexed with the effector domain of PKN/PRK1
.
Mol Cell
4
,
793
803
28
Hutchinson
,
C.L.
,
Lowe
,
P.N.
,
McLaughlin
,
S.H.
,
Mott
,
H.R.
and
Owen
,
D.
(
2011
)
Mutational analysis reveals a single binding interface between RhoA and its effector, PRK1
.
Biochemistry
50
,
2860
2869
29
Modha
,
R.
,
Campbell
,
L.J.
,
Nietlispach
,
D.
,
Buhecha
,
H.R.
,
Owen
,
D.
and
Mott
,
H.R.
(
2008
)
The Rac1 polybasic region is required for interaction with its effector PRK1
.
J. Biol. Chem.
283
,
1492
1500
30
Sophocleous
,
G.
,
Wood
,
G.
,
Owen
,
D.
and
Mott
,
H.R.
(
2020
)
1H, 15n and 13C resonance assignments of the HR1c domain of PRK1, a protein kinase C-related kinase
.
Biomol NMR Assign.
14
,
245
250
31
Corbalán-García
,
S.
and
Gómez-Fernández
,
J.C.
(
2010
)
The C2 domains of classical and novel PKCs as versatile decoders of membrane signals
.
BioFactors
36
,
1
7
32
Palmer
,
R.H.
,
Dekker L
,
V.
,
Woscholski
,
R.
,
Le Good
,
J.A.
,
Gigg
,
R.
and
Parker
,
P.J.
(
1995
)
Activation of PRK1 by phosphatidylinositol 4,5-bisphosphate and phosphatidylinositol 3,4,5-Trisphosphate
.
J. Biol. Chem.
270
,
22412
6
33
Ponting
,
C.P.
and
Parker
,
P.J.
(
1996
)
Extending the C2 domain family: C2s in PKCs δ, ε, η, θ, phospholipases, GAPs, and perforin
.
Protein Sci.
5
,
162
166
34
Nalefski
,
E.A.
and
Falke
,
J.J.
(
1996
)
The C2 domain calcium-binding motif: structural and functional diversity
.
Protein Sci.
5
,
2375
2390
35
Braverman
,
L.E.
and
Quilliam
,
L.A.
(
1999
)
Identification of Grb4/Nckβ, a Src homology 2 and 3 domain-containing adapter protein having similar binding and biological properties to Nck
.
J. Biol. Chem.
274
,
5542
5549
36
Shibata
,
H.
,
Oishi
,
K.
,
Yamagiwa
,
A.
,
Matsumoto
,
M.
,
Mukai
,
H.
and
Ono
,
Y.
(
2001
)
PKNβ interacts with the SH3 domains of Graf and a novel Graf related protein, Graf2, which are GTPase activating proteins for Rho family
.
J. Biochem.
130
,
23
31
37
Chamberlain
,
P.
,
Delker
,
S.
,
Pagarigan
,
B.
,
Mahmoudi
,
A.
,
Jackson
,
P.
,
Abbasian
,
M.
et al. (
2014
)
Crystal structures of PRK1 in complex with the clinical compounds lestaurtinib and tofacitinib reveal ligand induced conformational changes
.
PLoS One
9
,
e103638
38
Nolen
,
B.
,
Taylor
,
S.
and
Ghosh
,
G.
(
2004
)
Regulation of protein kinases: controlling activity through activation segment conformation
.
Mol Cell
15
,
661
675
39
Leroux
,
A.E.
,
Schulze
,
J.O.
and
Biondi
,
R.M.
(
2018
)
AGC kinases, mechanisms of regulation and innovative drug development
.
Semin Cancer Biol.
48
,
1
17
40
Pearce
,
L.R.
,
Komander
,
D.
and
Alessi
,
D.R.
(
2010
)
The nuts and bolts of AGC protein kinases
.
Nat. Rev. Mol. Cell Biol.
11
,
9
22
41
Arencibia
,
J.M.
,
Pastor-Flores
,
D.
,
Bauer
,
A.F.
,
Schulze
,
J.O.
and
Biondi
,
R.M.
(
2013
)
AGC protein kinases: from structural mechanism of regulation to allosteric drug development for the treatment of human diseases
.
Biochim. Biophys. Acta
1834
,
1302
1321
42
Dettori
,
R.
,
Sonzegni
,
S.
,
Meyer
,
L.
,
Lopez-Garcia
,
L.A.
,
Morrice
,
N.A.
,
Zeuzem
,
S.
et al. (
2009
)
Regulation of the interaction between protein kinase C-related protein kinase 2 (PRK2) and its upstream kinase, 3-phosphoinositide-dependent protein kinase 1 (PDK1)
.
J. Biol. Chem.
284
,
30318
30327
43
Yang
,
C.S.
,
Melhuish
,
T.A.
,
Spencer
,
A.
,
Ni
,
L.
,
Hao
,
Y.
,
Jividen
,
K.
et al. (
2017
)
The protein kinase C super-family member PKN is regulated by mTOR and influences differentiation during prostate cancer progression
.
Prostate
77
,
1452
1467
44
Lim
,
W.G.
,
Chen
,
X.
,
Liu
,
J.-P.
,
Tan
,
B.J.
,
Zhou
,
S.
,
Smith
,
A.
et al. (
2008
)
The C-terminus of PRK2/PKNγ is required for optimal activation by RhoA in a GTP-dependent manner
.
Arch. Biochem. Biophys.
479
,
170
178
45
Zeng
,
R.
,
Wang
,
Z.
,
Li
,
X.
,
Chen
,
Y.
,
Yang
,
S.
and
Dong
,
J.
(
2020
)
Cyclin-dependent kinase 1-mediated phosphorylation of protein kinase N1 promotes anchorage-independent growth and migration
.
Cell Signal.
69
,
109546
46
Kitagawa
,
M.
,
Shibata
,
H.
,
Toshimori
,
M.
,
Mukai
,
H.
and
Ono
,
Y.
(
1996
)
The role of the unique motifs in the amino-terminal region of PKN on its enzymatic activity
.
Biochem. Biophys. Res. Commun.
220
,
963
968
47
Collazos
,
A.
,
Michael
,
N.
,
Whelan
,
R.D.H.
,
Kelly
,
G.
,
Mellor
,
H.
,
Pang
,
L.C.H.
et al. (
2011
)
Site recognition and substrate screens for PKN family proteins
.
Biochem. J.
438
,
535
543
48
Shiga
,
K.
,
Takayama
,
K.
,
Futaki
,
S.
,
Hutti
,
J.E.
,
Cantley
,
L.C.
,
Ueki
,
K.
et al. (
2010
)
Development of an intracellularly acting inhibitory peptide selective for PKN
.
Biochem. J.
425
,
445
453
49
Flynn
,
P.
,
Mellor
,
H.
,
Casamassima
,
A.
and
Parker
,
P.J.
(
2000
)
Rho GTPase control of protein kinase C-related protein kinase activation by 3-phosphoinositide-dependent protein kinase
.
J. Biol. Chem.
275
,
11064
11070
50
Lu
,
Y.
and
Settleman
,
J.
(
1999
)
The Drosophila Pkn protein kinase is a Rho/Rac effector target required for dorsal closure during embryogenesis
.
Genes Dev.
13
,
1168
1180
51
Bauer
,
A.F.
,
Sonzogni
,
S.
,
Meyer
,
L.
,
Zeuzem
,
S.
,
Piiper
,
A.
,
Biondi
,
R.M.
et al. (
2012
)
Regulation of protein kinase C-related protein kinase 2 (PRK2) by an intermolecular PRK2-PRK2 interaction mediated by Its N-terminal domain
.
J. Biol. Chem.
287
,
20590
20602
52
Falk
,
M.D.
,
Liu
,
W.
,
Bolanos
,
B.
,
Unsal-Kacmaz
,
K.
,
Klippel
,
A.
,
Grant
,
S.
et al. (
2014
)
Enzyme kinetics and distinct modulation of the protein kinase N family of kinases by lipid activators and small molecule inhibitors
.
Biosci. Rep.
34
,
93
106
53
Mukai
,
H.
,
Kitagawa
,
M.
,
Shibata
,
H.
,
Takanaga
,
H.
,
Mori
,
K.
,
Shimakawa
,
M.
et al. (
1994
)
Activation of PKN, a novel 120-kDa protein kinase with leucine zipper-like sequences, by unsaturated fatty acids and by limited proteolysis
.
Biochem. Biophys. Res. Commun.
204
,
348
356
54
Cryns
,
V.L.
,
Byun
,
Y.
,
Rana
,
A.
,
Mellor
,
H.
,
Lustig
,
K.D.
,
Ghanem
,
L.
et al. (
1997
)
Specific proteolysis of the kinase protein kinase C-related kinase 2 by caspase-3 during apoptosis
.
J. Biol. Chem.
272
,
29449
29453
55
Takahashi
,
M.
,
Mukai
,
H.
,
Toshimori
,
M.
,
Miyamoto
,
M.
and
Ono
,
Y.
(
1998
)
Proteolytic activation of PKN by caspase-3 or related protease during apoptosis
.
Proc. Natl. Acad. Sci. U.S.A.
95
,
11566
11571
56
Ueyama
,
T.
,
Ren
,
Y.
,
Sakai
,
N.
,
Takahashi
,
M.
,
Ono
,
Y.
,
Kondoh
,
T.
et al. (
2001
)
Generation of a constitutively active fragment of PKN in microglia/macrophages after middle cerebral artery occlusion in rats
.
J. Neurochem.
79
,
903
913
57
Yoshinaga
,
C.
,
Mukai
,
H.
,
Toshimori
,
M.
,
Miyamoto
,
M.
and
Ono
,
Y.
(
1999
)
Mutational analysis of the regulatory mechanism of PKN: the regulatory region of PKN contains an arachidonic acid-sensitive autoinhibitory domain
.
J. Biochem.
126
,
475
484
58
Yu
,
W.
,
Liu
,
J.
,
Morrice
,
N.A.
and
Wettenhall
,
R.E.H.
(
1997
)
Isolation and characterization of a structural homologue of human PRK2 from rat liver: distinguishing substrate and lipid activator specificities
.
J. Biol. Chem.
272
,
10030
4
59
Morrice
,
N.A.
,
Gabrielli
,
B.
,
Kemp
,
B.E.
and
Wettenhall
,
R.E.H.
(
1994
)
A cardiolipin-activated protein kinase from rat liver structurally distinct from the protein kinases C
.
J. Biol. Chem.
269
,
20040
6
PMID:
[PubMed]
60
Peng
,
B.
,
Morrice
,
N.A.
,
Groenen
,
L.C.
and
Wettenhall
,
R.E.H.
(
1996
)
Phosphorylation events associated with different states of activation of a hepatic cardiolipin/protease-activated protein kinase
.
J. Biol. Chem.
271
,
32233
32240
61
Mukai
,
H.
and
Ono
,
Y.
(
2006
)
Purification and kinase assay of PKN
.
Methods Enzymol.
406
,
234
250
62
Lin
,
J.L.J.
(
2019
)
Characterization of the novel cardiolipin binding regions identified on the protease and lipid activated PKC-related kinase 1
.
Protein Sci.
28
,
1473
1486
63
Amano
,
M.
,
Hamaguchi
,
T.
,
Shohag
,
M.H.
,
Kozawa
,
K.
,
Kato
,
K.
,
Zhang
,
X.
et al. (
2015
)
Kinase-interacting substrate screening is a novel method to identify kinase substrates
.
J. Cell Biol.
209
,
895
912
64
Mukai
,
H.
,
Toshimori
,
M.
,
Shibata
,
H.
,
Takanaga
,
H.
,
Kitagawa
,
M.
,
Miyahara
,
M.
et al. (
1997
)
Interaction of PKN with α-actinin
.
J. Biol. Chem.
272
,
4740
4746
65
Lim
,
M.A.
,
Yang
,
L.
,
Zheng
,
Y.
,
Wu
,
H.
,
Dong
,
L.Q.
and
Liu
,
F.
(
2004
)
Roles of PDK-1 and PKN in regulating cell migration and cortical actin formation of PTEN-knockout cells
.
Oncogene
23
,
9348
9358
66
Bourguignon
,
L.Y.W.
,
Gilad
,
E.
,
Peyrollier
,
K.
,
Brightman
,
A.
and
Swanson
,
R.A.
(
2007
)
Hyaluronan-CD44 interaction stimulates Rac1 signaling and PKNγ kinase activation leading to cytoskeleton function and cell migration in astrocytes
.
J. Neurochem.
101
,
1002
1017
67
Dong
,
L.Q.
,
Landa
,
L.R.
,
Wick
,
M.J.
,
Zhu
,
L.
,
Mukai
,
H.
,
Ono
,
Y.
et al. (
2000
)
Phosphorylation of protein kinase N by phosphoinositide-dependent protein kinase-1 mediates insulin signals to the actin cytoskeleton
.
Proc. Natl. Acad. Sci. U.S.A.
97
,
5089
5094
68
Aleku
,
M.
,
Schulz
,
P.
,
Keil
,
O.
,
Santel
,
A.
,
Schaeper
,
U.
,
Dieckhoff
,
B.
et al. (
2008
)
Atu027, a liposomal small interfering RNA formulation targeting protein kinase N3, inhibits cancer progression
.
Cancer Res.
68
,
9788
9798
69
Möpert
,
K.
,
Löffler
,
K.
,
Röder
,
N.
,
Kaufmann
,
J.
and
Santel
,
A.
(
2012
)
Depletion of protein kinase N3 (PKN3) impairs actin and adherens junctions dynamics and attenuates endothelial cell activation
.
Eur. J. Cell Biol.
91
,
694
705
70
Hetmanski
,
J.H.R.
,
de Belly
,
H.
,
Busnelli
,
I.
,
Waring
,
T.
,
Nair
,
R.V.
,
Sokleva
,
V.
et al. (
2019
)
Membrane tension orchestrates rear retraction in matrix-directed cell migration
.
Dev. Cell
51
,
460
475
71
Matsuzawa
,
K.
,
Kosako
,
H.
,
Inagaki
,
N.
,
Shibata
,
H.
,
Mukai
,
H.
,
Ono
,
Y.
et al. (
1997
)
Domain-specific phosphorylation of vimentin and glial fibrillary acidic protein by PKN
.
Biochem. Biophys. Res. Commun.
234
,
621
625
72
Manser
,
C.
,
Stevenson
,
A.
,
Banner
,
S.
,
Davies
,
J.
,
Tudor
,
E.L.
,
Ono
,
Y.
et al. (
2008
)
Deregulation of PKN1 activity disrupts neurofilament organisation and axonal transport
.
FEBS Lett.
582
,
2303
2308
73
Kawamata
,
T.
,
Taniguchi
,
T.
,
Mukai
,
H.
,
Kitagawa
,
M.
,
Hashimoto
,
T.
,
Maeda
,
K.
et al. (
1998
)
A protein kinase, PKN, accumulates in Alzheimer neurofibrillary tangles and associated endoplasmic reticulum-derived vesicles and phosphorylates tau protein
.
J. Neurosci.
18
,
7402
7410
74
Taniguchi
,
T.
,
Kawamata
,
T.
,
Mukai
,
H.
,
Hasegawa
,
H.
,
Isagawa
,
T.
,
Yasuda
,
M.
et al. (
2001
)
Phosphorylation of Tau is regulated by PKN
.
J. Biol. Chem.
276
,
10025
10031
75
Uehara
,
S.
,
Udagawa
,
N.
,
Mukai
,
H.
,
Ishihara
,
A.
,
Maeda
,
K.
,
Yamashita
,
T.
et al. (
2017
)
Protein kinase N3 promotes bone resorption by osteoclasts in response to Wnt5a-Ror2 signaling
.
Sci. Signal.
10
,
eaan0023
76
Uehara
,
S.
,
Udagawa
,
N.
and
Kobayashi
,
Y.
(
2019
)
Regulation of osteoclast function via Rho-Pkn3-c-Src pathways
.
J. Oral Biosci.
61
,
135
140
77
Gavard
,
J.
and
Gutkind
,
J.S.
(
2008
)
Protein kinase C-related kinase and ROCK are required for thrombin-induced endothelial cell permeability downstream from Gα12/13 and Gα11/q
.
J. Biol. Chem.
283
,
29888
29896
78
Fischer
,
A.
,
Stuckas
,
H.
,
Gluth
,
M.
,
Russell
,
T.D.
,
Rudolph
,
M.C.
,
Beeman
,
N.E.
et al. (
2007
)
Impaired tight junction sealing and precocious involution in mammary glands of PKN1 transgenic mice
.
J. Cell Sci.
120
,
2272
2283
79
Mehruba
,
M.
,
Siddique
,
S.M.
and
Mukai
,
H.
(
2020
)
PKN1 controls the aggregation, spheroid formation, and viability of mouse embryonic fibroblasts in suspension culture
.
Biochem. Biophys. Res. Commun.
523
,
398
404
80
Calautti
,
E.
,
Grossi
,
M.
,
Mammucari
,
C.
,
Aoyama
,
Y.
,
Pirro
,
M.
,
Ono
,
Y.
et al. (
2002
)
Fyn tyrosine kinase is a downstream mediator of Rho/PRK2 function in keratinocyte cell-cell adhesion
.
J. Cell Biol.
156
,
137
148
81
Lee
,
S.J.
,
Hwang
,
J.
,
Jeong
,
H.J.
,
Yoo
,
M.
,
Go
,
G.Y.
,
Lee
,
J.R.
et al. (
2016
)
PKN2 and Cdo interact to activate AKT and promote myoblast differentiation
.
Cell Death Dis.
7
,
e2431
82
Yasui
,
T.
,
Sakakibara-Yada
,
K.
,
Nishimura
,
T.
,
Morita
,
K.
,
Tada
,
S.
,
Mosialos
,
G.
et al. (
2012
)
Protein kinase N1, a cell inhibitor of Akt kinase, has a central role in quality control of germinal center formation
.
Proc. Natl. Acad. Sci. U.S.A.
109
,
21022
7
83
zur Nedden
,
S.
,
Eith
,
R.
,
Schwarzer
,
C.
,
Zanetti
,
L.
,
Seitter
,
H.
,
Fresser
,
F.
et al. (
2018
)
Protein kinase N1 critically regulates cerebellar development and long-term function
.
J. Clin. Invest.
128
,
2076
2088
84
Wick
,
M.J.
,
Dong
,
L.Q.
,
Riojas
,
R.A.
,
Ramos
,
F.J.
and
Liu
,
F.
(
2000
)
Mechanism of phosphorylation of protein kinase B/Akt by a constitutively active 3-phosphoinositide-dependent protein kinase-1
.
J. Biol. Chem.
275
,
40400
6
85
Metzger
,
E.
,
Yin
,
N.
,
Wissmann
,
M.
,
Kunowska
,
N.
,
Fischer
,
K.
,
Friedrichs
,
N.
et al. (
2008
)
Phosphorylation of histone H3 at threonine 11 establishes a novel chromatin mark for transcriptional regulation
.
Nat. Cell Biol.
10
,
53
60
86
O'Sullivan
,
A.G.
,
Mulvaney
,
E.P.
,
Hyland
,
P.B.
and
Kinsella
,
B.T.
(
2015
)
Protein kinase C-related kinase 1 and 2 play an essential role in thromboxane-mediated neoplastic responses in prostate cancer
.
Oncotarget
6
,
26437
26456
87
O'Sullivan
,
A.G.
,
Mulvaney
,
E.P.
and
Kinsella
,
B.T.
(
2017
)
Regulation of protein kinase C-related kinase (PRK) signalling by the TPα and TPβ isoforms of the human thromboxane A2 receptor: Implications for thromboxane- and androgen- dependent neoplastic and epigenetic responses in prostate cancer
.
Biochim. Biophys. Acta – Mol. Basis Dis.
1863
,
838
856
88
Harrison
,
B.C.
,
Huynh
,
K.
,
Lundgaard
,
G.L.
,
Helmke
,
S.M.
,
Perryman
,
M.B.
and
McKinsey
,
T.A.
(
2010
)
Protein kinase C-related kinase targets nuclear localization signals in a subset of class IIa histone deacetylases
.
FEBS Lett.
584
,
1103
1110
89
Mukai
,
H.
,
Miyahara
,
M.
,
Sunakawa
,
H.
,
Shibata
,
H.
,
Toshimori
,
M.
,
Kitagawa
,
M.
et al. (
1996
)
Translocation of PKN from the cytosol to the nucleus induced by stresses
.
Proc. Natl. Acad. Sci. U.S.A.
93
,
10195
9
90
Shibata
,
H.
,
Oda
,
H.
,
Mukai
,
H.
,
Oishi
,
K.
,
Misaki
,
K.
,
Ohkubo
,
H.
et al. (
1999
)
Interaction of PKN with a neuron-specific basic helix–loop–helix transcription factor, NDRF/NeuroD2
.
Mol. Brain Res.
74
,
126
134
91
Morissette
,
M.R.
,
Sah
,
V.P.
,
Glembotski
,
C.C.
and
Brown
,
J.H.
(
2000
)
The Rho effector, PKN, regulates ANF gene transcription in cardiomyocytes through a serum response element
.
Am. J. Physiol. - Hear Circ. Physiol.
278
,
H1769
H1774
92
Cottone
,
G.
,
Baldi
,
A.
,
Palescandolo
,
E.
,
Manente
,
L.
,
Penta
,
R.
,
Paggi
,
M.G.
et al. (
2006
)
Pkn is a novel partner of cyclin T2a in muscle differentiation
.
J. Cell Physiol.
207
,
232
237
93
Jilg
,
C.A.
,
Ketscher
,
A.
,
Metzger
,
E.
,
Hummel
,
B.
,
Willmann
,
D.
,
Rüsseler
,
V.
et al. (
2014
)
PRK1/PKN1 controls migration and metastasis of androgen-independent prostate cancer cells
.
Oncotarget
5
,
12646
12664
94
Singh
,
N.K.
,
Kundumani-Sridharan
,
V.
,
Kumar
,
S.
,
Verma
,
S.K.
,
Kotla
,
S.
,
Mukai
,
H.
et al. (
2012
)
Protein kinase N1 is a novel substrate of NFATc1-mediated cyclin D1-CDK6 activity and modulates vascular smooth muscle cell division and migration leading to inward blood vessel wall remodeling
.
J. Biol. Chem.
287
,
36291
36304
95
Singh
,
N.K.
,
Janjanam
,
J.
and
Rao
,
G.N.
(
2017
)
P115 RhoGEF activates the Rac1 GTPase signaling cascade in MCP1 chemokine-induced vascular smooth muscle cell migration and proliferation
.
J. Biol. Chem.
292
,
14080
14091
96
Wang
,
F.
,
Zhan
,
R.
,
Chen
,
L.
,
Dai
,
X.
,
Wang
,
W.
,
Guo
,
R.
et al. (
2017
)
Rhoa promotes epidermal stem cell proliferation via PKN1-cyclin D1 signaling
.
PLoS One
12
,
e0172613
97
Jamilloux
,
Y.
,
Magnotti
,
F.
,
Belot
,
A.
and
Henry
,
T.
(
2018
)
The pyrin inflammasome: from sensing RhoA GTPases-inhibiting toxins to triggering autoinflammatory syndromes
.
Pathog. Dis.
76
,
1
9
98
Park
,
Y.H.
,
Wood
,
G.
,
Kastner
,
D.L.
and
Chae
,
J.J.
(
2016
)
Pyrin inflammasome activation and RhoA signaling in the autoinflammatory diseases FMF and HIDS
.
Nat. Immunol.
17
,
914
921
99
Mashud
,
R.
,
Nomachi
,
A.
,
Hayakawa
,
A.
,
Kubouchi
,
K.
,
Danno
,
S.
,
Hirata
,
T.
et al. (
2017
)
Impaired lymphocyte trafficking in mice deficient in the kinase activity of PKN1
.
Sci. Rep.
7
,
7663
100
Siddique
,
S.M.
,
Kubouchi
,
K.
,
Shinmichi
,
Y.
,
Sawada
,
N.
,
Sugiura
,
R.
,
Itoh
,
Y.
et al. (
2019
)
PKN1 kinase-negative knock-in mice develop splenomegaly and leukopenia at advanced age without obvious autoimmune-like phenotypes
.
Sci. Rep.
9
,
13977
101
Standaert
,
M.
,
Bandyopadhyay
,
G.
,
Galloway
,
L.
,
Ono
,
Y.
,
Mukai
,
H.
and
Farese
,
R.
(
1998
)
Comparative effects of GTPγS and insulin on the activation of Rho, phosphatidylinositol 3-kinase, and protein kinase N in rat adipocytes
.
J. Biol. Chem.
273
,
7470
7477
102
Banno
,
Y.
(
2002
)
Regulation and possible role of mammalian phospholipase D in cellular functions
.
J. Biochem.
131
,
301
306
103
Oishi
,
K.
,
Takahashi
,
M.
,
Mukai
,
H.
,
Banno
,
Y.
,
Nakashima
,
S.
,
Kanaho
,
Y.
et al. (
2001
)
PKN regulates phospholipase D1 through direct interaction
.
J. Biol. Chem.
276
,
18096
18101
104
Isagawa
,
T.
,
Mukai
,
H.
,
Oishi
,
K.
,
Taniguchi
,
T.
,
Hasegawa
,
H.
,
Kawamata
,
T.
et al. (
2000
)
Dual effects of PKNα and protein kinase C on phosphorylation of tau protein by glycogen synthase kinase-3β
.
Biochem. Biophys. Res. Commun.
273
,
209
212
105
Ruby
,
M.A.
,
Riedl
,
I.
,
Massart
,
J.
,
Åhlin
,
M.
and
Zierath
,
J.R.
(
2017
)
Protein kinase N2 regulates AMP kinase signaling and insulin responsiveness of glucose metabolism in skeletal muscle
.
Am. J. Physiol. – Endocrinol. Metab.
313
,
E483
E491
106
Wallroth
,
A.
,
Koch
,
P.A.
,
Marat
,
A.L.
,
Krause
,
E.
and
Haucke
,
V.
(
2019
)
Protein kinase N controls a lysosomal lipid switch to facilitate nutrient signalling via mTORC1
.
Nat. Cell Biol.
21
,
1093
1101
107
Quétier
,
I.
,
Marshall
,
J.J.T.
,
Spencer-Dene
,
B.
,
Lachmann
,
S.
,
Casamassima
,
A.
,
Franco
,
C.
et al. (
2016
)
Knockout of the PKN family of Rho effector kinases reveals a Non-redundant role for PKN2 in developmental mesoderm expansion
.
Cell Rep.
14
,
440
448
108
Danno
,
S.
,
Kubouchi
,
K.
,
Mehruba
,
M.
,
Abe
,
M.
,
Natsume
,
R.
,
Sakimura
,
K.
et al. (
2017
)
PKN2 is essential for mouse embryonic development and proliferation of mouse fibroblasts
.
Genes Cells
22
,
220
236
109
Tsankova
,
A.
,
Pham
,
T.T.
,
Garcia
,
D.S.
,
Otte
,
F.
and
Cabernard
,
C.
(
2017
)
Cell polarity regulates biased myosin activity and dynamics during asymmetric cell division via Drosophila Rho kinase and protein kinase N
.
Dev. Cell.
42
,
143
155
110
Sass
,
G.L.
and
Ostrow
,
B.D.
(
2014
)
Disruption of the protein kinase N gene of Drosophila melanogaster results in the recessive delorean allele (pkndln) With a negative impact on wing morphogenesis
.
G3 (Bethesda)
4
,
643
656
111
Ferreira
,
T.
,
Prudêncio
,
P.
and
Martinho
,
R.G.
(
2014
)
Drosophila protein kinase N (Pkn) is a negative regulator of actin-myosin activity during oogenesis
.
Dev. Biol.
394
,
277
291
112
Leenders
,
F.
,
Möpert
,
K.
,
Schmiedeknecht
,
A.
,
Santel
,
A.
,
Czauderna
,
F.
,
Aleku
,
M.
et al. (
2004
)
PKN3 is required for malignant prostate cell growth downstream of activated PI 3-kinase
.
EMBO J.
23
,
3303
3313
113
Galgano
,
M.T.
,
Conaway
,
M.
,
Spencer
,
A.M.
,
Paschal
,
B.M.
and
Frierson
,
H.F.
(
2009
)
PRK1 distribution in normal tissues and carcinomas: overexpression and activation in ovarian serous carcinoma
.
Hum. Pathol.
40
,
1434
1440
114
Chen
,
L.
,
Shern
,
J.F.
,
Wei
,
J.S.
,
Yohe
,
M.E.
,
Song
,
Y.K.
,
Hurd
,
L.
et al. (
2015
)
Clonality and evolutionary history of rhabdomyosarcoma
.
PLoS Genet.
11
,
e1005075
115
Kastnerova
,
L.
,
Luzar
,
B.
,
Goto
,
K.
,
Grishakov
,
V.
,
Gatalica
,
Z.
,
Kamarachev
,
J.
et al. (
2019
)
Secretory carcinoma of the skin: report of 6 cases, including a case with a novel NFIX-PKN1 translocation
.
Am. J. Surg. Pathol.
43
,
1092
1098
116
Stransky
,
N.
,
Cerami
,
E.
,
Schalm
,
S.
,
Kim
,
J.L.
and
Lengauer
,
C.
(
2014
)
The landscape of kinase fusions in cancer
.
Nat. Commun.
5
,
4846
117
Donati
,
M.
,
Kastnerova
,
L.
,
Cempírková
,
D.
,
Vaněček
,
T.
,
Michal
,
M.
and
Kazakov D
,
V.
(
2020
)
Vulvar pigmented epithelioid melanocytoma with a novel HTT-PKN1 fusion: a case report
.
Am. J. Dermatopathol.
42
,
544
546
118
Lachmann
,
S.
,
Jevons
,
A.
,
De Rycker
,
M.
,
Casamassima
,
A.
,
Radtke
,
S.
,
Collazos
,
A.
et al. (
2011
)
Regulatory domain selectivity in the cell-type specific PKN-dependence of cell migration
.
PLoS One
6
,
e21732
119
Rajagopalan
,
P.
,
Nanjappa
,
V.
,
Patel
,
K.
,
Jain
,
A.P.
,
Mangalaparthi
,
K.K.
,
Patil
,
A.H.
et al. (
2018
)
Role of protein kinase N2 (PKN2) in cigarette smoke-mediated oncogenic transformation of oral cells
.
J. Cell Commun. Signal.
12
,
709
721
120
Cheng
,
Y.
,
Zhu
,
Y.
,
Xu
,
J.
,
Yang
,
M.
,
Chen
,
P.
,
Xu
,
W.
et al. (
2018
)
PKN2 in colon cancer cells inhibits M2 phenotype polarization of tumor-associated macrophages via regulating DUSP6-Erk1/2 pathway
.
Mol. Cancer
17
,
1
16
121
Patel
,
H.
,
Li
,
J.
,
Herrero
,
A.
,
Kroboth
,
J.
,
Byron
,
A.
,
Von Kriegsheim
,
A.
et al. (
2020
)
Novel roles of PRK1 and PRK2 in cilia and cancer biology
.
Sci. Rep.
10
,
3902
122
Guen
,
V.J.
,
Gamble
,
C.
,
Perez
,
D.E.
,
Bourassa
,
S.
,
Zappel
,
H.
,
Gärtner
,
J.
et al. (
2016
)
STAR syndrome-associated CDK10/Cyclin M regulates actin network architecture and ciliogenesis
.
Cell Cycle
15
,
678
688
123
Gemperle
,
J.
,
Dibus
,
M.
,
Koudelková
,
L.
,
Rosel
,
D.
and
Brábek
,
J.
(
2019
)
The interaction of p130Cas with PKN3 promotes malignant growth
.
Mol. Oncol.
13
,
264
289
124
Kitagawa
,
M.
,
Mukai
,
H.
,
Takahashi
,
M.
and
Ono
,
Y.
(
1998
)
The role of PKN in the regulation of αB-crystallin expression via heat shock transcription factor 1
.
Biochem. Biophys. Res. Commun.
252
,
561
565
125
Takagi
,
H.
,
Hsu
,
C.P.
,
Kajimoto
,
K.
,
Shao
,
D.
,
Yang
,
Y.
,
Maejima
,
Y.
et al. (
2010
)
Activation of PKN mediates survival of cardiac myocytes in the heart during ischemia/reperfusion
.
Circ. Res.
107
,
642
649
126
Kajimoto
,
K.
,
Shao
,
D.
,
Takagi
,
H.
,
Maceri
,
G.
,
Zablocki
,
D.
,
Mukai
,
H.
et al. (
2011
)
Hypotonic swelling-induced activation of PKN1 mediates cell survival in cardiac myocytes
.
Am. J. Physiol. Circ. Physiol.
300
,
H191
H200
127
Francois
,
A.A.
,
Obasanjo-Blackshire
,
K.
,
Clark
,
J.E.
,
Boguslavskyi
,
A.
,
Holt
,
M.R.
,
Parker
,
P.J.
et al. (
2018
)
Loss of protein kinase novel 1 (PKN1) is associated with mild systolic and diastolic contractile dysfunction, increased phospholamban Thr17 phosphorylation, and exacerbated ischaemia-reperfusion injury
.
Cardiovasc. Res.
114
,
138
157
128
Fu
,
S.
,
Ping
,
P.
,
Wang
,
F.
and
Luo
,
L.
(
2018
)
Synthesis, secretion, function, metabolism and application of natriuretic peptides in heart failure
.
J. Biol. Eng.
12
,
1
21
129
Jensen
,
D.R.
,
Martin
,
D.M.
,
Gebarski
,
S.
,
Sahoo
,
T.
,
Brundage
,
E.K.
,
Chinault
,
A.C.
et al. (
2009
)
A novel chromosome 19p13.12 deletion in a child with multiple congenital anomalies
.
Am. J. Med. Genet. Part A
149A
,
396
402
130
Deaton
,
R.A.
,
Su
,
C.
,
Valencia
,
T.G.
and
Grant
,
S.R.
(
2005
)
Transforming growth factor-β1-induced expression of smooth muscle marker genes involves activation of PKN and p38 MAPK
.
J. Biol. Chem.
280
,
31172
31181
131
Cariolato
,
L.
,
Cavin
,
S.
and
Diviani
,
D.
(
2011
)
A-kinase anchoring protein (AKAP)-Lbc anchors a PKN-based signaling complex involved in α1-adrenergic receptor-induced p38 activation
.
J. Biol. Chem.
286
,
7925
7937
132
López I
,
P.
,
Cariolato
,
L.
,
Maric
,
D.
,
Gillet
,
L.
,
Abriel
,
H.
and
Diviani
,
D.
(
2013
)
A-kinase anchoring protein Lbc coordinates a p38 activating signaling complex controlling compensatory cardiac hypertrophy
.
Mol. Cell. Biol.
33
,
2903
2917
133
Sakaguchi
,
T.
,
Takefuji
,
M.
,
Wettschureck
,
N.
,
Hamaguchi
,
T.
,
Amano
,
M.
,
Kato
,
K.
et al. (
2019
)
Protein kinase N promotes stress-induced cardiac dysfunction through phosphorylation of myocardin-related transcription factor A and disruption of Its interaction with actin
.
Circulation
140
,
1737
1752
134
Keszei
,
A.F.A.
,
Tang
,
X.
,
McCormick
,
C.
,
Zeqiraj
,
E.
,
Rohde
,
J.R.
,
Tyers
,
M.
et al. (
2014
)
Structure of an SspH1-PKN1 complex reveals the basis for host substrate recognition and mechanism of activation for a bacterial E3 ubiquitin ligase
.
Mol. Cell. Biol.
34
,
362
373
135
Mishra
,
J.P.
,
Cohen
,
D.
,
Zamperone
,
A.
,
Nesic
,
D.
,
Muesch
,
A.
and
Stein
,
M.
(
2015
)
Caga of Helicobacter pylori interacts with and inhibits the serine-threonine kinase PRK2
.
Cell Microbiol.
17
,
1670
1682
136
Kim
,
S.J.
,
Kim
,
J.H.
,
Kim
,
Y.G.
,
Lim
,
H.S.
and
Oh
,
J.W.
(
2004
)
Protein kinase C-related kinase 2 regulates hepatitis C virus RNA polymerase function by phosphorylation
.
J. Biol. Chem.
279
,
50031
50041
137
Kim
,
M.G.
,
Moon
,
J.S.
,
Kim
,
E.J.
,
Lee
,
S.H.
and
Oh
,
J.W.
(
2012
)
Destabilization of PDK1 by Hsp90 inactivation suppresses hepatitis C virus replication through inhibition of PRK2-mediated viral RNA polymerase phosphorylation
.
Biochem. Biophys. Res. Commun.
421
,
112
118
138
Han
,
S.H.
,
Kim
,
S.J.
,
Kim
,
E.J.
,
Kim
,
T.E.
,
Moon
,
J.S.
,
Kim
,
G.W.
et al. (
2014
)
Phosphorylation of hepatitis C virus RNA polymerases Ser29 and Ser42 by protein kinase C-Related kinase 2 regulates viral RNA replication
.
J. Virol.
88
,
11240
11252
139
McPhee
,
J.B.
,
Mena
,
P.
and
Bliska
,
J.B.
(
2010
)
Delineation of regions of the Yersinia YopM protein required for interaction with the RSK1 and PRK2 host kinases and their requirement for interleukin-10 production and virulence
.
Infect. Immun.
78
,
3529
3539
140
Höfling
,
S.
,
Scharnert
,
J.
,
Cromme
,
C.
,
Bertrand
,
J.
,
Pap
,
T.
,
Schmidt
,
M.A.
et al. (
2014
)
Manipulation of pro-inflammatory cytokine production by the bacterial cell-penetrating effector protein YopM is independent of its interaction with host cell kinases RSK1 and PRK2
.
Virulence
5
,
761
771
141
Chung
,
L.K.
,
Park
,
Y.H.
,
Zheng
,
Y.
,
Brodsky
,
I.E.
Hearing
,
P.
,
Kastner
,
D.L.
et al. (
2016
)
The Yersinia virulence factor YopM hijacks host kinases to inhibit type III effector-triggered activation of the pyrin inflammasome
.
Cell Host Microbe.
20
,
296
306
142
Ratner
,
D.
,
Orning
,
M.P.A.
,
Proulx
,
M.K.
,
Wang
,
D.
,
Gavrilin
,
M.A.
,
Wewers
,
M.D.
et al. (
2016
)
The Yersinia pestis effector yopM inhibits pyrin inflammasome activation
.
PLoS Pathog.
12
,
e1006035
143
Köhler
,
J.
,
Erlenkamp
,
G.
,
Eberlin
,
A.
,
Rumpf
,
T.
,
Slynko
,
I.
,
Metzger
,
E.
et al. (
2012
)
Lestaurtinib inhibits histone phosphorylation and androgen-dependent gene expression in prostate cancer cells
.
PLoS One
7
,
e34973
144
Slynko
,
I.
,
Schmidtkunz
,
K.
,
Rumpf
,
T.
,
Klaeger
,
S.
,
Heinzlmeir
,
S.
,
Najar
,
A.
et al. (
2016
)
Identification of highly potent protein kinase C-Related kinase 1 inhibitors by virtual screening, binding free energy rescoring, and in vitro testing
.
ChemMedChem
11
,
2084
2094
145
Kim
,
S.J.
,
Kim
,
J.H.
,
Sun
,
J.M.
,
Kim
,
M.G.
and
Oh
,
J.W.
(
2009
)
Suppression of hepatitis C virus replication by protein kinase C-related kinase 2 inhibitors that block phosphorylation of viral RNA polymerase
.
J. Viral Hepat.
16
,
697
704
146
Lee
,
S.H.
,
Moon
,
J.S.
,
Pak
,
B.Y.
,
Kim
,
G.W.
,
Lee
,
W.
,
Cho
,
H.
et al. (
2018
)
HA1077 displays synergistic activity with daclatasvir against hepatitis C virus and suppresses the emergence of NS5A resistance-associated substitutions in mice
.
Sci. Rep.
8
,
12469
147
Browne
,
C.M.
,
Jiang
,
B.
,
Ficarro
,
S.B.
,
Doctor
,
Z.M.
,
Johnson
,
J.L.
,
Card
,
J.D.
et al. (
2019
)
A chemoproteomic strategy for direct and proteome-wide covalent inhibitor target-site identification
.
J. Am. Chem. Soc.
141
,
191
203
148
Scott
,
F.
,
Fala
,
A.M.
,
Pennicott
,
L.E.
,
Reuillon
,
T.D.
,
Massirer
,
K.B.
,
Elkins
,
J.M.
et al. (
2020
)
Development of 2-(4-pyridyl)-benzimidazoles as PKN2 chemical tools to probe cancer
.
Bioorganic. Med. Chem. Lett.
30
,
127040
149
Karaman
,
M.W.
,
Herrgard
,
S.
,
Treiber
,
D.K.
,
Gallant
,
P.
,
Atteridge
,
C.E.
,
Campbell
,
B.T.
et al. (
2008
)
A quantitative analysis of kinase inhibitor selectivity
.
Nat. Biotechnol.
26
,
127
132
150
Arencibia
,
J.M.
,
Fröhner
,
W.
,
Krupa
,
M.
,
Pastor-Flores
,
D.
,
Merker
,
P.
,
Oellerich
,
T.
et al. (
2017
)
An allosteric inhibitor scaffold targeting the PIF-pocket of atypical protein kinase C isoforms
.
ACS Chem. Biol.
12
,
564
573
151
Moon
,
J.S.
,
Lee
,
S.H.
,
Han
,
S.H.
,
Kim
,
E.J.
,
Cho
,
H.
,
Lee
,
W.
et al. (
2016
)
Inhibition of hepatitis C virus in mouse models by lipidoid nanoparticle-mediated systemic delivery of siRNA against PRK2
.
Nanomedicine
12
,
1489
1498
152
Park
,
H.J.
,
Jeon
,
E.J.
,
Lee
,
J.S.
,
Hong
,
S.H.
,
Cho
,
A.N.
,
Lee
,
J.
et al. (
2016
)
Galactosylated lipidoid nanoparticles for delivery of small interfering RNA to inhibit hepatitis C viral replication In vivo
.
Adv. Healthc. Mater.
5
,
2931
2941
153
Strumberg
,
D.
,
Schultheis
,
B.
,
Traugott
,
U.
,
Vank
,
C.
,
Santel
,
A.
,
Keil
,
O.
et al. (
2012
)
Phase I clinical development of Atu027, a siRNA formulation targeting PKN3 in patients with advanced solid tumors
.
Int. J. Clin. Pharmacol. Ther.
50
,
76
78
154
Takakura
,
K.
,
Kawamura
,
A.
,
Torisu
,
Y.
,
Koido
,
S.
,
Yahagi
,
N.
and
Saruta
,
M.
(
2019
)
The clinical potential of oligonucleotide therapeutics against pancreatic cancer
.
Int. J. Mol. Sci.
20
,
3331
155
Schultheis
,
B.
,
Strumberg
,
D.
,
Kuhlmann
,
J.
,
Wolf
,
M.
,
Link
,
K.
,
Seufferlein
,
T.
et al. (
2020
)
Safety, efficacy and pharcacokinetics of targeted therapy with the liposomal RNA interference therapeutic Atu027 combined with gemcitabine in patients with pancreatic adenocarcinoma
.
A randomized phase Ib/IIa study. Cancers (Basel)
12
,
3130
156
Misaki
,
K.
,
Mukai
,
H.
,
Yoshinaga
,
C.
,
Oishi
,
K.
,
Isagawa
,
T.
,
Takahashi
,
M.
et al. (
2001
)
PKN delays mitotic timing by inhibition of Cdc25C: Possible involvement of PKN in the regulation of cell division
.
Proc. Natl. Acad. Sci. U.S.A.
98
,
125
129
157
Yuan
,
Q.
,
Ren
,
C.
,
Xu
,
W.
,
Petri
,
B.
,
Zhang
,
J.
,
Zhang
,
Y.
et al. (
2017
)
PKN1 directs polarized RAB21 vesicle trafficking via RPH3A and Is important for neutrophil adhesion and ischemia-reperfusion injury
.
Cell Rep.
19
,
2586
2597
158
Gao
,
Q.
,
Kumar
,
A.
,
Srinivasan
,
S.
,
Singh
,
L.
,
Mukai
,
H.
,
Ono
,
Y.
et al. (
2000
)
PKN binds and phosphorylates human papillomavirus E6 oncoprotein
.
J. Biol. Chem.
275
,
14824
14830
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY). Open access for this article was enabled by the participation of University of Cambridge in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.