Next Article in Journal
L-carnitine and Ginkgo biloba Supplementation In Vivo Ameliorates HCD-Induced Steatohepatitis and Dyslipidemia by Regulating Hepatic Metabolism
Previous Article in Journal
Tumor Microenvironment Modulates Invadopodia Activity of Non-Selected and Acid-Selected Pancreatic Cancer Cells and Its Sensitivity to Gemcitabine and C18-Gemcitabine
Previous Article in Special Issue
The Post-Translational Role of UFMylation in Physiology and Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pin1-Catalyzed Conformation Changes Regulate Protein Ubiquitination and Degradation

by
Jessica Jeong
1,2,†,
Muhammad Usman
1,2,†,
Yitong Li
1,2,†,
Xiao Zhen Zhou
1,3,4,* and
Kun Ping Lu
1,2,*
1
Departments of Biochemistry and Oncology, Schulich School of Medicine & Dentistry, Western University, London, ON N6A 5C1, Canada
2
Robarts Research Institute, Western University, London, ON N6A 5B7, Canada
3
Department of Pathology and Laboratory Medicine, and Oncology, Schulich School of Medicine & Dentistry, Western University, London, ON N6A 5C1, Canada
4
Lawson Health Research Institute, Western University, London, ON N6C 2R5, Canada
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2024, 13(9), 731; https://doi.org/10.3390/cells13090731
Submission received: 22 February 2024 / Revised: 12 April 2024 / Accepted: 14 April 2024 / Published: 23 April 2024
(This article belongs to the Special Issue Advances in Ubiquitination and Deubiquitination Research)

Abstract

:
The unique prolyl isomerase Pin1 binds to and catalyzes cis–trans conformational changes of specific Ser/Thr-Pro motifs after phosphorylation, thereby playing a pivotal role in regulating the structure and function of its protein substrates. In particular, Pin1 activity regulates the affinity of a substrate for E3 ubiquitin ligases, thereby modulating the turnover of a subset of proteins and coordinating their activities after phosphorylation in both physiological and disease states. In this review, we highlight recent advancements in Pin1-regulated ubiquitination in the context of cancer and neurodegenerative disease. Specifically, Pin1 promotes cancer progression by increasing the stabilities of numerous oncoproteins and decreasing the stabilities of many tumor suppressors. Meanwhile, Pin1 plays a critical role in different neurodegenerative disorders via the regulation of protein turnover. Finally, we propose a novel therapeutic approach wherein the ubiquitin–proteasome system can be leveraged for therapy by targeting pathogenic intracellular targets for TRIM21-dependent degradation using stereospecific antibodies.

1. Introduction

Protein phosphorylation is one of the most important and universal regulatory post-translational modifications (PTMs) in the cell, occurring as a response to intracellular and extracellular stimuli to initiate signaling cascades or alter protein–protein interactions. In eukaryotes, this reversible mechanism consists of the covalent addition of a phosphate group at the serine, threonine, or tyrosine side chain of a protein and is mediated by kinase and phosphatase enzymes. Phosphorylation plays a crucial role in regulating a diverse range of cellular activities, one of which is ubiquitination. Specifically, phosphorylation serves to modulate a substrate’s receptivity to ubiquitination and influence the activity of substrate-ubiquitinating enzymes, particularly in relation to the ubiquitin–proteasome system (UPS) [1].
The UPS is the major enzymatic pathway in eukaryotes for intracellular protein degradation, consisting of the covalent tagging of proteins with ubiquitin (Ub) and the subsequent degradation by the 26S proteasome. Ubiquitination is orchestrated by at least three enzymes: E1: a Ub-activating enzyme; E2: a Ub-conjugating enzyme; and E3: a Ub ligase. Initially, E1 will use adenosine triphosphate (ATP) to activate Ub, creating a reactive Ub thioester at the protein C-terminus that is conjugated to E1. This activated Ub will undergo trans thiolation onto E2 before it is finally transferred onto the target substrate by E2 and E3 enzymatic activity [2]. Much of the specificity of this post-translational modification pathway is owed to this final step, wherein the E3 Ub ligase acts as a matchmaker to specifically recognize and bind the substrate to be ubiquitinated, as evidenced by the hundreds of E3 analogs required to fulfill conjugation specificity, compared to the single mammalian E1 and dozens of E2s [3,4]. E3 regulation of the UPS is particularly significant because the 26S proteasome cleaves proteins exclusively based on the Ub marker and not based on any specific amino acid sequence, like other proteases. This allows it to degrade a wide range of substrates with a high degree of specificity. Typically, substrates undergo post-translational modification with the addition of a Ub monomer (monoubiquitination); however, Ub itself can be a target for further ubiquitination, allowing the processive formation of a short Ub chain (oligoubiquitination) or a longer Ub chain (polyubiquitination) [5]. Alternatively, another class of ubiquitination factors, E4 enzymes, may also catalyze the elongation of the Ub chain [6].
One pivotal regulator of phosphorylation-mediated ubiquitination is Pin1 (protein interacting with NIMA (never in mitosis A)-1), a ubiquitously expressed and highly conserved eukaryotic peptidyl-prolyl isomerase (PPIase). Pin1 uniquely binds and catalyzes cis–trans conformational changes at specific phosphorylated Ser/Thr-Pro motifs [7,8,9], thereby modulating the activities, functions, and stabilities of a broad range of substrates after phosphorylation. Notable Pin1 substrates include several master regulators, such as p65/RelA nuclear factor-kappaB (NF-κB) [10], cyclin-dependent kinases (CDKs) [11], and the tumor suppressor protein p53 [12]. Accordingly, as Pin1 is a critical and dynamic signaling regulator for many substrates in essential functions, such as the cell cycle, immunity, and metabolism [13,14,15], it is widely distributed at both the cellular and tissue levels, depending on its substrates [16]. Thus, Pin1 can be found in both the nucleus and cytoplasm of many cell types depending on the presence of its substates, including stromal, parenchymal, and stem cells [17,18,19,20].
Pin1 is especially relevant in our discussion of phosphorylation-mediated ubiquitination given that a significant portion of the overall phosphorylation activity occurs on Ser/Thr-Pro residues, accounting for one-quarter of the sites identified in global phosphorylation studies, with pSer-Pro sites outnumbering those of pThr-Pro at a ratio of five to one [21]. Hence, Pin1 is a critical mediator in the crosstalk between ubiquitination and phosphorylation, acting as a unique and pivotal post-phosphorylation mechanism for a broad range of substrates by which the protein stability and UPS interactions can be sterically modulated (Table 1).
Pin1 functionality is critically dependent on the nature of its substrates and cellular pathways. Pin1 can interconvert from cis to trans conformation or vice versa depending on the identity of the substrate and relevant structural differences, as Ser/Thr-Pro motifs may have different preferences for being in a cis or trans configuration after phosphorylation, depending on the substrate [121]. The trans conformation of peptide bonds tends to be more energetically favorable and common due to the reduced steric hinderance from adjacent amino acids [122,123]. Consequently, Pin1 typically catalyzes cis–trans isomerizations in its substrates, although trans–cis substrates also exist [124,125,126]. For example, the cadherin CDH1 is one notable substrate wherein Pin1 trans–cis isomerization contributes to changes in stability [126].

2. Pin1 and the Ubiquitin–Proteasome System

Pin1 interactions with the UPS are multi-faceted and complex, as Pin1 may regulate multiple components in a pathway using any combination of several mechanisms (Figure 1). Specifically, Pin1-catalyzed isomerization may increase or decrease the stability of a protein through a variety of mechanisms, such as by (1) modulating a substrate’s compatibility with the E3 ligase through steric alterations, (2) negatively regulating substrate E3 ligases, and (3) modifying the length of a protein’s Ub chain. An interesting note is that the involvement of Pin1 in the UPS suggests that cis or trans conformation is a major factor in the regulation of Ub-mediated proteolysis.
Firstly, Pin1 increases the rate of cis–trans isomerization so that the resulting substrate has an altered affinity for E3 ligase ubiquitination. Isomerization may expose specific structural motifs that can be recognized by the E3 ligase, leading to the ubiquitination and subsequent degradation of the target substrate. Alternatively, the new conformation may bind poorly with the E3 ligase and consequently increase the substrate’s half-life [121]. For example, Pin1 recognizes and binds phosphorylated RARα, consequently promoting its protein degradation and turnover by the UPS [90]. In contrast, the opposite is also possible, where Pin1 conformational changes can enhance the protein stability. For instance, Pin1 binds and stabilizes the many members of the p53 family of transcription factors, including Tap63α and ΔNp63α [60]. Pin1 isomerization is thought to alter the cyclization of the region surrounding the PpxY motif of p63, preventing the binding of the WWP1 E3 ligase and ubiquitination [60]. As a result, Pin1 protects the levels of Tap63α and ΔNp63α and enables them to perform apoptotic and proliferative activities, respectively [60].
Secondly, Pin1 can negatively regulate many E3 ligases. One such example is the Fbw7 Skp1-Cullin-F-box (SCF)-type complex, a well-known tumor suppressor E3 ligase that targets various oncoproteins for degradation, such as c-Myc, cyclin E, XBP1, and Notch1/4 [17,127]. Pin1 promotes the self-ubiquitination and degradation of phosphorylated Fbw7 by disrupting its dimerization, likely through Fbw7 isomerization [23]. Thus, Pin1 indirectly promotes cancer development by eliminating the Fbw7 inhibition of oncogenic activities, such as the self-renewal of cancer stem cells through Notch signaling [17]. Similarly, Pin1 depletion sensitized in vitro cancer cells to Taxol by upregulating Fwb7 and subsequently decreasing the levels of the oncogene Mcl-1 [23]. Pin1 also interacts with phosphorylated c-Jun, an oncogenic transcription factor, to promote its Fbw7-mediated ubiquitination and consequent degradation [116].
Thirdly, Pin1 can regulate the protein fate and stability by controlling the length of the Ub chain. This can determine a substrate’s ability to be recognized by Ub-binding proteins [128], as well as dictate whether the Ub-conjugated substrate will proceed towards degradation pathways or associated non-proteolytic activities, such as DNA repair or signal transduction [129]. While short mono-/oligo-Ub chains commonly lead to non-proteolytic signaling activities and long poly-Ub chains lead to degradation, these functions of Ub can occur sequentially through the elongation of the Ub chain [129]. Specifically, substrates may initially undergo mono-/oligoubiquitination for cell signaling before being polyubiquitinated and degraded. While the Ub elongation factor E4 ligase often facilitates the lengthening of these chains [129,130,131], Pin1 also participates in polyubiquitination. For example, CtlP is a DNA damage factor that is recruited to the sites of double-stranded breaks to promote homologous recombination (HR) over other lower-fidelity repair mechanisms, like non-homologous end joining (NHEJ). However, the Pin1 isomerization of phosphorylated CtlP promotes its polyubiquitination and degradation [61]. By modulating the CtlP stability, Pin1 contributes to the phosphorylation-dependent regulation of DNA double-stranded repair mechanisms, indicating that abnormal Pin1 levels, like in cancer, can compromise the efficacy of the HR and NHEJ repair mechanisms and, consequently, the genomic integrity [61]. Another well-known example involves the Pin1 substrate p53, a master regulator of tumor suppressor genes, and the E3 ligase Mdm2. High Pin1 is associated with p53 monoubiquitination and nuclear export, while Pin1 inhibition results in the polyubiquitination of p53 by Mdm2 and subsequent degradation [132]. Thus, Pin1 and Mdm2 act in concert to regulate the p53 levels.
Of note, Pin1 may also regulate phosphorylation-dependent ubiquitination by regulating other PTMs, like phosphorylation, sumoylation, and acetylation, that can crosstalk with ubiquitination to coordinate cellular processes. For example, Pin1 promotes the degradation of SUV39H1, a histone methyltransferase, inhibiting its ability to induce histone 3 trimethylation and resulting in breast cancer tumor progression [109]. Similarly, post-translational modifications to Pin1 itself can regulate its activity, with the dysregulation of Pin1 PTMs contributing to the pathogenesis of cancer and Alzheimer’s Disease, as reviewed in detail by Chen and his colleagues [133].
Ultimately, by controlling the fate of phosphoproteins, Pin1 is a unique and crucial regulator of many human diseases and cellular events. In this review, we will highlight recent advancements in the understanding of Pin1-regulated ubiquitination and its role in cancer and neurodegenerative diseases. Then, we will propose how the UPS can be leveraged in the treatment of these Pin1-associated diseases by targeting pathogenic proteins for TRIM21-mediated degradation via stereospecific antibodies.

3. Pin1-Regulated Protein Ubiquitination in Cancer

Pin1 is tightly regulated in physiological settings and typically acts to coordinate cellular processes, like cell cycle progression and DNA repair. Typically, Pin1 is expressed at low levels in most normal cells, such as fibroblasts and epithelial cells, although it may be highly expressed in some cell subsets and certain tissues [134]. However, in most solid malignancies, like breast, prostate, and lung cancer [134,135,136], Pin1 has been found to be commonly overexpressed when compared to their physiological counterparts [134]. One study found that in 38 out of 60 human cancer types tested, Pin1 was overexpressed in 10% of these cases, suggesting that Pin1 overexpression is a critical event for tumorigenesis and may serve as an amplifier of oncogenic signaling [134]. While the mechanisms behind Pin overexpression are not fully understood, it has been posited that this trend is a result of the breakdown of the transcriptional and post-transcriptional mechanisms that usually keep Pin1 tightly regulated [137,138]. For instance, the deregulation of E2F during breast cancer promotes Pin1 expression [137], while the downregulation of the inhibitory phosphorylation of Pin1 leads to the accumulation of active Pin1 [138].
Indeed, with few exceptions, Pin1 promotes cancer progression and cancer stem cell expansion by increasing the stabilities of over 70 oncoproteins while decreasing the stabilities of over 35 tumor suppressors (Table 1) [139]. Consequently, the Pin1 levels in cancer are often correlated with poor clinical outcomes and act as a key prognostic marker in many cancer types [40,54,135,140,141]. Thus, targeting Pin1 has emerged as a viable clinical strategy, with one recent study finding that Pin1 inhibition in a pancreatic cancer model disrupted multiple cancer pathways and the immunosuppressive microenvironment, rendering it eradicable by immunochemotherapies [142]. In this section, we will explore recent discoveries in the Pin1 enhancement of oncogene stability and tumor suppressor degradation.

3.1. Pin1 Enhances Oncogenic Protein Stability

Excess Pin1 activity can divert substrates from degradation, enabling the enhanced stabilities of oncoproteins and the promotion of tumorigenesis by acting on cancer stem cells as well as stromal cells. One key hallmark of cancer cells is their ability to adapt the metabolic processes of both their internal environment and the tumor microenvironment to accommodate their increased proliferation and transformation [143]. One way in which Pin1-regulated ubiquitination contributes to the metabolic reprogramming of cancer cells is by stabilizing critical transcription factors, such as hypoxia inducible factor (HIF), as reviewed recently by Nakatsu and colleagues [144]. Under oxygen-deficient conditions, HIF, a master regulator of oxygen homeostasis, upregulates the stress response to promote tumor survival. HIF is composed of two subunits: one alpha subunit of which there are three subtypes (HIF-1α, HIF-2α, HIF-3α) and one beta subunit (HIF-1β) [145]. The most prominently studied subunits are the oxygen-dependent alpha subunits HIF-1α and HIF-2α, which are both necessary for cancer cell viability in the hypoxic tumor microenvironment [84]. In normoxic conditions, HIF-1α is regulated by prolyl hydroxylase domain-containing proteins (PHD) and the Von Hippel–Lindau tumor suppressor protein (pVHL), which each coordinate to promote the rapid ubiquitination and degradation of HIF-1α [146]. Consequently, HIF-1α has a short half-life of under 5 min in normoxia [147]. During hypoxia, however, HIF-1α is phosphorylated at Ser451E, allowing Pin1 to catalyze its conformational change [63,64,65]. The phosphorylation of HIF-1α is essential for its transcriptional activity and disrupts interaction with PHD- and pVHL-mediated protein degradation, significantly enhancing the stability of this oncoprotein [148]. Pin1 also interacts with CDK1-phophorylated pVHL, recruiting the E3 ligase WSB1 and facilitating pVHL ubiquitination and degradation [11]. Interrupting the Pin1/CDK1/pVHL axis results in decreased cancer cell proliferation, migration, invasion, and chemoresistance, and therefore it could be therapeutically valuable in treating cancers with wild-type VHL [11].
Furthermore, Pin1 interacts with the CDK1/2 kinase and Cul3-KLHL20 Ub ligase to degrade promyelocytic leukemia (PML), which is a tumor suppressor that typically prevents HIF-1α translation by suppressing mTOR after HIF-1α activation [100,148]. When PML is dephosphorylated at Ser518 by the phosphatase SCP1, Pin1 and KLHL20 are unable to mediate the ubiquitination and degradation of PML, thereby inhibiting the progression and growth of both in vitro and in vivo clear-cell renal carcinoma models [100]. Restored PML levels suppress the mTOR-HIF pathway, enhancing the effects of the mTOR inhibitor Temsirolimus and indicating the potential for combination treatments that target PML degradation and the mTOR pathway [149]. Additionally, Pin1 facilitates the sumoylation of PML by SUMO1 in glioma stem cells, which enables PML to interact with and stabilize c-Myc, thereby permitting the survival and carcinogenic potential of glioma stem cells [150]. Similarly, in HIF-2α, Pin1 directly binds to the subunit’s phosphorylated Ser790 site located in the nuclear export signal domain [84]. Phosphorylation at this site is essential for the stability and transactivation of HIF-2α and enables its activity even in normoxic conditions, enabling the progression of breast cancer cells [84].
Recently, Pin1 was found to promote tumor cell survival by stabilizing nuclear factor erythroid 2–related factor 2 (Nrf2). The transcription factor Nrf2 controls the expressions of genes that provide defense against both internal and external stressors, including xenobiotics and reactive oxygen species (ROS) [151]. While the transient activation of Nrf2 protects the cell from oncogenic insults, its accumulation in cancer cells allows for the constitutive induction of antioxidant and detoxifying pathways, promoting cancer proliferation in hypoxic environments and resistance to drug therapies [152]. Accordingly, Nrf2 is tightly regulated by a complex network of pathways, with the Keap1-Cul3-Rbx1 axis being the most prominent. Keap1 is an adapter subunit of the Cul3/Rbx1 E3 ligase and sequesters Nrf2 in the cytosol, leading to its continuous ubiquitination and degradation [153]. Upon encountering stressors like oxidative or electrophilic agents, Nrf2 dissociates from Keap1, allowing Pin1 to directly interact with its Ser215, Ser408, and Ser577 sites [154]. Consequently, Nrf2 is stabilized and localized to the nucleus, allowing it to transactivate cognate oncogenic pathways [83,154]. Interestingly, Pin1 also binds to the phosphorylated Ser104 and Thr277 sites of Keap1, potentially allowing Pin1 to compete with Keap1 for Nrf2 binding [154]. Furthermore, Pin1 may interact with other oncoproteins to promote Nrf2 activity. Pin1 interacts with c-Myc to bind to the Nrf2 promoter, enhancing the positive effect of c-Myc on the promoter activity [154]. Additionally, H-Ras, also known as transforming protein p21, is an oncogenic GTPase that regulates cell survival and differentiation and is correlated with increased Pin1-Nrf2 interaction [155]. Downstream, Pin1 upregulates the glutathione peroxidase 4 (GPX4) expression via Nrf2, promoting chemotherapy resistance [156]. Lastly, Pin1 has also been found to regulate the extracellular matrix composition and redox balance in pancreatic cancer through the NRF2/ARE pathway [157].
Pin1 has also been identified as a novel regulator and protein stabilizer in the oncogenic Hippo pathway, which controls the organ size by promoting cell proliferation, apoptosis, and cell stemness [158]. The Hippo pathway is divided into a tumor-suppressing serine–threonine kinase regulator module and an oncogenic transcriptional module, the latter in which Yes-associated protein (YAP) and PDZ-binding protein (TAZ) transcription factors participate [158]. Cancer cells specifically leverage YAP/TAZ to upregulate genes associated with increased stem cell renewal, cell proliferation, and apoptosis resistance. The amplification of the effects of other transcription factors, including AP-1, E2F, Myc, and CTGF, is also facilitated by YAP/TAZ, which functions as an integrator of several carcinogenic pathways, including the Wnt pathway [75,158]. Ortega and colleagues conducted a thorough review of this subject [158]. Pin1 promotes tumorigenesis by stabilizing YAP and TAZ, as well as by promoting their nuclear localization and transactivation [75,76], resulting in increased drug resistance and tumorigenicity in a breast cancer model [76]. While the exact mechanism is still being clarified, Pin1 interacts with YAP and TAZ in a phosphorylation-independent manner, suggesting that it regulates YAP/TAZ indirectly [76]. The downregulation of Pin1-YAP/TAZ activity using Cinobufacini injections is an effective treatment for osteosarcoma [159].
With regard to other components of the Hippo pathway, Pin1 also interacts with LATS1 and LATS2, two upstream kinases that phosphorylate and inhibit YAP and TAZ in the context of the anti-tubulin drug response [160]. The administration of anti-tubulin chemotherapies hyperactivates the CDK1-mediated phosphorylation of LATS2 and subsequent interactions with Pin1; the authors of this study purport that Pin1 catalyzes the conformational change of LATS1/2, switching its preferred substrate from YAP/TAZ to Pin1 and augmenting anti-tubulin-induced apoptosis [160]. Furthermore, Pin1 induces the ubiquitin-mediated degradation of STK3, a kinase that activates LATS1/2, resulting in increased tumorigenicity in melanoma mouse models [75]. As illustrated, Pin1 stabilizes a broad range of oncogenes, some of which include BRD4 in oncogenic gene expression [71], estrogen receptor-alpha in tumorigenic signaling [57], and tissue factors that can contribute to angiogenesis [72,73].

3.2. Pin1 Decreases Tumor Suppressor Protein Stability

Conversely, Pin1 may also disrupt the balance between oncoproteins and tumor suppressors by promoting the degradation of key tumor suppressors. Pin1 is a well-known regulator of many CDK substrates, including cyclin E, GRK2, tau, and TRF-1 [128], and it has recently been identified as a regulator of CDK substrates, like tripartite motif family 59 (TRIM59). TRIM59 is an E3 ligase and CDK5 substrate that mediates the CDK5 progression of glioblastoma. TRIM59 is phosphorylated at Ser308 by the CDK5-activated epidermal growth factor receptor (EGFR), which then recruits Pin1 to catalyze the cis–trans isomerization of TRIM59 [74]. This exposes the nuclear localization sequence of TRIM59 and facilitates its nuclear localization, wherein TRIM59 subsequently degrades macroH2A1, a tumor-suppressive histone variant [74], and inhibits TC45 from the dephosphorylation of STAT3 [74,161]. Both these mechanisms result in the sustained upregulation of STAT3 and increased glioblastoma tumorigenicity, and they are linked to a poor clinical prognosis in glioblastoma patients [74]. Similarly, TRIM32 is an E3 ligase and CDK2 substrate that mediates the CDK2 promotion of radio resistance in triple-negative breast cancer [85]. Radiotherapy stimulates the CDK2 phosphorylation of Ser328 and Ser339 on TRIM32, resulting in the subsequent Pin1 isomerization and nuclear localization of TRIM32 [85]. Nuclear TRIM32 inhibits TC45, leading to upregulated STAT3 and radio resistance in triple-negative breast cancer [85].
In addition to CDK substrates, two major regulators of cell cycle progression that interact closely with CDKs are the retinoblastoma protein (Rb) and anaphase-promoting complex (APC/C), both of which functionally collaborate to control the G1/S transition [162]. Both the APC/C and Rb are Pin1-regulated tumor suppressors that are largely inactivated in cancer [126]. Rb inhibits S-phase entry by sequestering the key transcriptions factors necessary for DNA replication during the S phase [163]. However, near the end of G1, Rb is hyperphosphorylated by cyclin–CDK complexes so that it releases chromatin-modifying enzymes and transcription factors that mediate the G1/S transition [163]. In addition, Pin1 catalyzes a conformation change in Rb, facilitating the release of the E2F transcription factor while also enabling the Rb to become fully phosphorylated [164,165,166].
Yet, Rb is frequently disrupted in many aggressive human cancers, such as triple-negative breast cancer, due to the constitutive expression of CDKs [167]. Thus, the APC/C becomes the sole regulator of the G1/S transition in these Rb-deficient cancers [126]. Recently, the involvement of Pin1 in regulating the APC/CCDH1 was uncovered. When bound to its CDH1 co-activator, the APC/CCDH1 E3 ligase typically blocks cell cycle entry by targeting various critical mitotic proteins for degradation [126]. Near the end of the G1 phase, CDKs become more active and phosphorylate CDH1 at Ser163 [126]. Pin1 then binds to CDH1 to catalyze a trans–cis isomerization that renders the co-activator resistant to dephosphorylation, inactivating the APC/CCDH1 and allowing mitotic proteins to accumulate and propel the cell into the S phase [126]. However, when CDH1 is not dephosphorylated, such as in the G0/G1 phase, the APC/CCDH1 is active and recognizes the D-box motif in the PPIase domain of Pin1, ubiquitinating Pin1 and other mitotic proteins for proteolysis and preventing S-phase entry [126]. Therefore, when Pin1 is overexpressed and CDK is constitutively expressed, like in cancer cells, the G1/S regulatory checkpoint breaks down as the APC/CCDH1 is inhibited and the cell proliferates unchecked [126].
Interestingly, the APC/CCDH1-mediated degradation of Pin1 explains why effective Pin1 inhibitors, like Sulfopin and ATRA-ATO, induce Pin1 degradation [68,142,168,169,170]. By blocking Pin1 enzymatic activity, the balance of dephosphorylated CDH1 to phosphorylated CDH1 may increase and promote APC/CCDH1 activation and the degradation of Pin1 [126]. Significantly, by combining Pin1 and CDK4 inhibitors to prevent APC/CCDH1 inactivation, the APC/CCDH1 induces the degradation of Pin1 and other mitotic proteins, forcing the cells into cell cycle arrest and triggering increased anti-tumor immune activity [126]. This combination treatment was highly effective in delaying cancer progression in a triple-negative breast cancer model, both disrupting the immunosuppressive tumor microenvironment and enhancing the anti-tumor immunity [126].
As illustrated, Pin1 regulates a substantial range of CDK substrates, including SEPT9 for cytokinesis completion [171], Smad3 for oncogenesis in CDK triple-negative breast cancer [172,173], and even p27, a CDK inhibitor that is regulated by Pin1 [42]. However, beyond CDK substrates, Pin1 can also directly regulate CDKs themselves, specifically CDK10. CDK10 directly interacts with the WW domain of Pin1 and becomes ubiquitinated and degraded [105]. Among other substrates, CDK10 inhibits the phosphorylation and activation of Raf-1, a key regulator for cell proliferation and survival [174]. The downregulation of CDK10 facilitates increased Raf-1 phosphorylation, resulting in increased Raf-1 activity and tamoxifen resistance in breast cancer cells [105].
Furthermore, Pin1 has also been found to regulate multiple components of the DNA repair pathway. Recently, Pin1 has also been found to promote the stability of FAAP20 in the Fanconi anemia pathway [79], to counteract excessive chromatin ubiquitination by promoting the sumoylation of the RNF168 E3 ligase [175], and to enhance the genoprotective activity of the BRCA1-BARD Ub ligase complex [78]. The Pin1 stabilization of BRCA1 also contributes to the maintenance of genomic integrity and checkpoint controls. Specifically, after ionizing radiation triggers the phosphorylation of BRCA1, Pin1 isomerizes BRCA1 at p-Ser1191 and prevents its ubiquitination at Lys1037, thereby promoting BRCA1 activity at nuclear DNA repair foci [80]. The inhibition of Pin1 removes BRCA1-mediated radio resistance and homologous recombination repair mechanisms, sensitizing breast cancer cells to radiation and PARP inhibitor treatments, which both induce stress and DNA damage that the cells can no longer fully alleviate [80].
Lastly, Pin1 was recently found to induce the ubiquitination and degradation of RUNX3, a tumor suppressor transcription factor that controls apoptosis, cell differentiation, and metastasis [110]. Pin1 binds to any of the four phosphorylated Ser/Thr-Pro sites on RUNX3 and downregulates its transcriptional activity and stability, thereby promoting breast cancer progression [110]. Similarly, tamoxifen resistance is conferred on and enhanced in breast cancer cells by Pin1-mediated SGK1 degradation [176].

4. Pin1-Regulated Ubiquitination in Neurodegenerative Disease

Compared to other cells in the body, Pin1 is highly expressed throughout the central and peripheral nervous systems [86,177]. While the role of Pin1 in regulating physiological neuronal functions was previously not well known, within the last decade, there has been mounting evidence of its integral role in promoting neuronal differentiation, axonal growth, synaptic transmission, and apoptosis [178]. In contrast, Pin1 attenuation in neurodegenerative diseases has been long well established, with Pin1 being most famously known to suppress Alzheimer’s Disease (AD) by regulating tau and amyloid precursor protein turnover and degradation [179,180]. In addition, the Pin1 modulation of proteins central to other neurodegenerative diseases, such as Parkinson’s Disease (PD) and Huntington’s Disease (HD), is an emerging area of research that should not be neglected. In this section, we will examine recent advancements in our understanding of Pin1’s role in negatively regulating the stability and ubiquitin-mediated degradation of proteins associated with neurodegenerative diseases, specifically AD, PD, and HD.

4.1. Pin1-Regulated Ubiquitination in Alzheimer’s Disease

AD is classically characterized by the abnormal aggregation of tau into neurofibrillary tangles and APP into amyloid-beta plaques. Firstly, physiological tau plays a crucial role in promoting microtubule stability and axonal outgrowth; however, when tau becomes hyperphosphorylated, it dissociates from microtubules and aggregates into the pathological neurofibrillary tangles that give rise to the synaptic dysfunction and neurodegenerative symptoms of AD [181]. Pin1 regulates tau by directly binding pThr231-Pro to catalyze a cis–trans conformation change, facilitating pThr231 dephosphorylation via protein phosphatase 2A and CDK5-p25 [182]. However, Pin1 is reduced in the AD brain by tangle sequestration [104,108], downregulation [87,183,184], phosphorylation at S71 by DAPK1 [185,186,187], oxidation at C113 [188,189], or even degradation induced by the environmental pollutant cobalt [190].
Indeed, the recent development of conformation-specific antibodies has offered the first opportunity to distinguish the cis from trans conformation of tau in AD pathogenesis. Specifically, cis-P-tau does not participate in microtubule assembly, is prone to aggregation, and resists dephosphorylation and Ub-mediated degradation, as opposed to trans-P-tau, which cannot self-associate into paired helical filaments and instead preferentially binds to microtubules to stabilize axonal transport [191]. This neuroprotective role of Pin1 is exemplified in gene manipulation studies, wherein Pin1 knockout induced tauopathy and increased tau stability in murine models while Pin1 overexpression suppressed tauopathic phenotypes [192]. Thus, Pin1 protects against tangle formation by inducing a cis–trans conformation change that restores the biological activity of phosphorylated tau to bind and promote microtubule assembly [86].
During AD, however, neurons are commonly exposed to oxidative stress, and, consequently, Pin1 may be aberrantly modified by oxidation, leading to a loss of function and polyubiquitination [107,179,193]. Furthermore, the loss of Pin1 in synapses during AD progression affects the ubiquitination of postsynaptic density proteins, a huge protein complex associated with the membrane of postsynaptic excitatory synapses that regulates synaptic plasticity [194,195]. This results in the polyubiquitination and degradation of the Shank3 protein, which regulates the organization of the postsynaptic density, ultimately leading to an increased vulnerability to the toxic effects of amyloid-beta and glutamate while also decreasing synaptic plasticity during AD progression [194].
Moerover, Pin1 is also a critical regulator of the amyloid precursor protein (APP), from which amyloid-beta (Aβ), the main component of AD-chatracteristic amyloid plaques, is derived. Physiologically, the APP plays an important role in brain development, memory, and synaptic plasticity, although its full role is not yet fully understood due to the complex manner in which the APP is processed by various secretases [196]. Pin1 binds the phosphorylated Thr668-Pro motif of the APP and catalyzes a cis–trans conformation change that reduces Aβ peptide secretion [197]. Mechanistically, it is hypothesized that, as opposed to trans-APP, the cis-APP conformation favors amyloid formation by β-secretase cleavage [197,198]. Moreover, another mechanism by which Pin1 protects against AD is by upregulating the UPS-mediated APP turnover by promoting the degradation of Glycogen Synthase Kinase-3β (GSK3β), a kinase that regulates the ubiquitination of several cancer or neurodegenerative-associated proteins in conjunction with Pin1 [177]. Other GSK3β substrates have been reviewed by Liou and colleagues [121]. By binding to Thr330-Pro, Pin1 inhibits the GSK3β phosphorylation of the APP, allowing the APP to subsequently be degraded [177]. Pin1 may also regulate GSK3β substrates; for instance, in a rat hippocampal cell culture model, Pin1 isomerized GSK3β-phosphorylated cis-HIF-1α to mediate the polyubiquitination and consequent degradation of trans-HIF-1α by the UPS [63]. This becomes therapeutically relevant in AD, wherein Pin1 and GSK3β may be downregulated under hypoxic or ischemic conditions, allowing cis-HIF-1α to accumulate and contribute to the increased expression of β-secretase 1 and the consequent amyloidogenic APP processing [63,199,200].

4.2. Pin1-Regulated Ubiquitination in Parkinson’s Disease

PD pathogenesis centers around the accumulation of alpha-synuclein (α-syn), a pre-synaptic neuronal protein that aberrantly misfolds into Lewy-body aggregates that perturb dopaminergic signaling, ultimately resulting in neuronal death and neurodegeneration [201,202]. Although α-syn plays a physiological role in neurotransmitter release, elevated levels of α-syn drive toxicity; thus, it is crucial for synuclein production and clearance to be tightly regulated and balanced [202]. Pin1 is commonly localized to Lewy bodies in PD brain tissue, indirectly interacting with α-syn to inhibit its degradation, thereby increasing the half-life and insolubility of α-syn and facilitating the formation of Lewy bodies [203]. In a PD murine model, the overexpression of Pin1 increased α-synuclein aggregation and upregulated pro-apoptotic cascades in dopaminergic neurons, contributing to PD onset and the associated neurodegeneration [204]. Treatment with a Pin1 inhibitor ameliorated PD-associated motor deficits, neurochemical depletion, and dopaminergic neuron degeneration in this experimental PD animal model [204]. Furthermore, Pin1 directly binds synphilin-1, a protein that participates in Lewy-body formation with α-syn, at the phosphorylated Ser-211-Pro and Ser-215-Pro motifs to promote its interaction with α-syn [204,205].

4.3. Pin1-Regulated Ubiquitination in Huntington’s Disease

HD is a neurodegenerative disease caused by mutant Huntingtin (Htt) protein with an elongated segment of glutamine residues [206]. Although Htt plays a critical physiological role in early development as well as in post-developmental activities, like axonal trafficking and anti-apoptotic signaling [207], mutant Htt tends to form intranuclear and cytoplasmic aggregates in neurons, especially those in the striatum, which results in cytotoxicity and HD-associated neurodegenerative symptoms [208,209]. Pin1 is playing an emerging role in HD pathogenesis. For instance, in an HD murine model, Pin1 was essential for mediating the downstream apoptotic effects of Htt, specifically by binding to the phosphorylated Ser46-Pro motif of p53, a master tumor suppressor protein, and stabilizing it into an activatable form [210]. This implies that mutant Htt may indirectly induce neuronal death by acting as an upstream inducer of Pin1-isomerized p53, rather than directly interact with it [210]. Additionally, p53 has been shown to upregulate mutant Htt expression, suggesting a positive-feedback loop [210]. Further studies by the same group using HdhQ111 knock-in mice demonstrated that Pin1 is involved in HD pathogenesis throughout the lifespan of the animal [211]. In early life, Pin1 participates in the DNA damage response, presumably by stabilizing p53 [211]. For mid-age mice, Pin1 regulates hormone synthesis as well as Wnt/β-catenin signaling by stabilizing the substrate β-catenin to control neuronal differentiation [114,211]. To note, the mutant Htt stabilization of β-catenin has been demonstrated to also contribute to striatal neurotoxicity in a Drosophila HD model [212].
However, the ablation of Pin1 in this same HdhQ111 model also decreased the mutant Htt aggregate load in the mouse striatum [211]. Similarly, a subsequent study by the same group found that Pin1 upregulated the mutant Htt clearance and reduced aggregation in an in vitro model through indirect interactions [115]. In particular, Pin1 indirectly promotes the degradation of both wild-type and mutant Htt through the UPS or autophagy [115].
Ultimately, the conflicting roles of Pin1 and the associated ubiquitination-mediated degradation in HD pathogenesis reveal the exciting therapeutic potential for targeting Pin1 manipulation, especially as previous work has demonstrated that, despite an initial impairment, the UPS functions normally during HD [213,214,215,216,217]. Moreover, moderate depletion of wild-type and mutant Htt to levels no lower than 50% does not introduce any major detrimental effects, as the mutant Htt retains some physiological activity [218,219,220,221,222,223]. This was evident in a recent clinical trial wherein a maximum reduction of 42% of mutant Htt did not yield any clinically relevant adverse effects [224]. Nevertheless, much investigation is required to elucidate the exact nature of these mechanisms in HD.

4.4. Pin1-Regulated Ubiquitination in Other Neurological Disorders

Beyond classical neurodegenerative diseases, Pin1-mediated ubiquitination has emerged as a regulator of adjacent neurological disorders, such as epilepsy, stroke, spinal cord injury, retinal disease, and even physiological aging, although its role is much less defined.
In epilepsy, early studies on the role of Pin1, as recently reviewed by Chen and colleagues [225], suggest that Pin1 regulates the stabilities of several disease-related proteins. For example, Pin1 binds and isomerizes protein kinase C, which is tenuously associated with the neurotransmitter balance and neuronal hyperexcitability, subsequently priming it for agonist-induced, ubiquitin-mediated degradation [107,226,227]. To note, Pin1 also regulates synaptic transmission by destabilizing and promoting the degradation of the postsynaptic density protein 95 (PSD-95), an integral regulator of synaptic plasticity and excitatory signaling, the dysregulation of which has been implicated in epileptic seizure generation [118,227,228]. As PSD-95 anchors NMDA receptors in the postsynaptic membrane, Pin1 isomerization induces a structural change that interferes with this interaction and results in a decrease in NMDA receptors, negatively affecting NMDA signaling and the dendritic spine morphology [229].
Pin1 also regulates the Notch pathway in ischemic stroke pathogenesis to promote neuronal death. In particular, Pin1 interaction with the Notch1 intracellular domain (NCID) potentiates γ-secretase cleavage that inhibits the Fwb7-induced polyubiquitination of NCID1, thereby preventing its ubiquitin-mediated proteolysis and enabling its enhancement of neuronal death after simulated ischemia in an in vitro model [45]. The subsequent treatment of Pin1-knockout mice with Pin1 inhibitors demonstrated reduced brain damage and improved functional outcomes after the induction of focal ischemic stroke [45]. In addition, Pin1 not only enhances Notch1 activity, but Notch1 also directly induces Pin1 transcription, forming a positive-feedback loop [39]. Notably, Pin1 also regulates other key players in ischemic stroke, such as NF-kβ, HIF-1α, and p53, although the crosstalk between these other molecules should not be understated [230]. For example, Pin1 directly binds to the phosphorylated Ser33/46-Pro of p53 and stabilizes it, allowing it to induce downstream apoptotic gene expression and the mitochondrial cell death pathways [28]. Specifically, rescued p53 subseqeuntly interacts with Notch to promote apoptotic pathways in neurons, comprising their viability and resistance to ischemic damage and ultimately promoting pathogenesis after ischemic stroke [231].
Moreover, in the context of spinal cord injury, Pin1 positively regulates the stability of several anti-apoptotic proteins that resist neuronal death after traumatic injury. Normally, Pin1 plays a physiological role in binding myeloid cell leukemia sequence-1 (Mcl-1) at phosphorylated Thr163-Pro and preventing its ubiquitination [232]. However, upon injury, Pin1 directly binds the phosphorylated Ser178-Pro motif in the death domain-associated protein and promotes its rapid degradation by the UPS [97]. This activates the ASK1/JNK pathway, specifically JNK3 [233,234], which facilitates the phosphorylation of Mcl-1 at Ser121-Pro and a subsequent conformational change that leads to Pin1 dissociation [232]. The consequent Mcl-1 degradation inhibits the B-cell lymphoma 2 protein, enabling the initiation of mitochondrial-mediated apoptosis via cytochrome c release [33,232]. Conversely, Pin1 upregulates neuronal apoptosis by stabilizing BIMEL via binding at phosphorylated Ser65-Pro; uniquely, this pathway suppresses cell death in non-neural cells but increases apoptosis in neurons, likely due to the presence of the neuron-specific JNK scaffold protein JIP, which promotes Pin1 binding to BIMEL [33].
These examples highlight the involvement of Pin1 in various neurological disorders, especially as Pin1 regulates various pro-apoptotic proteins, as described above, that are involved in neuronal death. Paradoxically, Pin1-mediated ubiquitination seems to play context-dependent, contrasting roles that are still not fully understood—for instance, promoting cell death in stroke, but binding protectively in spinal cord injury. Delineating the specific protein networks governed by Pin1 could provide invaluable roadmaps to understand disease triggers and guide therapeutic development. Thus, Pin1-mediated ubiquitination provides an intriguing perspective to developing therapeutics for a wide range of neurological dysfunctions.

5. Conformation-Specific Antibodies Target Intracellular Proteins for TRIM21-Mediated and Ubiquitination-Mediated Proteolysis

The neutralization ability of antibodies was long thought to be limited to the extracellular space, as they could not penetrate the cell membrane. With few exceptions, natural antibodies could not be used in live cells without permeabilization, making a wide range of intracellular targets beyond the reach of antibody-mediated detection, visualization, and inhibition [235]. Currently, most therapeutic antibodies are mostly restricted to extracellular membrane proteins or secreted proteins, which are estimated to make up 8000–9000 proteins [236] of the estimated 20,000 proteins in the human proteome [237]. This implies that targeting the intracellular proteome could unlock around two-thirds of the human proteome available for therapeutic intervention [236]. Among these intracellular targets are notorious cancer-driving intracellular targets, such as NF-kB and c-Myc, that have insofar been undruggable by small molecular inhibitors [238,239]. Although many antibodies cannot naturally penetrate live cells and require further engineering to acquire such a property [240], it has been shown that a subset of human and mouse antibodies produced in autoimmune disease patients [241,242,243,244,245] and in labs [246,247,248] can penetrate live cells and capture their intracellular antigens without engineering [234,249,250]. Monoclonal antibodies against many intracellular proteins, including oncoproteins [251,252], tau [253,254,255,256,257,258,259], TDP43 [260], and RANs [261], to name a few, are used as therapeutics without engineering, with some being evaluated in human trials [262]. Thus, intracellular antibodies represent a burgeoning extension of antibody therapeutics and research tools, allowing the study and treatment of many intracellular disease-driving targets.
Moreover, the recent discovery of TRIM21 introduces a mechanism by which intracellular antibodies may mediate their effects. TRIM21 is a cytosolic E3 Ub ligase and antibody receptor that acts as a final line of defense against invading viruses. Its physiological role is to intercept and neutralize antibody-coated viruses that have evaded extracellular immune defenses and entered the cell [263]. Specifically, TRIM21 recognizes the Fc of antibodies bound to invading pathogens and catalyzes K63-ubiquitin chain formation, stimulating proinflammatory pathways and anti-viral cell activity [264]. These pathogens are rapidly targeted for degradation by the proteasome and trigger innate immune activity, allowing TRIM21 to act as a bridge between adaptive and innate immune mechanisms [265].
In the context of Pin1 substrates, natural monoclonal antibodies have been developed that are able to reliably target cis or trans protein isomers inside the cell [266,267,268,269,270,271]. By targeting only one specific conformation of a protein for TRIM21-mediated protein degradation, it is possible to selectively deplete the toxic disease-driving isomers while retaining the beneficial forms (Figure 2). If used in combination with Pin1 inhibitors like Sulfopin [170] and a combination of arsenic trioxide and all-trans retinoic acid [169], which both block the further Pin1 generation of the target isomer, intracellular antibodies represent a fascinating prospect for treating certain Pin1-related pathologies.
Consistent the findings that Pin1 protects against tau-associated neurodegenerative diseases by catalyzing cis–trans isomerization, the accumulation of cis-P-tau has been shown to be an early biomarker and etiological factor for the neurodegenerative effects of traumatic and vascular brain injury and long-term sequelae like AD, chronic traumatic encephalopathy, and vascular dementia [266,267,268,269,270,271]. After severe or repetitive TBI in mice, cis P-tau mAb treatment not only eliminates cis P-tau231 and cistauosis but also prevents short- and long-term neuropathology and brain dysfunction [267,268,269]. In htau mice of AD, eliminating cis P-tau231 at a young age with cis mAb treatment prevented age-dependent tauopathy, neuronal loss, and learning and memory impairment [270]. Moreover, even after an NFT-like pathology and cognitive loss had already developed in aged htau mice, treating the mice with cis mAb still prevented cis P-tau231 accumulation and rescued the neuronal loss and brain atrophy [270]. Humanized cis P-tau mAb also reduced tau seeding and attenuated neuropathological and behavioral endpoints in transgenic mice expressing inducible mutant P301L tau, which has been linked with familial frontotemporal dementia [271]. In stroke models, eliminating cis P-tau231 using cis mAb rescues most stroke-like pathologies and brain dysfunction, and the efficacy is still strikingly obvious even after reducing the cerebral blood flow by ~50% for 6 months, which was confirmed by the unbiased single-cell transcriptomic profiling of VCID mice [270]. Thus, cis P-tau mAb might provide a prophylactic therapy that blocks the onset of neurodegeneration or a therapeutic intervention that is given after disease onset to reduce its severity in AD, TBI/CTE, and VCID [270]. A phase 1 trial of the i.v.-administered humanized cis P-tau231 mAb (PNT001) did not show obvious side effects and also produced CSF concentrations of PNT001 that suggest a potentially therapeutic effect [272].
Notably, TRIM21 mediates the neutralization of antibody-bound tau by promoting tau y ubiquitination and proteolysis via the proteasome and VCP unfoldase [268,273]. In mouse models, the antibody targeting of extracellular tau aggregates relied on TRIM21-mediated degradation after the internalization of the tau–antibody complex [274]. Antibody protection was lost in TRIM21-deficient mice [274].
Meanwhile, in immunotherapy for other diseases, antibodies have found massive success in pivotal advances such as immune checkpoint inhibitors and antibody–drug conjugates, yet their targets are still mostly relegated to the extracellular domain. The advent of stereospecific antibody technology that can capture intracellular targets and target them for TRIM21-mediated degradation introduces a huge potential for the treatment of a wide range of diseases in which dysfunctional Pin1 leads to the accumulation of a pathological protein. Because an abundance of Pin1 substrates, their regulatory mechanisms, and their pathological interactions have already been identified, future work will likely center around selecting appropriate molecular targets and developing stereospecific antibodies for translational experiments.

Funding

This work was supported by the Canada Foundation for Innovation (CFI) (grants #43257 and #43822), Canadian Institutes of Health Research (#502128, #506475 and #506500) and Ontario Institute for Cancer Research (#1240) to K.P.L. and X.Z.Z.

Acknowledgments

We thank the lab members of the corresponding authors for their critical reading of the manuscript and for providing insightful suggestions. All figures were generated with BioRender.com (accessed on 1 February 2024).

Conflicts of Interest

X.Z.Z. and K.P.L. are the scientific founders and former scientific advisors of Pinteon, in which they own equity, as well as the inventors of several issued patents and/or pending patent applications related to Pin1. Their interests were reviewed and are managed by Western University in accordance with its conflict-of-interest policy. The remaining authors declare no conflicts of interest.

References

  1. Nguyen, L.K.; Kolch, W.; Kholodenko, B.N. When ubiquitination meets phosphorylation: A systems biology perspective of EGFR/MAPK signalling. Cell Commun. Signal. 2013, 11, 52. [Google Scholar] [CrossRef] [PubMed]
  2. Lecker, S.H.; Goldberg, A.L.; Mitch, W.E. Protein degradation by the ubiquitin-proteasome pathway in normal and disease states. J. Am. Soc. Nephrol. 2006, 17, 1807–1819. [Google Scholar] [CrossRef] [PubMed]
  3. David, Y.; Ternette, N.; Edelmann, M.J.; Ziv, T.; Gayer, B.; Sertchook, R.; Dadon, Y.; Kessler, B.M.; Navon, A. E3 ligases determine ubiquitination site and conjugate type by enforcing specificity on E2 enzymes. J. Biol. Chem. 2011, 286, 44104–44115. [Google Scholar] [CrossRef] [PubMed]
  4. Kaiser, P.; Fon, E.A. Expanding horizons at Big Sky. Symposium on ubiquitin and signaling. EMBO Rep. 2007, 8, 817–822. [Google Scholar] [CrossRef]
  5. Li, W.; Ye, Y. Polyubiquitin chains: Functions, structures, and mechanisms. Cell. Mol. Life Sci. 2008, 65, 2397–2406. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, C.; van Dyk, D.; Xu, P.; Choe, V.; Pan, H.; Peng, J.; Andrews, B.; Rao, H. Ubiquitin chain elongation enzyme Ufd2 regulates a subset of Doa10 substrates. J. Biol. Chem. 2010, 285, 10265–10272. [Google Scholar] [CrossRef]
  7. Theuerkorn, M.; Fischer, G.; Schiene-Fischer, C. Prolyl cis/trans isomerase signalling pathways in cancer. Curr. Opin. Pharmacol. 2011, 11, 281–287. [Google Scholar] [CrossRef]
  8. Ping Lu, K.; Hanes, S.D.; Hunter, T. A human peptidyl-prolyl isomerase essential for regulation of mitosis. Nature 1996, 380, 544–547. [Google Scholar] [CrossRef]
  9. Lummis, S.C.; Beene, D.L.; Lee, L.W.; Lester, H.A.; Broadhurst, R.W.; Dougherty, D.A. Cis–trans isomerization at a proline opens the pore of a neurotransmitter-gated ion channel. Nature 2005, 438, 248–252. [Google Scholar] [CrossRef]
  10. Ryo, A.; Suizu, F.; Yoshida, Y.; Perrem, K.; Liou, Y.C.; Wulf, G.; Rottapel, R.; Yamaoka, S.; Lu, K.P. Regulation of NF-kappaB signaling by Pin1-dependent prolyl isomerization and ubiquitin-mediated proteolysis of p65/RelA. Mol. Cell 2003, 12, 1413–1426. [Google Scholar] [CrossRef]
  11. Chen, J.; Li, M.; Liu, Y.; Guan, T.; Yang, X.; Wen, Y.; Zhu, Y.; Xiao, Z.; Shen, X.; Zhang, H.; et al. PIN1 and CDK1 cooperatively govern pVHL stability and suppressive functions. Cell Death Differ. 2023, 30, 1082–1095. [Google Scholar] [CrossRef] [PubMed]
  12. Hu, H.; Wulf, G.M. The amplifier effect: How Pin1 empowers mutant p53. Breast Cancer Res. 2011, 13, 315. [Google Scholar] [CrossRef]
  13. Lin, C.H.; Li, H.Y.; Lee, Y.C.; Calkins, M.J.; Lee, K.H.; Yang, C.N.; Lu, P.J. Landscape of Pin1 in the cell cycle. Exp. Biol. Med. 2015, 240, 403–408. [Google Scholar] [CrossRef] [PubMed]
  14. Esnault, S.; Shen, Z.J.; Malter, J.S. Pinning down signaling in the immune system: The role of the peptidyl-prolyl isomerase Pin1 in immune cell function. Crit. Rev. Immunol. 2008, 28, 45–60. [Google Scholar] [CrossRef] [PubMed]
  15. Nakatsu, Y.; Matsunaga, Y.; Yamamotoya, T.; Ueda, K.; Inoue, Y.; Mori, K.; Sakoda, H.; Fujishiro, M.; Ono, H.; Kushiyama, A.; et al. Physiological and Pathogenic Roles of Prolyl Isomerase Pin1 in Metabolic Regulations via Multiple Signal Transduction Pathway Modulations. Int. J. Mol. Sci. 2016, 17, 1495. [Google Scholar] [CrossRef] [PubMed]
  16. Pu, W.; Zheng, Y.; Peng, Y. Prolyl Isomerase Pin1 in Human Cancer: Function, Mechanism, and Significance. Front. Cell Dev. Biol. 2020, 8, 168. [Google Scholar] [CrossRef] [PubMed]
  17. Rustighi, A.; Zannini, A.; Tiberi, L.; Sommaggio, R.; Piazza, S.; Sorrentino, G.; Nuzzo, S.; Tuscano, A.; Eterno, V.; Benvenuti, F.; et al. Prolyl-isomerase Pin1 controls normal and cancer stem cells of the breast. EMBO Mol. Med. 2014, 6, 99–119. [Google Scholar] [CrossRef] [PubMed]
  18. Langer, E.M.; English, I.A.; Kresse, K.M.; MacPherson, K.; Allen-Petersen, B.L.; Daniel, C.J.; Adey, A.; Sears, R.C. Abstract LT012: The prolyl isomerase PIN1 plays a critical role in fibroblast plasticity to impact pancreatic cancer. Cancer Res. 2021, 81, LT012. [Google Scholar] [CrossRef]
  19. Caligiuri, I.; Vincenzo, C.; Asano, T.; Kumar, V.; Rizzolio, F. The metabolic crosstalk between PIN1 and the tumour microenvironment. Semin. Cancer Biol. 2023, 91, 143–157. [Google Scholar] [CrossRef]
  20. Nakatsu, Y.; Matsunaga, Y.; Yamamotoya, T.; Ueda, K.; Inoue, M.K.; Mizuno, Y.; Nakanishi, M.; Sano, T.; Yamawaki, Y.; Kushiyama, A.; et al. Prolyl isomerase Pin1 suppresses thermogenic programs in adipocytes by promoting degradation of transcriptional co-activator PRDM16. Cell Rep. 2019, 26, 3221–3230.e3. [Google Scholar] [CrossRef]
  21. Ubersax, J.A.; Ferrell Jr, J.E. Mechanisms of specificity in protein phosphorylation. Nat. Rev. Mol. Cell Biol. 2007, 8, 530–541. [Google Scholar] [CrossRef]
  22. Ryo, A.; Nakamura, M.; Wulf, G.; Liou, Y.C.; Lu, K.P. Pin1 regulates turnover and subcellular localization of beta-catenin by inhibiting its interaction with APC. Nat. Cell Biol. 2001, 3, 793–801. [Google Scholar] [CrossRef] [PubMed]
  23. Min, S.H.; Lau, A.W.; Lee, T.H.; Inuzuka, H.; Wei, S.; Huang, P.; Shaik, S.; Lee, D.Y.; Finn, G.; Balastik, M.; et al. Negative regulation of the stability and tumor suppressor function of Fbw7 by the Pin1 prolyl isomerase. Mol. Cell 2012, 46, 771–783. [Google Scholar] [CrossRef] [PubMed]
  24. Gregory, M.A.; Qi, Y.; Hann, S.R. Phosphorylation by glycogen synthase kinase-3 controls c-myc proteolysis and subnuclear localization. J. Biol. Chem. 2003, 278, 51606–51612. [Google Scholar] [CrossRef] [PubMed]
  25. Welcker, M.; Orian, A.; Jin, J.; Grim, J.A.; Harper, J.W.; Eisenman, R.N.; Clurman, B.E. The Fbw7 tumor suppressor regulates glycogen synthase kinase 3 phosphorylation-dependent c-Myc protein degradation. Proc. Natl. Acad. Sci. USA 2004, 101, 9085–9090. [Google Scholar] [CrossRef] [PubMed]
  26. Farrell, A.S.; Pelz, C.; Wang, X.; Daniel, C.J.; Wang, Z.; Su, Y.; Janghorban, M.; Zhang, X.; Morgan, C.; Impey, S.; et al. Pin1 regulates the dynamics of c-Myc DNA binding to facilitate target gene regulation and oncogenesis. Mol. Cell. Biol. 2013, 33, 2930–2949. [Google Scholar] [CrossRef] [PubMed]
  27. Liou, Y.-C.; Ryo, A.; Huang, H.-K.; Lu, P.-J.; Bronson, R.; Fujimori, F.; Uchida, T.; Hunter, T.; Lu, K.P. Loss of Pin1 function in the mouse causes phenotypes resembling cyclin D1-null phenotypes. Proc. Natl. Acad. Sci. USA 2002, 99, 1335–1340. [Google Scholar] [CrossRef]
  28. Wulf, G.M.; Liou, Y.C.; Ryo, A.; Lee, S.W.; Lu, K.P. Role of Pin1 in the regulation of p53 stability and p21 transactivation, and cell cycle checkpoints in response to DNA damage. J. Biol. Chem. 2002, 277, 47976–47979. [Google Scholar] [CrossRef] [PubMed]
  29. Zheng, H.; You, H.; Zhou, X.Z.; Murray, S.A.; Uchida, T.; Wulf, G.; Gu, L.; Tang, X.; Lu, K.P.; Xiao, Z.-X.J. The prolyl isomerase Pin1 is a regulator of p53 in genotoxic response. Nature 2002, 419, 849–853. [Google Scholar] [CrossRef]
  30. Zacchi, P.; Gostissa, M.; Uchida, T.; Salvagno, C.; Avolio, F.; Volinia, S.; Ronai, Z.; Blandino, G.; Schneider, C.; Del Sal, G. The prolyl isomerase Pin1 reveals a mechanism to control p53 functions after genotoxic insults. Nature 2002, 419, 853–857. [Google Scholar] [CrossRef]
  31. Basu, A.; Das, M.; Qanungo, S.; Fan, X.-J.; DuBois, G.; Haldar, S. Proteasomal degradation of human peptidyl prolyl isomerase pin1-pointing phospho Bcl2 toward dephosphorylation. Neoplasia 2002, 4, 218–227. [Google Scholar] [CrossRef] [PubMed]
  32. Mantovani, F.; Piazza, S.; Gostissa, M.; Strano, S.; Zacchi, P.; Mantovani, R.; Blandino, G.; Del Sal, G. Pin1 links the activities of c-Abl and p300 in regulating p73 function. Mol. Cell 2004, 14, 625–636. [Google Scholar] [CrossRef] [PubMed]
  33. Becker, E.B.; Bonni, A. Pin1 mediates neural-specific activation of the mitochondrial apoptotic machinery. Neuron 2006, 49, 655–662. [Google Scholar] [CrossRef] [PubMed]
  34. Bernis, C.; Vigneron, S.; Burgess, A.; Labbé, J.C.; Fesquet, D.; Castro, A.; Lorca, T. Pin1 stabilizes Emi1 during G2 phase by preventing its association with SCF(betatrcp). EMBO Rep. 2007, 8, 91–98. [Google Scholar] [CrossRef] [PubMed]
  35. Pang, R.; Lee, T.K.; Poon, R.T.; Fan, S.T.; Wong, K.B.; Kwong, Y.; Tse, E. Pin1 interacts with a specific serine-proline motif of hepatitis B virus X-protein to enhance hepatocarcinogenesis. Gastroenterology 2007, 132, 1088–1103. [Google Scholar] [CrossRef] [PubMed]
  36. Lam, P.B.; Burga, L.N.; Wu, B.P.; Hofstatter, E.W.; Lu, K.P.; Wulf, G.M. Prolyl isomerase Pin1 is highly expressed in Her2-positive breast cancer and regulates erbB2 protein stability. Mol. Cancer 2008, 7, 91. [Google Scholar] [CrossRef]
  37. Khanal, P.; Namgoong, G.M.; Kang, B.S.; Woo, E.-R.; Choi, H.S. The prolyl isomerase Pin1 enhances HER-2 expression and cellular transformation via its interaction with mitogen-activated protein kinase/extracellular signal-regulated kinase kinase 1. Mol. Cancer Ther. 2010, 9, 606–616. [Google Scholar] [CrossRef]
  38. Ding, Q.; Huo, L.; Yang, J.Y.; Xia, W.; Wei, Y.; Liao, Y.; Chang, C.J.; Yang, Y.; Lai, C.C.; Lee, D.F.; et al. Down-regulation of myeloid cell leukemia-1 through inhibiting Erk/Pin 1 pathway by sorafenib facilitates chemosensitization in breast cancer. Cancer Res. 2008, 68, 6109–6117. [Google Scholar] [CrossRef] [PubMed]
  39. Rustighi, A.; Tiberi, L.; Soldano, A.; Napoli, M.; Nuciforo, P.; Rosato, A.; Kaplan, F.; Capobianco, A.; Pece, S.; Di Fiore, P.P.; et al. The prolyl-isomerase Pin1 is a Notch1 target that enhances Notch1 activation in cancer. Nat. Cell Biol. 2009, 11, 133–142. [Google Scholar] [CrossRef]
  40. Liao, Y.; Wei, Y.; Zhou, X.; Yang, J.Y.; Dai, C.; Chen, Y.J.; Agarwal, N.K.; Sarbassov, D.; Shi, D.; Yu, D.; et al. Peptidyl-prolyl cis/trans isomerase Pin1 is critical for the regulation of PKB/Akt stability and activation phosphorylation. Oncogene 2009, 28, 2436–2445. [Google Scholar] [CrossRef]
  41. van der Horst, A.; Khanna, K.K. The peptidyl-prolyl isomerase Pin1 regulates cytokinesis through Cep55. Cancer Res. 2009, 69, 6651–6659. [Google Scholar] [CrossRef]
  42. Cheng, C.W.; Leong, K.W.; Ng, Y.M.; Kwong, Y.L.; Tse, E. The peptidyl-prolyl isomerase PIN1 relieves cyclin-dependent kinase 2 (CDK2) inhibition by the CDK inhibitor p27. J. Biol. Chem. 2017, 292, 21431–21441. [Google Scholar] [CrossRef] [PubMed]
  43. Zhou, W.; Yang, Q.; Low, C.B.; Karthik, B.C.; Wang, Y.; Ryo, A.; Yao, S.Q.; Yang, D.; Liou, Y.-C. Pin1 catalyzes conformational changes of Thr-187 in p27Kip1 and mediates its stability through a polyubiquitination process. J. Biol. Chem. 2009, 284, 23980–23988. [Google Scholar] [CrossRef] [PubMed]
  44. Fan, G.; Fan, Y.; Gupta, N.; Matsuura, I.; Liu, F.; Zhou, X.Z.; Lu, K.P.; Gélinas, C. Peptidyl-prolyl isomerase Pin1 markedly enhances the oncogenic activity of the rel proteins in the nuclear factor-kappaB family. Cancer Res. 2009, 69, 4589–4597. [Google Scholar] [CrossRef] [PubMed]
  45. Baik, S.H.; Fane, M.; Park, J.H.; Cheng, Y.L.; Yang-Wei Fann, D.; Yun, U.J.; Choi, Y.; Park, J.S.; Chai, B.H.; Park, J.S.; et al. Pin1 promotes neuronal death in stroke by stabilizing Notch intracellular domain. Ann. Neurol. 2015, 77, 504–516. [Google Scholar] [CrossRef] [PubMed]
  46. Jeong, S.J.; Ryo, A.; Yamamoto, N. The prolyl isomerase Pin1 stabilizes the human T-cell leukemia virus type 1 (HTLV-1) Tax oncoprotein and promotes malignant transformation. Biochem. Biophys. Res. Commun. 2009, 381, 294–299. [Google Scholar] [CrossRef] [PubMed]
  47. Peloponese, J.M., Jr.; Yasunaga, J.; Kinjo, T.; Watashi, K.; Jeang, K.T. Peptidylproline cis-trans-isomerase Pin1 interacts with human T-cell leukemia virus type 1 tax and modulates its activation of NF-kappaB. J. Virol. 2009, 83, 3238–3248. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, T.; Schneider, R.A.; Shah, V.; Huang, Y.; Likhotvorik, R.I.; Keshvara, L.; Hoyt, D.G. Protein Never in Mitosis Gene A Interacting-1 regulates calpain activity and the degradation of cyclooxygenase-2 in endothelial cells. J. Inflamm. 2009, 6, 20. [Google Scholar] [CrossRef]
  49. Jeong, H.G.; Pokharel, Y.R.; Lim, S.C.; Hwang, Y.P.; Han, E.H.; Yoon, J.-H.; Ahn, S.-G.; Lee, K.Y.; Kang, K.W. Novel role of Pin1 induction in type II collagen-mediated rheumatoid arthritis. J. Immunol. 2009, 183, 6689–6697. [Google Scholar] [CrossRef]
  50. Fujimoto, Y.; Shiraki, T.; Horiuchi, Y.; Waku, T.; Shigenaga, A.; Otaka, A.; Ikura, T.; Igarashi, K.; Aimoto, S.; Tate, S.I.; et al. Proline cis/trans-isomerase Pin1 regulates peroxisome proliferator-activated receptor gamma activity through the direct binding to the activation function-1 domain. J. Biol. Chem. 2010, 285, 3126–3132. [Google Scholar] [CrossRef]
  51. Moretto-Zita, M.; Jin, H.; Shen, Z.; Zhao, T.; Briggs, S.P.; Xu, Y. Phosphorylation stabilizes Nanog by promoting its interaction with Pin1. Proc. Natl. Acad. Sci. USA 2010, 107, 13312–13317. [Google Scholar] [CrossRef] [PubMed]
  52. Manganaro, L.; Lusic, M.; Gutierrez, M.I.; Cereseto, A.; Del Sal, G.; Giacca, M. Concerted action of cellular JNK and Pin1 restricts HIV-1 genome integration to activated CD4+ T lymphocytes. Nat. Med. 2010, 16, 329–333. [Google Scholar] [CrossRef]
  53. Nishi, M.; Akutsu, H.; Masui, S.; Kondo, A.; Nagashima, Y.; Kimura, H.; Perrem, K.; Shigeri, Y.; Toyoda, M.; Okayama, A.; et al. A distinct role for Pin1 in the induction and maintenance of pluripotency. J. Biol. Chem. 2011, 286, 11593–11603. [Google Scholar] [CrossRef] [PubMed]
  54. Girardini, J.E.; Napoli, M.; Piazza, S.; Rustighi, A.; Marotta, C.; Radaelli, E.; Capaci, V.; Jordan, L.; Quinlan, P.; Thompson, A.; et al. A Pin1/mutant p53 axis promotes aggressiveness in breast cancer. Cancer Cell 2011, 20, 79–91. [Google Scholar] [CrossRef] [PubMed]
  55. Marcucci, R.; Brindle, J.; Paro, S.; Casadio, A.; Hempel, S.; Morrice, N.; Bisso, A.; Keegan, L.P.; Del Sal, G.; O’connell, M.A. Pin1 and WWP2 regulate GluR2 Q/R site RNA editing by ADAR2 with opposing effects. EMBO J. 2011, 30, 4211–4222. [Google Scholar] [CrossRef]
  56. Kato, H.; Naito, M.; Saito, T.; Hideyama, T.; Suzuki, Y.; Kimura, T.; Kwak, S.; Aizawa, H. Prolyl Isomerase Pin1 Expression in the Spinal Motor Neurons of Patients With Sporadic Amyotrophic Lateral Sclerosis. J. Clin. Neurol. 2022, 18, 463–469. [Google Scholar] [CrossRef] [PubMed]
  57. Rajbhandari, P.; Finn, G.; Solodin, N.M.; Singarapu, K.K.; Sahu, S.C.; Markley, J.L.; Kadunc, K.J.; Ellison-Zelski, S.J.; Kariagina, A.; Haslam, S.Z.; et al. Regulation of estrogen receptor α N-terminus conformation and function by peptidyl prolyl isomerase Pin1. Mol. Cell Biol. 2012, 32, 445–457. [Google Scholar] [CrossRef]
  58. Lucchetti, C.; Caligiuri, I.; Toffoli, G.; Giordano, A.; Rizzolio, F. The prolyl isomerase Pin1 acts synergistically with CDK2 to regulate the basal activity of estrogen receptor α in breast cancer. PLoS ONE 2013, 8, e55355. [Google Scholar] [CrossRef]
  59. Bitomsky, N.; Conrad, E.; Moritz, C.; Polonio-Vallon, T.; Sombroek, D.; Schultheiss, K.; Glas, C.; Greiner, V.; Herbel, C.; Mantovani, F.; et al. Autophosphorylation and Pin1 binding coordinate DNA damage-induced HIPK2 activation and cell death. Proc. Natl. Acad. Sci. USA 2013, 110, E4203–E4212. [Google Scholar] [CrossRef]
  60. Li, C.; Chang, D.L.; Yang, Z.; Qi, J.; Liu, R.; He, H.; Li, D.; Xiao, Z.X. Pin1 modulates p63α protein stability in regulation of cell survival, proliferation and tumor formation. Cell Death Dis. 2013, 4, e943. [Google Scholar] [CrossRef]
  61. Steger, M.; Murina, O.; Hühn, D.; Ferretti, L.P.; Walser, R.; Hänggi, K.; Lafranchi, L.; Neugebauer, C.; Paliwal, S.; Janscak, P.; et al. Prolyl isomerase PIN1 regulates DNA double-strand break repair by counteracting DNA end resection. Mol. Cell 2013, 50, 333–343. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, H.-C.; Chuang, J.-Y.; Jeng, W.-Y.; Liu, C.-I.; Wang, A.H.-J.; Lu, P.-J.; Chang, W.-C.; Hung, J.-J. Pin1-mediated Sp1 phosphorylation by CDK1 increases Sp1 stability and decreases its DNA-binding activity during mitosis. Nucleic Acids Res. 2014, 42, 13573–13587. [Google Scholar] [CrossRef] [PubMed]
  63. Lonati, E.; Brambilla, A.; Milani, C.; Masserini, M.; Palestini, P.; Bulbarelli, A. Pin1, a new player in the fate of HIF-1α degradation: An hypothetical mechanism inside vascular damage as Alzheimer’s disease risk factor. Front. Cell. Neurosci. 2014, 8, 1. [Google Scholar] [CrossRef]
  64. Jalouli, M.; Déry, M.A.; Lafleur, V.N.; Lamalice, L.; Zhou, X.Z.; Lu, K.P.; Richard, D.E. The prolyl isomerase Pin1 regulates hypoxia-inducible transcription factor (HIF) activity. Cell. Signal. 2014, 26, 1649–1656. [Google Scholar] [CrossRef]
  65. Han, H.-j.; Saeidi, S.; Kim, S.-J.; Piao, J.-Y.; Lim, S.; Guillen-Quispe, Y.N.; Choi, B.Y.; Surh, Y.-J. Alternative regulation of HIF-1α stability through Phosphorylation on Ser451. Biochem. Biophys. Res. Commun. 2021, 545, 150–156. [Google Scholar] [CrossRef] [PubMed]
  66. Rajbhandari, P.; Schalper, K.A.; Solodin, N.M.; Ellison-Zelski, S.J.; Ping Lu, K.; Rimm, D.L.; Alarid, E.T. Pin1 modulates ERα levels in breast cancer through inhibition of phosphorylation-dependent ubiquitination and degradation. Oncogene 2014, 33, 1438–1447. [Google Scholar] [CrossRef]
  67. Hellmuth, S.; Rata, S.; Brown, A.; Heidmann, S.; Novak, B.; Stemmann, O. Human chromosome segregation involves multi-layered regulation of separase by the peptidyl-prolyl-isomerase Pin1. Mol. Cell 2015, 58, 495–506. [Google Scholar] [CrossRef] [PubMed]
  68. Wei, S.; Kozono, S.; Kats, L.; Nechama, M.; Li, W.; Guarnerio, J.; Luo, M.; You, M.H.; Yao, Y.; Kondo, A.; et al. Active Pin1 is a key target of all-trans retinoic acid in acute promyelocytic leukemia and breast cancer. Nat. Med. 2015, 21, 457–466. [Google Scholar] [CrossRef]
  69. Balastik, M.; Zhou, X.Z.; Alberich-Jorda, M.; Weissova, R.; Žiak, J.; Pazyra-Murphy, M.F.; Cosker, K.E.; Machonova, O.; Kozmikova, I.; Chen, C.H.; et al. Prolyl Isomerase Pin1 Regulates Axon Guidance by Stabilizing CRMP2A Selectively in Distal Axons. Cell Rep. 2015, 13, 812–828. [Google Scholar] [CrossRef]
  70. Kang, S.W.; Lee, E.; Cho, E.; Seo, J.H.; Ko, H.W.; Kim, E.Y. Drosophila peptidyl-prolyl isomerase Pin1 modulates circadian rhythms via regulating levels of PERIOD. Biochem. Biophys. Res. Commun. 2015, 463, 235–240. [Google Scholar] [CrossRef]
  71. Hu, X.; Dong, S.H.; Chen, J.; Zhou, X.Z.; Chen, R.; Nair, S.; Lu, K.P.; Chen, L.F. Prolyl isomerase PIN1 regulates the stability, transcriptional activity and oncogenic potential of BRD4. Oncogene 2017, 36, 5177–5188. [Google Scholar] [CrossRef] [PubMed]
  72. Ettelaie, C.; Collier, M.E.W.; Featherby, S.; Greenman, J.; Maraveyas, A. Peptidyl-prolyl isomerase 1 (Pin1) preserves the phosphorylation state of tissue factor and prolongs its release within microvesicles. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2018, 1865, 12–24. [Google Scholar] [CrossRef] [PubMed]
  73. Kurakula, K.; Koenis, D.S.; Herzik, M.A., Jr.; Liu, Y.; Craft, J.W., Jr.; van Loenen, P.B.; Vos, M.; Tran, M.K.; Versteeg, H.H.; Goumans, M.T.H.; et al. Structural and cellular mechanisms of peptidyl-prolyl isomerase Pin1-mediated enhancement of Tissue Factor gene expression, protein half-life, and pro-coagulant activity. Haematologica 2018, 103, 1073–1082. [Google Scholar] [CrossRef] [PubMed]
  74. Sang, Y.; Li, Y.; Zhang, Y.; Alvarez, A.A.; Yu, B.; Zhang, W.; Hu, B.; Cheng, S.-Y.; Feng, H. CDK5-dependent phosphorylation and nuclear translocation of TRIM59 promotes macroH2A1 ubiquitination and tumorigenicity. Nat. Commun. 2019, 10, 4013. [Google Scholar] [CrossRef]
  75. Kim, G.; Bhattarai, P.Y.; Lim, S.C.; Kim, J.Y.; Choi, H.S. PIN1 facilitates ubiquitin-mediated degradation of serine/threonine kinase 3 and promotes melanoma development via TAZ activation. Cancer Lett. 2021, 499, 164–174. [Google Scholar] [CrossRef] [PubMed]
  76. Khanal, P.; Yeung, B.; Zhao, Y.; Yang, X. Identification of Prolyl isomerase Pin1 as a novel positive regulator of YAP/TAZ in breast cancer cells. Sci. Rep. 2019, 9, 6394. [Google Scholar] [CrossRef] [PubMed]
  77. Ueda, K.; Nakatsu, Y.; Yamamotoya, T.; Ono, H.; Inoue, Y.; Inoue, M.-K.; Mizuno, Y.; Matsunaga, Y.; Kushiyama, A.; Sakoda, H.; et al. Prolyl isomerase Pin1 binds to and stabilizes acetyl CoA carboxylase 1 protein, thereby supporting cancer cell proliferation. Oncotarget 2019, 10, 1637–1648. [Google Scholar] [CrossRef] [PubMed]
  78. Daza-Martin, M.; Starowicz, K.; Jamshad, M.; Tye, S.; Ronson, G.E.; MacKay, H.L.; Chauhan, A.S.; Walker, A.K.; Stone, H.R.; Beesley, J.F.J.; et al. Isomerization of BRCA1–BARD1 promotes replication fork protection. Nature 2019, 571, 521–527. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, J.; Chan, B.; Tong, M.; Paung, Y.; Jo, U.; Martin, D.; Seeliger, M.; Haley, J.; Kim, H. Prolyl isomerization of FAAP20 catalyzed by PIN1 regulates the Fanconi anemia pathway. PLoS Genet. 2019, 15, e1007983. [Google Scholar] [CrossRef]
  80. Luo, M.L.; Zheng, F.; Chen, W.; Liang, Z.M.; Chandramouly, G.; Tan, J.; Willis, N.A.; Chen, C.H.; Taveira, M.O.; Zhou, X.Z.; et al. Inactivation of the Prolyl Isomerase Pin1 Sensitizes BRCA1-Proficient Breast Cancer to PARP Inhibition. Cancer Res. 2020, 80, 3033–3045. [Google Scholar] [CrossRef]
  81. Miyakawa, K.; Matsunaga, S.; Khatun, H.; Yamaoka, Y.; Watashi, K.; Sugiyama, M.; Kimura, H.; Wakita, T.; Ryo, A. Prolyl Isomerase Pin1 Regulates the Stability of Hepatitis B Virus Core Protein. Front. Cell Dev. Biol. 2020, 8, 26. [Google Scholar]
  82. Napoletano, F.; Bravo, G.F.; Voto, I.A.P.; Santin, A.; Celora, L.; Campaner, E.; Dezi, C.; Bertossi, A.; Valentino, E.; Santorsola, M.; et al. The prolyl-isomerase PIN1 is essential for nuclear Lamin-B structure and function and protects heterochromatin under mechanical stress. Cell Rep. 2021, 36, 109694. [Google Scholar] [CrossRef] [PubMed]
  83. Saeidi, S.; Kim, S.J.; Guillen-Quispe, Y.N.; Jagadeesh, A.S.V.; Han, H.J.; Kim, S.H.; Zhong, X.; Piao, J.Y.; Kim, S.J.; Jeong, J.; et al. Peptidyl-prolyl cis-trans isomerase NIMA-interacting 1 directly binds and stabilizes Nrf2 in breast cancer. FASEB J. 2022, 36, e22068. [Google Scholar] [CrossRef] [PubMed]
  84. Guillen-Quispe, Y.N.; Kim, S.J.; Saeidi, S.; Zhou, T.; Zheng, J.; Kim, S.H.; Fang, X.; Chelakkot, C.; Rios-Castillo, M.E.; Shin, Y.K.; et al. Oxygen-independent stabilization of HIF-2α in breast cancer through direct interaction with peptidyl-prolyl cis-trans isomerase NIMA-interacting 1. Free Radic. Biol. Med. 2023, 207, 296–307. [Google Scholar] [CrossRef] [PubMed]
  85. Tang, J.; Li, J.; Lian, J.; Huang, Y.; Zhang, Y.; Lu, Y.; Zhong, G.; Wang, Y.; Zhang, Z.; Bai, X.; et al. CDK2-activated TRIM32 phosphorylation and nuclear translocation promotes radioresistance in triple-negative breast cancer. J. Adv. Res. 2023. [Google Scholar] [CrossRef]
  86. Lu, P.J.; Wulf, G.; Zhou, X.Z.; Lu, K.P. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature 1999, 399, 784–788. [Google Scholar] [CrossRef]
  87. Liou, Y.C.; Sun, A.; Ryo, A.; Zhou, X.Z.; Yu, Z.X.; Huang, H.K.; Uchida, T.; Bronson, R.; Bing, G.; Li, X.; et al. Role of the prolyl isomerase Pin1 in protecting against age-dependent neurodegeneration. Nature 2003, 424, 556–561. [Google Scholar] [CrossRef]
  88. Zhou, X.Z.; Kops, O.; Werner, A.; Lu, P.J.; Shen, M.; Stoller, G.; Küllertz, G.; Stark, M.; Fischer, G.; Lu, K.P. Pin1-dependent prolyl isomerization regulates dephosphorylation of Cdc25C and tau proteins. Mol. Cell 2000, 6, 873–883. [Google Scholar] [CrossRef]
  89. Hsu, T.; McRackan, D.; Vincent, T.S.; de Couet, H.G. Drosophila Pin1 prolyl isomerase Dodo is a MAP kinase signal responder during oogenesis. Nat. Cell Biol. 2001, 3, 538–543. [Google Scholar] [CrossRef]
  90. Brondani, V.; Schefer, Q.; Hamy, F.; Klimkait, T. The peptidyl-prolyl isomerase Pin1 regulates phospho-Ser77 retinoic acid receptor alpha stability. Biochem. Biophys. Res. Commun. 2005, 328, 6–13. [Google Scholar] [CrossRef]
  91. Yu, L.; Mohamed, A.J.; Vargas, L.; Berglöf, A.; Finn, G.; Lu, K.P.; Smith, C.I.E. Regulation of Bruton tyrosine kinase by the peptidylprolyl isomerase Pin1. J. Biol. Chem. 2006, 281, 18201–18207. [Google Scholar] [CrossRef] [PubMed]
  92. Saitoh, T.; Tun-Kyi, A.; Ryo, A.; Yamamoto, M.; Finn, G.; Fujita, T.; Akira, S.; Yamamoto, N.; Lu, K.P.; Yamaoka, S. Negative regulation of interferon-regulatory factor 3-dependent innate antiviral response by the prolyl isomerase Pin1. Nat. Immunol. 2006, 7, 598–605. [Google Scholar] [CrossRef] [PubMed]
  93. Yeh, E.S.; Lew, B.O.; Means, A.R. The loss of PIN1 deregulates cyclin E and sensitizes mouse embryo fibroblasts to genomic instability. J. Biol. Chem. 2006, 281, 241–251. [Google Scholar] [CrossRef] [PubMed]
  94. van Drogen, F.; Sangfelt, O.; Malyukova, A.; Matskova, L.; Yeh, E.; Means, A.R.; Reed, S.I. Ubiquitylation of cyclin E requires the sequential function of SCF complexes containing distinct hCdc4 isoforms. Mol. Cell 2006, 23, 37–48. [Google Scholar] [CrossRef] [PubMed]
  95. Bhaskaran, N.; van Drogen, F.; Ng, H.-F.; Kumar, R.; Ekholm-Reed, S.; Peter, M.; Sangfelt, O.; Reed, S.I. Fbw7α and Fbw7γ collaborate to shuttle cyclin E1 into the nucleolus for multiubiquitylation. Mol. Cell Biol. 2013, 33, 85–97. [Google Scholar] [CrossRef] [PubMed]
  96. Sangfelt, O.; Cepeda, D.; Malyukova, A.; van Drogen, F.; Reed, S.I. Both SCF(Cdc4alpha) and SCF(Cdc4gamma) are required for cyclin E turnover in cell lines that do not overexpress cyclin E. Cell Cycle 2008, 7, 1075–1082. [Google Scholar] [CrossRef] [PubMed]
  97. Ryo, A.; Hirai, A.; Nishi, M.; Liou, Y.C.; Perrem, K.; Lin, S.C.; Hirano, H.; Lee, S.W.; Aoki, I. A suppressive role of the prolyl isomerase Pin1 in cellular apoptosis mediated by the death-associated protein Daxx. J. Biol. Chem. 2007, 282, 36671–36681. [Google Scholar] [CrossRef] [PubMed]
  98. Stanya, K.J.; Liu, Y.; Means, A.R.; Kao, H.-Y. Cdk2 and Pin1 negatively regulate the transcriptional corepressor SMRT. J. Cell Biol. 2008, 183, 49–61. [Google Scholar] [CrossRef]
  99. Brenkman, A.B.; De Keizer, P.L.; Van Den Broek, N.J.; van der Groep, P.; Van Diest, P.J.; Van Der Horst, A.; Smits, A.M.; Burgering, B.M. The peptidyl-isomerase Pin1 regulates p27kip1 expression through inhibition of Forkhead box O tumor suppressors. Cancer Res. 2008, 68, 7597–7605. [Google Scholar] [CrossRef]
  100. Yuan, W.C.; Lee, Y.R.; Huang, S.F.; Lin, Y.M.; Chen, T.Y.; Chung, H.C.; Tsai, C.H.; Chen, H.Y.; Chiang, C.T.; Lai, C.K.; et al. A Cullin3-KLHL20 Ubiquitin ligase-dependent pathway targets PML to potentiate HIF-1 signaling and prostate cancer progression. Cancer Cell 2011, 20, 214–228. [Google Scholar] [CrossRef]
  101. Reineke, E.L.; Lam, M.; Liu, Q.; Liu, Y.; Stanya, K.J.; Chang, K.-S.; Means, A.R.; Kao, H.-Y. Degradation of the tumor suppressor PML by Pin1 contributes to the cancer phenotype of breast cancer MDA-MB-231 cells. Mol. Cell Biol. 2008, 28, 997–1006. [Google Scholar] [CrossRef] [PubMed]
  102. Watashi, K.; Khan, M.; Yedavalli, V.R.K.; Yeung, M.L.; Strebel, K.; Jeang, K.-T. Human immunodeficiency virus type 1 replication and regulation of APOBEC3G by peptidyl prolyl isomerase Pin1. J. Virol. 2008, 82, 9928–9936. [Google Scholar] [CrossRef] [PubMed]
  103. Nakano, A.; Koinuma, D.; Miyazawa, K.; Uchida, T.; Saitoh, M.; Kawabata, M.; Hanai, J.I.; Akiyama, H.; Abe, M.; Miyazono, K.; et al. Pin1 down-regulates transforming growth factor-beta (TGF-beta) signaling by inducing degradation of Smad proteins. J. Biol. Chem. 2009, 284, 6109–6115. [Google Scholar] [CrossRef] [PubMed]
  104. Penela, P.; Rivas, V.; Salcedo, A.; Mayor, F., Jr. G protein-coupled receptor kinase 2 (GRK2) modulation and cell cycle progression. Proc. Natl. Acad. Sci. USA 2010, 107, 1118–1123. [Google Scholar] [CrossRef] [PubMed]
  105. Khanal, P.; Yun, H.J.; Lim, S.C.; Ahn, S.G.; Yoon, H.E.; Kang, K.W.; Hong, R.; Choi, H.S. Proyl isomerase Pin1 facilitates ubiquitin-mediated degradation of cyclin-dependent kinase 10 to induce tamoxifen resistance in breast cancer cells. Oncogene 2012, 31, 3845–3856. [Google Scholar] [CrossRef] [PubMed]
  106. Marsolier, J.; Perichon, M.; DeBarry, J.D.; Villoutreix, B.O.; Chluba, J.; Lopez, T.; Garrido, C.; Zhou, X.Z.; Lu, K.P.; Fritsch, L.; et al. Theileria parasites secrete a prolyl isomerase to maintain host leukocyte transformation. Nature 2015, 520, 378–382. [Google Scholar] [CrossRef] [PubMed]
  107. Abrahamsen, H.; O’Neill, A.K.; Kannan, N.; Kruse, N.; Taylor, S.S.; Jennings, P.A.; Newton, A.C. Peptidyl-prolyl isomerase Pin1 controls down-regulation of conventional protein kinase C isozymes. J. Biol. Chem. 2012, 287, 13262–13278. [Google Scholar] [CrossRef]
  108. Lee, Y.C.; Que, J.; Chen, Y.C.; Lin, J.T.; Liou, Y.C.; Liao, P.C.; Liu, Y.P.; Lee, K.H.; Lin, L.C.; Hsiao, M.; et al. Pin1 acts as a negative regulator of the G2/M transition by interacting with the Aurora-A-Bora complex. J. Cell Sci. 2013, 126 Pt 21, 4862–4872. [Google Scholar] [CrossRef]
  109. Khanal, P.; Kim, G.; Lim, S.C.; Yun, H.J.; Lee, K.Y.; Choi, H.K.; Choi, H.S. Prolyl isomerase Pin1 negatively regulates the stability of SUV39H1 to promote tumorigenesis in breast cancer. FASEB J. 2013, 27, 4606–4618. [Google Scholar] [CrossRef]
  110. Nicole Tsang, Y.H.; Wu, X.W.; Lim, J.S.; Wee Ong, C.; Salto-Tellez, M.; Ito, K.; Ito, Y.; Chen, L.F. Prolyl isomerase Pin1 downregulates tumor suppressor RUNX3 in breast cancer. Oncogene 2013, 32, 1488–1496. [Google Scholar] [CrossRef]
  111. Hwang, Y.-C.; Yang, C.-H.; Lin, C.-H.; Ch’Ang, H.-J.; Chang, V.H.; Yu, W.C. Destabilization of KLF10, a tumor suppressor, relies on thr93 phosphorylation and isomerase association. Biochim. Biophys. Acta 2013, 1833, 3035–3045. [Google Scholar] [CrossRef] [PubMed]
  112. De Nicola, F.; Catena, V.; Rinaldo, C.; Bruno, T.; Iezzi, S.; Sorino, C.; Desantis, A.; Camerini, S.; Crescenzi, M.; Floridi, A.; et al. HIPK2 sustains apoptotic response by phosphorylating Che-1/AATF and promoting its degradation. Cell Death Dis. 2014, 5, e1414. [Google Scholar] [CrossRef] [PubMed]
  113. Nesti, E.; Corson, G.M.; McCleskey, M.; Oyer, J.A.; Mandel, G. C-terminal domain small phosphatase 1 and MAP kinase reciprocally control REST stability and neuronal differentiation. Proc. Natl. Acad. Sci. USA 2014, 111, E3929–E3936. [Google Scholar] [CrossRef] [PubMed]
  114. Nakamura, K.; Kosugi, I.; Lee, D.Y.; Hafner, A.; Sinclair, D.A.; Ryo, A.; Lu, K.P. Prolyl isomerase Pin1 regulates neuronal differentiation via β-catenin. Mol. Cell. Biol. 2012, 32, 2966–2978. [Google Scholar] [CrossRef] [PubMed]
  115. Carnemolla, A.; Michelazzi, S.; Agostoni, E. PIN1 Modulates Huntingtin Levels and Aggregate Accumulation: An In vitro Model. Front. Cell. Neurosci. 2017, 11, 121. [Google Scholar] [CrossRef] [PubMed]
  116. Csizmok, V.; Montecchio, M.; Lin, H.; Tyers, M.; Sunnerhagen, M.; Forman-Kay, J.D. Multivalent Interactions with Fbw7 and Pin1 Facilitate Recognition of c-Jun by the SCF(Fbw7) Ubiquitin Ligase. Structure 2018, 26, 28–39.e2. [Google Scholar] [CrossRef] [PubMed]
  117. Maarifi, G.; Smith, N.; Maillet, S.; Moncorgé, O.; Chamontin, C.; Edouard, J.; Sohm, F.; Blanchet, F.P.; Herbeuval, J.-P.; Lutfalla, G.; et al. TRIM8 is required for virus-induced IFN response in human plasmacytoid dendritic cells. Sci. Adv. 2019, 5, eaax3511. [Google Scholar] [CrossRef] [PubMed]
  118. Delgado, J.Y. An Alternative Pin1 Binding and Isomerization Site in the N-Terminus Domain of PSD-95. Front. Mol. Neurosci. 2020, 13, 31. [Google Scholar] [CrossRef] [PubMed]
  119. Nakatsu, Y.; Yamamotoya, T.; Okumura, M.; Ishii, T.; Kanamoto, M.; Naito, M.; Nakanishi, M.; Aoyama, S.; Matsunaga, Y.; Kushiyama, A.; et al. Prolyl isomerase Pin1 interacts with adipose triglyceride lipase and negatively controls both its expression and lipolysis. Metabolism 2021, 115, 154459. [Google Scholar] [CrossRef]
  120. Sun, Z.; Jiang, Q.; Gao, B.; Zhang, X.; Bu, L.; Wang, L.; Lin, Y.; Xie, W.; Li, J.; Guo, J. AKT Blocks SIK1-Mediated Repression of STAT3 to Promote Breast Tumorigenesis. Cancer Res. 2023, 83, 1264–1279. [Google Scholar] [CrossRef]
  121. Liou, Y.C.; Zhou, X.Z.; Lu, K.P. Prolyl isomerase Pin1 as a molecular switch to determine the fate of phosphoproteins. Trends Biochem. Sci. 2011, 36, 501–514. [Google Scholar] [CrossRef] [PubMed]
  122. Stewart, D.E.; Sarkar, A.; Wampler, J.E. Occurrence and role of cis peptide bonds in protein structures. J. Mol. Biol. 1990, 214, 253–260. [Google Scholar] [CrossRef]
  123. Pal, D.; Chakrabarti, P. Cis peptide bonds in proteins: Residues involved, their conformations, interactions and locations. J. Mol. Biol. 1999, 294, 271–288. [Google Scholar] [CrossRef]
  124. Schutkowski, M.; Bernhardt, A.; Zhou, X.Z.; Shen, M.; Reimer, U.; Rahfeld, J.-U.; Lu, K.P.; Fischer, G. Role of Phosphorylation in Determining the Backbone Dynamics of the Serine/Threonine-Proline Motif and Pin1 Substrate Recognition. Biochemistry 1998, 37, 5566–5575. [Google Scholar] [CrossRef] [PubMed]
  125. Hamelberg, D.; Shen, T.; McCammon, J.A. Phosphorylation Effects on cis/trans Isomerization and the Backbone Conformation of Serine−Proline Motifs:  Accelerated Molecular Dynamics Analysis. J. Am. Chem. Soc. 2005, 127, 1969–1974. [Google Scholar] [CrossRef] [PubMed]
  126. Ke, S.; Dang, F.; Wang, L.; Chen, J.Y.; Naik, M.T.; Thavamani, A.; Liu, Y.; Li, W.; Kim, N.; Naik, N.M.; et al. Reciprocal antagonism of PIN1 and APC/CCDH1 governs mitotic protein stability and cell cycle entry. Nat. Commun. 2023, 14, 3220. [Google Scholar] [CrossRef] [PubMed]
  127. Chae, U.; Lee, H.; Kim, B.; Jung, H.; Kim, B.M.; Lee, A.-H.; Lee, D.-S.; Min, S.-H. A negative feedback loop between XBP1 and Fbw7 regulates cancer development. Oncogenesis 2019, 8, 12. [Google Scholar] [CrossRef] [PubMed]
  128. Lutz, J.; Höllmüller, E.; Scheffner, M.; Marx, A.; Stengel, F. The Length of a Ubiquitin Chain: A General Factor for Selective Recognition by Ubiquitin-Binding Proteins. Angew. Chem. Int. Ed. 2020, 59, 12371–12375. [Google Scholar] [CrossRef] [PubMed]
  129. Siepe, D.; Jentsch, S. Prolyl isomerase Pin1 acts as a switch to control the degree of substrate ubiquitylation. Nat. Cell Biol. 2009, 11, 967–972. [Google Scholar] [CrossRef]
  130. Koegl, M.; Hoppe, T.; Schlenker, S.; Ulrich, H.D.; Mayer, T.U.; Jentsch, S. A novel ubiquitination factor, E4, is involved in multiubiquitin chain assembly. Cell 1999, 96, 635–644. [Google Scholar] [CrossRef]
  131. Hoppe, T. Multiubiquitylation by E4 enzymes: ‘One size’ doesn’t fit all. Trends Biochem. Sci. 2005, 30, 183–187. [Google Scholar] [CrossRef] [PubMed]
  132. Li, M.; Brooks, C.L.; Wu-Baer, F.; Chen, D.; Baer, R.; Gu, W. Mono- versus polyubiquitination: Differential control of p53 fate by Mdm2. Science 2003, 302, 1972–1975. [Google Scholar] [CrossRef] [PubMed]
  133. Chen, D.; Wang, L.; Lee, T.H. Post-translational Modifications of the Peptidyl-Prolyl Isomerase Pin1. Front. Cell Dev. Biol. 2020, 8, 129. [Google Scholar] [CrossRef] [PubMed]
  134. Bao, L.; Kimzey, A.; Sauter, G.; Sowadski, J.M.; Lu, K.P.; Wang, D.G. Prevalent overexpression of prolyl isomerase Pin1 in human cancers. Am. J. Pathol. 2004, 164, 1727–1737. [Google Scholar] [CrossRef] [PubMed]
  135. Ayala, G.; Wang, D.; Wulf, G.; Frolov, A.; Li, R.; Sowadski, J.; Wheeler, T.M.; Lu, K.P.; Bao, L. The prolyl isomerase Pin1 is a novel prognostic marker in human prostate cancer. Cancer Res. 2003, 63, 6244–6251. [Google Scholar] [PubMed]
  136. Tan, X.; Zhou, F.; Wan, J.; Hang, J.; Chen, Z.; Li, B.; Zhang, C.; Shao, K.; Jiang, P.; Shi, S.; et al. Pin1 expression contributes to lung cancer: Prognosis and carcinogenesis. Cancer Biol. Ther. 2010, 9, 111–119. [Google Scholar] [CrossRef] [PubMed]
  137. Ryo, A.; Liou, Y.C.; Wulf, G.; Nakamura, M.; Lee, S.W.; Lu, K.P. PIN1 is an E2F target gene essential for Neu/Ras-induced transformation of mammary epithelial cells. Mol. Cell. Biol. 2002, 22, 5281–5295. [Google Scholar] [CrossRef] [PubMed]
  138. Lu, P.J.; Zhou, X.Z.; Liou, Y.C.; Noel, J.P.; Lu, K.P. Critical role of WW domain phosphorylation in regulating phosphoserine binding activity and Pin1 function. J. Biol. Chem. 2002, 277, 2381–2384. [Google Scholar] [CrossRef]
  139. Zhou, X.Z.; Lu, K.P. The isomerase Pin1 controls numerous cancer-driving pathways and is a unique drug target. Nat. Rev. Cancer 2016, 16, 463–478. [Google Scholar] [CrossRef]
  140. Wulf, G.M.; Ryo, A.; Wulf, G.G.; Lee, S.W.; Niu, T.; Petkova, V.; Lu, K.P. Pin1 is overexpressed in breast cancer and cooperates with Ras signaling in increasing the transcriptional activity of c-Jun towards cyclin D1. EMBO J. 2001, 20, 3459–3472. [Google Scholar] [CrossRef]
  141. Knowlson, C.; Haddock, P.; Bingham, V.; McQuaid, S.; Mullan, P.B.; Buckley, N.E. Pin1 plays a key role in the response to treatment and clinical outcome in triple negative breast cancer. Ther. Adv. Med. Oncol. 2020, 12, 1758835920906047. [Google Scholar] [CrossRef]
  142. Koikawa, K.; Kibe, S.; Suizu, F.; Sekino, N.; Kim, N.; Manz, T.D.; Pinch, B.J.; Akshinthala, D.; Verma, A.; Gaglia, G.; et al. Targeting Pin1 renders pancreatic cancer eradicable by synergizing with immunochemotherapy. Cell 2021, 184, 4753–4771.e27. [Google Scholar] [CrossRef] [PubMed]
  143. Nong, S.; Han, X.; Xiang, Y.; Qian, Y.; Wei, Y.; Zhang, T.; Tian, K.; Shen, K.; Yang, J.; Ma, X. Metabolic reprogramming in cancer: Mechanisms and therapeutics. MedComm 2023, 4, e218. [Google Scholar] [CrossRef] [PubMed]
  144. Nakatsu, Y.; Yamamotoya, T.; Ueda, K.; Ono, H.; Inoue, M.-K.; Matsunaga, Y.; Kushiyama, A.; Sakoda, H.; Fujishiro, M.; Matsubara, A.; et al. Prolyl isomerase Pin1 in metabolic reprogramming of cancer cells. Cancer Lett. 2020, 470, 106–114. [Google Scholar] [CrossRef] [PubMed]
  145. Masoud, G.N.; Li, W. HIF-1α pathway: Role, regulation and intervention for cancer therapy. Acta Pharm. Sin. B 2015, 5, 378–389. [Google Scholar] [CrossRef] [PubMed]
  146. Jaakkola, P.; Mole, D.R.; Tian, Y.M.; Wilson, M.I.; Gielbert, J.; Gaskell, S.J.; von Kriegsheim, A.; Hebestreit, H.F.; Mukherji, M.; Schofield, C.J.; et al. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 2001, 292, 468–472. [Google Scholar] [CrossRef]
  147. Salceda, S.; Caro, J. Hypoxia-inducible factor 1alpha (HIF-1alpha) protein is rapidly degraded by the ubiquitin-proteasome system under normoxic conditions. Its stabilization by hypoxia depends on redox-induced changes. J. Biol. Chem. 1997, 272, 22642–22647. [Google Scholar] [CrossRef] [PubMed]
  148. Bernardi, R.; Guernah, I.; Jin, D.; Grisendi, S.; Alimonti, A.; Teruya-Feldstein, J.; Cordon-Cardo, C.; Simon, M.C.; Rafii, S.; Pandolfi, P.P. PML inhibits HIF-1alpha translation and neoangiogenesis through repression of mTOR. Nature 2006, 442, 779–785. [Google Scholar] [CrossRef]
  149. Lin, Y.C.; Lu, L.T.; Chen, H.Y.; Duan, X.; Lin, X.; Feng, X.H.; Tang, M.J.; Chen, R.H. SCP phosphatases suppress renal cell carcinoma by stabilizing PML and inhibiting mTOR/HIF signaling. Cancer Res. 2014, 74, 6935–6946. [Google Scholar] [CrossRef] [PubMed]
  150. Zhang, A.; Tao, W.; Zhai, K.; Fang, X.; Huang, Z.; Yu, J.S.; Sloan, A.E.; Rich, J.N.; Zhou, W.; Bao, S. Protein sumoylation with SUMO1 promoted by Pin1 in glioma stem cells augments glioblastoma malignancy. Neuro-Oncology 2020, 22, 1809–1821. [Google Scholar] [CrossRef]
  151. Ma, Q. Role of nrf2 in oxidative stress and toxicity. Annu. Rev. Pharmacol. Toxicol. 2013, 53, 401–426. [Google Scholar] [CrossRef]
  152. Jaramillo, M.C.; Zhang, D.D. The emerging role of the Nrf2-Keap1 signaling pathway in cancer. Genes Dev. 2013, 27, 2179–2191. [Google Scholar] [CrossRef] [PubMed]
  153. Li, R.; Jia, Z.; Zhu, H. Regulation of Nrf2 Signaling. React. Oxyg. Species 2019, 8, 312–322. [Google Scholar] [CrossRef]
  154. Liang, C.; Shi, S.; Liu, M.; Qin, Y.; Meng, Q.; Hua, J.; Ji, S.; Zhang, Y.; Yang, J.; Xu, J.; et al. PIN1 Maintains Redox Balance via the c-Myc/NRF2 Axis to Counteract Kras-Induced Mitochondrial Respiratory Injury in Pancreatic Cancer Cells. Cancer Res. 2019, 79, 133–145. [Google Scholar] [CrossRef]
  155. Saeidi, S.; Kim, S.J.; Han, H.J.; Kim, S.H.; Zheng, J.; Lee, H.B.; Han, W.; Noh, D.Y.; Na, H.K.; Surh, Y.J. H-Ras induces Nrf2-Pin1 interaction: Implications for breast cancer progression. Toxicol. Appl. Pharm. 2020, 402, 115121. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, Z.; Hu, Q.; Ye, S.; Xiang, L. Inhibition of the PIN1-NRF2/GPX4 axis imparts sensitivity to cisplatin in cervical cancer cells. Acta Biochim. Biophys. Sin. 2022, 54, 1325–1335. [Google Scholar] [CrossRef] [PubMed]
  157. Liang, R.; Song, G. Matrix stiffness-driven cancer progression and the targeted therapeutic strategy. Mechanobiol. Med. 2023, 1, 100013. [Google Scholar] [CrossRef]
  158. Ortega, Á.; Vera, I.; Diaz, M.P.; Navarro, C.; Rojas, M.; Torres, W.; Parra, H.; Salazar, J.; De Sanctis, J.B.; Bermúdez, V. The YAP/TAZ Signaling Pathway in the Tumor Microenvironment and Carcinogenesis: Current Knowledge and Therapeutic Promises. Int. J. Mol. Sci. 2021, 23, 430. [Google Scholar] [CrossRef] [PubMed]
  159. Chen, Y.; Wang, Y.; Zhai, Y.; Yuan, Y.; Wang, J.; Jin, Y.; Dang, L.; Song, L.; Chen, C.; Wang, Y. Cinobufacini injection suppresses the proliferation of human osteosarcoma cells by inhibiting PIN1-YAP/TAZ signaling pathway. Front. Pharmacol. 2023, 14, 1081363. [Google Scholar] [CrossRef]
  160. Yeung, B.; Khanal, P.; Mehta, V.; Trinkle-Mulcahy, L.; Yang, X. Identification of Cdk1-LATS-Pin1 as a Novel Signaling Axis in Anti-tubulin Drug Response of Cancer Cells. Mol. Cancer Res. 2018, 16, 1035–1045. [Google Scholar] [CrossRef]
  161. Sang, Y.; Li, Y.; Song, L.; Alvarez, A.A.; Zhang, W.; Lv, D.; Tang, J.; Liu, F.; Chang, Z.; Hatakeyama, S.; et al. TRIM59 Promotes Gliomagenesis by Inhibiting TC45 Dephosphorylation of STAT3. Cancer Res. 2018, 78, 1792–1804. [Google Scholar] [CrossRef]
  162. Ramanujn, A.; Tiwari, S. APC/C and retinoblastoma interaction: Cross-talk of retinoblastoma protein with the ubiquitin proteasome pathway. Biosci. Rep. 2016, 36, e00377. [Google Scholar] [CrossRef] [PubMed]
  163. Narasimha, A.M.; Kaulich, M.; Shapiro, G.S.; Choi, Y.J.; Sicinski, P.; Dowdy, S.F. Cyclin D activates the Rb tumor suppressor by mono-phosphorylation. eLife 2014, 3, e02872. [Google Scholar] [CrossRef]
  164. Burke, J.R.; Deshong, A.J.; Pelton, J.G.; Rubin, S.M. Phosphorylation-induced conformational changes in the retinoblastoma protein inhibit E2F transactivation domain binding. J. Biol. Chem. 2010, 285, 16286–16293. [Google Scholar] [CrossRef] [PubMed]
  165. Rizzolio, F.; Lucchetti, C.; Caligiuri, I.; Marchesi, I.; Caputo, M.; Klein-Szanto, A.J.; Bagella, L.; Castronovo, M.; Giordano, A. Retinoblastoma tumor-suppressor protein phosphorylation and inactivation depend on direct interaction with Pin1. Cell Death Differ. 2012, 19, 1152–1161. [Google Scholar] [CrossRef] [PubMed]
  166. Tong, Y.; Ying, H.; Liu, R.; Li, L.; Bergholz, J.; Xiao, Z.X. Pin1 inhibits PP2A-mediated Rb dephosphorylation in regulation of cell cycle and S-phase DNA damage. Cell Death Dis. 2015, 6, e1640. [Google Scholar] [CrossRef] [PubMed]
  167. Yin, L.; Duan, J.-J.; Bian, X.-W.; Yu, S.-c. Triple-negative breast cancer molecular subtyping and treatment progress. Breast Cancer Res. 2020, 22, 61. [Google Scholar] [CrossRef] [PubMed]
  168. Kozono, S.; Lin, Y.-M.; Seo, H.-S.; Pinch, B.; Lian, X.; Qiu, C.; Herbert, M.K.; Chen, C.-H.; Tan, L.; Gao, Z.J.; et al. Arsenic targets Pin1 and cooperates with retinoic acid to inhibit cancer-driving pathways and tumor-initiating cells. Nat. Commun. 2018, 9, 3069. [Google Scholar] [CrossRef] [PubMed]
  169. Campaner, E.; Rustighi, A.; Zannini, A.; Cristiani, A.; Piazza, S.; Ciani, Y.; Kalid, O.; Golan, G.; Baloglu, E.; Shacham, S.; et al. A covalent PIN1 inhibitor selectively targets cancer cells by a dual mechanism of action. Nat. Commun. 2017, 8, 15772. [Google Scholar] [CrossRef]
  170. Dubiella, C.; Pinch, B.J.; Koikawa, K.; Zaidman, D.; Poon, E.; Manz, T.D.; Nabet, B.; He, S.; Resnick, E.; Rogel, A.; et al. Sulfopin is a covalent inhibitor of Pin1 that blocks Myc-driven tumors in vivo. Nat. Chem. Biol. 2021, 17, 954–963. [Google Scholar] [CrossRef]
  171. Estey, M.P.; Di Ciano-Oliveira, C.; Froese, C.D.; Fung, K.Y.Y.; Steels, J.D.; Litchfield, D.W.; Trimble, W.S. Mitotic regulation of SEPT9 protein by cyclin-dependent kinase 1 (Cdk1) and Pin1 protein is important for the completion of cytokinesis. J. Biol. Chem. 2013, 288, 30075–30086. [Google Scholar] [CrossRef]
  172. Tarasewicz, E.; Rivas, L.; Hamdan, R.; Dokic, D.; Parimi, V.; Bernabe, B.P.; Thomas, A.; Shea, L.D.; Jeruss, J.S. Inhibition of CDK-mediated phosphorylation of Smad3 results in decreased oncogenesis in triple negative breast cancer cells. Cell Cycle 2014, 13, 3191–3201. [Google Scholar] [CrossRef] [PubMed]
  173. Thomas, A.L.; Lind, H.; Hong, A.; Dokic, D.; Oppat, K.; Rosenthal, E.; Guo, A.; Thomas, A.; Hamden, R.; Jeruss, J.S. Inhibition of CDK-mediated Smad3 phosphorylation reduces the Pin1-Smad3 interaction and aggressiveness of triple negative breast cancer cells. Cell Cycle 2017, 16, 1453–1464. [Google Scholar] [CrossRef] [PubMed]
  174. Dhillon, A.S.; Meikle, S.; Yazici, Z.; Eulitz, M.; Kolch, W. Regulation of Raf-1 activation and signalling by dephosphorylation. EMBO J. 2002, 21, 64–71. [Google Scholar] [CrossRef] [PubMed]
  175. Chauhan, A.S.; Garvin, A.J.; Jamshad, M.; Morris, J.R. Pin1-promoted SUMOylation of RNF168 restrains its chromatin accumulation. bioRxiv 2022. [Google Scholar] [CrossRef]
  176. Jo, A.; Yun, H.J.; Kim, J.Y.; Lim, S.C.; Choi, H.J.; Kang, B.S.; Choi, B.Y.; Choi, H.S. Prolyl isomerase PIN1 negatively regulates SGK1 stability to mediate tamoxifen resistance in breast cancer cells. Anticancer Res. 2015, 35, 785–794. [Google Scholar] [PubMed]
  177. Ma, S.L.; Pastorino, L.; Zhou, X.Z.; Lu, K.P. Prolyl isomerase Pin1 promotes amyloid precursor protein (APP) turnover by inhibiting glycogen synthase kinase-3β (GSK3β) activity: Novel mechanism for Pin1 to protect against Alzheimer disease. J. Biol. Chem. 2012, 287, 6969–6973. [Google Scholar] [CrossRef] [PubMed]
  178. Fagiani, F.; Govoni, S.; Racchi, M.; Lanni, C. The Peptidyl-prolyl Isomerase Pin1 in Neuronal Signaling: From Neurodevelopment to Neurodegeneration. Mol. Neurobiol. 2021, 58, 1062–1073. [Google Scholar] [CrossRef] [PubMed]
  179. Butterfield, D.A.; Abdul, H.M.; Opii, W.; Newman, S.F.; Joshi, G.; Ansari, M.A.; Sultana, R. Pin1 in Alzheimer’s disease. J. Neurochem. 2006, 98, 1697–1706. [Google Scholar] [CrossRef] [PubMed]
  180. Lu, K.P.; Zhou, X.Z. The prolyl isomerase PIN1: A pivotal new twist in phosphorylation signalling and disease. Nat. Rev. Mol. Cell Biol. 2007, 8, 904–916. [Google Scholar] [CrossRef]
  181. Wang, Y.; Mandelkow, E. Tau in physiology and pathology. Nat. Rev. Neurosci. 2016, 17, 5–21. [Google Scholar] [CrossRef]
  182. Smet, C.; Sambo, A.-V.; Wieruszeski, J.-M.; Leroy, A.; Landrieu, I.; Buée, L.; Lippens, G. The Peptidyl Prolyl cis/trans-Isomerase Pin1 Recognizes the Phospho-Thr212-Pro213 Site on Tau. Biochemistry 2004, 43, 2032–2040. [Google Scholar] [CrossRef] [PubMed]
  183. Malter, J.S. Pin1 and Alzheimer’s disease. Transl. Res. 2023, 254, 24–33. [Google Scholar] [CrossRef] [PubMed]
  184. Dakson, A.; Yokota, O.; Esiri, M.; Bigio, E.H.; Horan, M.; Pendleton, N.; Richardson, A.; Neary, D.; Snowden, J.S.; Robinson, A.; et al. Granular expression of prolyl-peptidyl isomerase PIN1 is a constant and specific feature of Alzheimer’s disease pathology and is independent of tau, Aβ and TDP-43 pathology. Acta Neuropathol. 2011, 121, 635–649. [Google Scholar] [CrossRef] [PubMed]
  185. Chen, D.; Mei, Y.; Kim, N.; Lan, G.; Gan, C.L.; Fan, F.; Zhang, T.; Xia, Y.; Wang, L.; Lin, C.; et al. Melatonin directly binds and inhibits death-associated protein kinase 1 function in Alzheimer’s disease. J. Pineal Res. 2020, 69, e12665. [Google Scholar] [CrossRef] [PubMed]
  186. Kim, N.; Wang, B.; Koikawa, K.; Nezu, Y.; Qiu, C.; Lee, T.H.; Zhou, X.Z. Inhibition of death-associated protein kinase 1 attenuates cis P-tau and neurodegeneration in traumatic brain injury. Prog. Neurobiol. 2021, 203, 102072. [Google Scholar] [CrossRef] [PubMed]
  187. Lee, T.H.; Chen, C.H.; Suizu, F.; Huang, P.; Schiene-Fischer, C.; Daum, S.; Zhang, Y.J.; Goate, A.; Chen, R.H.; Zhou, X.Z.; et al. Death-associated protein kinase 1 phosphorylates Pin1 and inhibits its prolyl isomerase activity and cellular function. Mol. Cell 2011, 42, 147–159. [Google Scholar] [CrossRef] [PubMed]
  188. Sultana, R.; Boyd-Kimball, D.; Poon, H.F.; Cai, J.; Pierce, W.M.; Klein, J.B.; Markesbery, W.R.; Zhou, X.Z.; Lu, K.P.; Butterfield, D.A. Oxidative modification and down-regulation of Pin1 in Alzheimer’s disease hippocampus: A redox proteomics analysis. Neurobiol. Aging 2006, 27, 918–925. [Google Scholar] [CrossRef] [PubMed]
  189. Chen, C.H.; Li, W.; Sultana, R.; You, M.H.; Kondo, A.; Shahpasand, K.; Kim, B.M.; Luo, M.L.; Nechama, M.; Lin, Y.M.; et al. Pin1 cysteine-113 oxidation inhibits its catalytic activity and cellular function in Alzheimer’s disease. Neurobiol. Dis. 2015, 76, 13–23. [Google Scholar] [CrossRef]
  190. Zheng, F.; Li, Y.; Zhang, F.; Sun, Y.; Zheng, C.; Luo, Z.; Wang, Y.L.; Aschner, M.; Zheng, H.; Lin, L.; et al. Cobalt induces neurodegenerative damages through Pin1 inactivation in mice and human neuroglioma cells. J. Hazard. Mater. 2021, 419, 126378. [Google Scholar] [CrossRef]
  191. Nakamura, K.; Greenwood, A.; Binder, L.; Bigio, E.H.; Denial, S.; Nicholson, L.; Zhou, X.Z.; Lu, K.P. Proline isomer-specific antibodies reveal the early pathogenic tau conformation in Alzheimer’s disease. Cell 2012, 149, 232–244. [Google Scholar] [CrossRef]
  192. Lim, J.; Balastik, M.; Lee, T.H.; Nakamura, K.; Liou, Y.C.; Sun, A.; Finn, G.; Pastorino, L.; Lee, V.M.; Lu, K.P. Pin1 has opposite effects on wild-type and P301L tau stability and tauopathy. J. Clin. Investig. 2008, 118, 1877–1889. [Google Scholar] [CrossRef]
  193. Tramutola, A.; Triani, F.; Di Domenico, F.; Barone, E.; Cai, J.; Klein, J.B.; Perluigi, M.; Butterfield, D.A. Poly-ubiquitin profile in Alzheimer disease brain. Neurobiol. Dis. 2018, 118, 129–141. [Google Scholar] [CrossRef]
  194. Xu, L.; Ren, Z.; Chow, F.E.; Tsai, R.; Liu, T.; Rizzolio, F.; Boffo, S.; Xu, Y.; Huang, S.; Lippa, C.F.; et al. Pathological Role of Peptidyl-Prolyl Isomerase Pin1 in the Disruption of Synaptic Plasticity in Alzheimer’s Disease. Neural Plast. 2017, 2017, 3270725. [Google Scholar] [CrossRef]
  195. Dosemeci, A.; Weinberg, R.J.; Reese, T.S.; Tao-Cheng, J.H. The Postsynaptic Density: There Is More than Meets the Eye. Front. Synaptic Neurosci. 2016, 8, 23. [Google Scholar] [CrossRef]
  196. Müller, U.C.; Deller, T.; Korte, M. Not just amyloid: Physiological functions of the amyloid precursor protein family. Nat. Rev. Neurosci. 2017, 18, 281–298. [Google Scholar] [CrossRef] [PubMed]
  197. Pastorino, L.; Sun, A.; Lu, P.J.; Zhou, X.Z.; Balastik, M.; Finn, G.; Wulf, G.; Lim, J.; Li, S.H.; Li, X.; et al. The prolyl isomerase Pin1 regulates amyloid precursor protein processing and amyloid-beta production. Nature 2006, 440, 528–534. [Google Scholar] [CrossRef] [PubMed]
  198. MacLeod, R.; Hillert, E.K.; Cameron, R.T.; Baillie, G.S. The role and therapeutic targeting of α-, β- and γ-secretase in Alzheimer’s disease. Future Sci. OA 2015, 1, Fso11. [Google Scholar] [CrossRef]
  199. Zhang, X.; Zhou, K.; Wang, R.; Cui, J.; Lipton, S.A.; Liao, F.-F.; Xu, H.; Zhang, Y.-W. Hypoxia-inducible Factor 1α (HIF-1α)-mediated Hypoxia Increases BACE1 Expression and β-Amyloid Generation. J. Biol. Chem. 2007, 282, 10873–10880. [Google Scholar] [CrossRef] [PubMed]
  200. Bulbarelli, A.; Lonati, E.; Cazzaniga, E.; Gregori, M.; Masserini, M. Pin1 affects Tau phosphorylation in response to Abeta oligomers. Mol. Cell. Neurosci. 2009, 42, 75–80. [Google Scholar] [CrossRef]
  201. Stefanis, L. α-Synuclein in Parkinson’s disease. Cold Spring Harb. Perspect. Med. 2012, 2, a009399. [Google Scholar] [CrossRef]
  202. Calabresi, P.; Mechelli, A.; Natale, G.; Volpicelli-Daley, L.; Di Lazzaro, G.; Ghiglieri, V. Alpha-synuclein in Parkinson’s disease and other synucleinopathies: From overt neurodegeneration back to early synaptic dysfunction. Cell Death Dis. 2023, 14, 176. [Google Scholar] [CrossRef] [PubMed]
  203. Ryo, A.; Togo, T.; Nakai, T.; Hirai, A.; Nishi, M.; Yamaguchi, A.; Suzuki, K.; Hirayasu, Y.; Kobayashi, H.; Perrem, K.; et al. Prolyl-isomerase Pin1 accumulates in lewy bodies of parkinson disease and facilitates formation of alpha-synuclein inclusions. J. Biol. Chem. 2006, 281, 4117–4125. [Google Scholar] [CrossRef] [PubMed]
  204. Ghosh, A.; Saminathan, H.; Kanthasamy, A.; Anantharam, V.; Jin, H.; Sondarva, G.; Harischandra, D.S.; Qian, Z.; Rana, A.; Kanthasamy, A.G. The peptidyl-prolyl isomerase Pin1 up-regulation and proapoptotic function in dopaminergic neurons: Relevance to the pathogenesis of Parkinson disease. J. Biol. Chem. 2013, 288, 21955–21971. [Google Scholar] [CrossRef] [PubMed]
  205. Shishido, T.; Nagano, Y.; Araki, M.; Kurashige, T.; Obayashi, H.; Nakamura, T.; Takahashi, T.; Matsumoto, M.; Maruyama, H. Synphilin-1 has neuroprotective effects on MPP(+)-induced Parkinson’s disease model cells by inhibiting ROS production and apoptosis. Neurosci. Lett. 2019, 690, 145–150. [Google Scholar] [CrossRef]
  206. MacDonald, M.E.; Ambrose, C.M.; Duyao, M.P.; Myers, R.H.; Lin, C.; Srinidhi, L.; Barnes, G.; Taylor, S.A.; James, M.; Groot, N.; et al. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. Cell 1993, 72, 971–983. [Google Scholar] [CrossRef]
  207. Schulte, J.; Littleton, J.T. The biological function of the Huntingtin protein and its relevance to Huntington’s Disease pathology. Curr. Trends Neurol. 2011, 5, 65–78. [Google Scholar]
  208. Saudou, F.; Humbert, S. The Biology of Huntingtin. Neuron 2016, 89, 910–926. [Google Scholar] [CrossRef]
  209. Arrasate, M.; Finkbeiner, S. Protein aggregates in Huntington’s disease. Exp. Neurol. 2012, 238, 1–11. [Google Scholar] [CrossRef]
  210. Grison, A.; Mantovani, F.; Comel, A.; Agostoni, E.; Gustincich, S.; Persichetti, F.; Del Sal, G. Ser46 phosphorylation and prolyl-isomerase Pin1-mediated isomerization of p53 are key events in p53-dependent apoptosis induced by mutant huntingtin. Proc. Natl. Acad. Sci. USA 2011, 108, 17979–17984. [Google Scholar] [CrossRef]
  211. Agostoni, E.; Michelazzi, S.; Maurutto, M.; Carnemolla, A.; Ciani, Y.; Vatta, P.; Roncaglia, P.; Zucchelli, S.; Leanza, G.; Mantovani, F.; et al. Effects of Pin1 Loss in Hdh(Q111) Knock-in Mice. Front. Cell. Neurosci. 2016, 10, 110. [Google Scholar] [CrossRef]
  212. Godin, J.D.; Poizat, G.; Hickey, M.A.; Maschat, F.; Humbert, S. Mutant huntingtin-impaired degradation of beta-catenin causes neurotoxicity in Huntington’s disease. EMBO J. 2010, 29, 2433–2445. [Google Scholar] [CrossRef] [PubMed]
  213. Bennett, E.J.; Bence, N.F.; Jayakumar, R.; Kopito, R.R. Global impairment of the ubiquitin-proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation. Mol. Cell 2005, 17, 351–365. [Google Scholar] [CrossRef] [PubMed]
  214. Bett, J.S.; Cook, C.; Petrucelli, L.; Bates, G.P. The ubiquitin-proteasome reporter GFPu does not accumulate in neurons of the R6/2 transgenic mouse model of Huntington’s disease. PLoS ONE 2009, 4, e5128. [Google Scholar] [CrossRef] [PubMed]
  215. Mitra, S.; Tsvetkov, A.S.; Finkbeiner, S. Single neuron ubiquitin-proteasome dynamics accompanying inclusion body formation in huntington disease. J. Biol. Chem. 2009, 284, 4398–4403. [Google Scholar] [CrossRef] [PubMed]
  216. Ortega, Z.; Díaz-Hernández, M.; Maynard, C.J.; Hernández, F.; Dantuma, N.P.; Lucas, J.J. Acute polyglutamine expression in inducible mouse model unravels ubiquitin/proteasome system impairment and permanent recovery attributable to aggregate formation. J. Neurosci. 2010, 30, 3675–3688. [Google Scholar] [CrossRef] [PubMed]
  217. Bett, J.S.; Goellner, G.M.; Woodman, B.; Pratt, G.; Rechsteiner, M.; Bates, G.P. Proteasome impairment does not contribute to pathogenesis in R6/2 Huntington’s disease mice: Exclusion of proteasome activator REGgamma as a therapeutic target. Hum. Mol. Genet. 2006, 15, 33–44. [Google Scholar] [CrossRef] [PubMed]
  218. Yu, D.; Pendergraff, H.; Liu, J.; Kordasiewicz, H.B.; Cleveland, D.W.; Swayze, E.E.; Lima, W.F.; Crooke, S.T.; Prakash, T.P.; Corey, D.R. Single-stranded RNAs use RNAi to potently and allele-selectively inhibit mutant huntingtin expression. Cell 2012, 150, 895–908. [Google Scholar] [CrossRef] [PubMed]
  219. Wild, E.J.; Tabrizi, S.J. Targets for future clinical trials in Huntington’s disease: What’s in the pipeline? Mov. Disord. 2014, 29, 1434–1445. [Google Scholar] [CrossRef] [PubMed]
  220. Duyao, M.P.; Auerbach, A.B.; Ryan, A.; Persichetti, F.; Barnes, G.T.; McNeil, S.M.; Ge, P.; Vonsattel, J.P.; Gusella, J.F.; Joyner, A.L.; et al. Inactivation of the mouse Huntington’s disease gene homolog Hdh. Science 1995, 269, 407–410. [Google Scholar] [CrossRef]
  221. White, J.K.; Auerbach, W.; Duyao, M.P.; Vonsattel, J.P.; Gusella, J.F.; Joyner, A.L.; MacDonald, M.E. Huntingtin is required for neurogenesis and is not impaired by the Huntington’s disease CAG expansion. Nat. Genet. 1997, 17, 404–410. [Google Scholar] [CrossRef]
  222. Cattaneo, E.; Rigamonti, D.; Goffredo, D.; Zuccato, C.; Squitieri, F.; Sipione, S. Loss of normal huntingtin function: New developments in Huntington’s disease research. Trends Neurosci. 2001, 24, 182–188. [Google Scholar] [CrossRef] [PubMed]
  223. Reiner, A.; Dragatsis, I.; Zeitlin, S.; Goldowitz, D. Wild-type huntingtin plays a role in brain development and neuronal survival. Mol. Neurobiol. 2003, 28, 259–276. [Google Scholar] [CrossRef]
  224. Tabrizi, S.J.; Leavitt, B.R.; Landwehrmeyer, G.B.; Wild, E.J.; Saft, C.; Barker, R.A.; Blair, N.F.; Craufurd, D.; Priller, J.; Rickards, H.; et al. Targeting Huntingtin Expression in Patients with Huntington’s Disease. N. Engl. J. Med. 2019, 380, 2307–2316. [Google Scholar] [CrossRef] [PubMed]
  225. Chen, Y.; Hou, X.; Pang, J.; Yang, F.; Li, A.; Lin, S.; Lin, N.; Lee, T.H.; Liu, H. The role of peptidyl-prolyl isomerase Pin1 in neuronal signaling in epilepsy. Front. Mol. Neurosci. 2022, 15, 1006419. [Google Scholar] [CrossRef] [PubMed]
  226. Merlin, L.R. Impact of protein kinase C activation on status epilepticus and epileptogenesis: Oh, what a tangled web. Epilepsy Curr. 2008, 8, 101–103. [Google Scholar] [CrossRef] [PubMed]
  227. Yang, Y.; Igumenova, T.I. The C-Terminal V5 Domain of Protein Kinase Cα Is Intrinsically Disordered, with Propensity to Associate with a Membrane Mimetic. PLoS ONE 2013, 8, e65699. [Google Scholar] [CrossRef] [PubMed]
  228. Colledge, M.; Snyder, E.M.; Crozier, R.A.; Soderling, J.A.; Jin, Y.; Langeberg, L.K.; Lu, H.; Bear, M.F.; Scott, J.D. Ubiquitination Regulates PSD-95 Degradation and AMPA Receptor Surface Expression. Neuron 2003, 40, 595–607. [Google Scholar] [CrossRef] [PubMed]
  229. Antonelli, R.; De Filippo, R.; Middei, S.; Stancheva, S.; Pastore, B.; Ammassari-Teule, M.; Barberis, A.; Cherubini, E.; Zacchi, P. Pin1 Modulates the Synaptic Content of NMDA Receptors via Prolyl-Isomerization of PSD-95. J. Neurosci. 2016, 36, 5437–5447. [Google Scholar] [CrossRef] [PubMed]
  230. Arumugam, T.V.; Baik, S.H.; Balaganapathy, P.; Sobey, C.G.; Mattson, M.P.; Jo, D.G. Notch signaling and neuronal death in stroke. Prog. Neurobiol. 2018, 165–167, 103–116. [Google Scholar] [CrossRef]
  231. Balaganapathy, P.; Baik, S.H.; Mallilankaraman, K.; Sobey, C.G.; Jo, D.G.; Arumugam, T.V. Interplay between Notch and p53 promotes neuronal cell death in ischemic stroke. J. Cereb. Blood Flow Metab. 2018, 38, 1781–1795. [Google Scholar] [CrossRef]
  232. Li, Q.M.; Tep, C.; Yune, T.Y.; Zhou, X.Z.; Uchida, T.; Lu, K.P.; Yoon, S.O. Opposite regulation of oligodendrocyte apoptosis by JNK3 and Pin1 after spinal cord injury. J. Neurosci. 2007, 27, 8395–8404. [Google Scholar] [CrossRef] [PubMed]
  233. Chang, H.Y.; Nishitoh, H.; Yang, X.; Ichijo, H.; Baltimore, D. Activation of apoptosis signal-regulating kinase 1 (ASK1) by the adapter protein Daxx. Science 1998, 281, 1860–1863. [Google Scholar] [CrossRef] [PubMed]
  234. Cheon, S.Y.; Cho, K.J.; Kim, S.Y.; Kam, E.H.; Lee, J.E.; Koo, B.N. Blockade of Apoptosis Signal-Regulating Kinase 1 Attenuates Matrix Metalloproteinase 9 Activity in Brain Endothelial Cells and the Subsequent Apoptosis in Neurons after Ischemic Injury. Front. Cell. Neurosci. 2016, 10, 213. [Google Scholar] [CrossRef] [PubMed]
  235. Slastnikova, T.A.; Ulasov, A.V.; Rosenkranz, A.A.; Sobolev, A.S. Targeted Intracellular Delivery of Antibodies: The State of the Art. Front. Pharmacol. 2018, 9, 1208. [Google Scholar] [CrossRef] [PubMed]
  236. Miersch, S.; Sidhu, S.S. Intracellular targeting with engineered proteins. F1000Research 2016, 5. [Google Scholar] [CrossRef] [PubMed]
  237. Aebersold, R.; Agar, J.N.; Amster, I.J.; Baker, M.S.; Bertozzi, C.R.; Boja, E.S.; Costello, C.E.; Cravatt, B.F.; Fenselau, C.; Garcia, B.A.; et al. How many human proteoforms are there? Nat. Chem. Biol. 2018, 14, 206–214. [Google Scholar] [CrossRef]
  238. Lazo, J.S.; Sharlow, E.R. Drugging Undruggable Molecular Cancer Targets. Annu. Rev. Pharmacol. Toxicol. 2016, 56, 23–40. [Google Scholar] [CrossRef] [PubMed]
  239. Crews, C.M. Targeting the undruggable proteome: The small molecules of my dreams. Chem. Biol. 2010, 17, 551–555. [Google Scholar] [CrossRef] [PubMed]
  240. Herrmann, A.; Nagao, T.; Zhang, C.; Lahtz, C.; Li, Y.J.; Yue, C.; Mulfarth, R.; Yu, H. An effective cell-penetrating antibody delivery platform. JCI Insight 2019, 4, e127474. [Google Scholar] [CrossRef]
  241. Alarcon-Segovia, D.; Ruiz-Arguelles, A.; Fishbein, E. Antibody to nuclear ribonucleoprotein penetrates live human mononuclear cells through Fc receptors. Nature 1978, 271, 67–69. [Google Scholar] [CrossRef]
  242. Alarcon-Segovia, D.; Ruiz-Arguelles, A.; Llorente, L. Antibody penetration into living cells. II. Anti-ribonucleoprotein IgG penetrates into Tgamma lymphocytes causing their deletion and the abrogation of suppressor function. J. Immunol. 1979, 122, 1855–1862. [Google Scholar] [CrossRef] [PubMed]
  243. Foster, M.H.; Kieber-Emmons, T.; Ohliger, M.; Madaio, M.P. Molecular and structural analysis of nuclear localizing anti-DNA lupus antibodies. Immunol. Res. 1994, 13, 186–206. [Google Scholar] [CrossRef]
  244. Douglas, J.N.; Gardner, L.A.; Levin, M.C. Antibodies to an Intracellular Antigen Penetrate Neuronal Cells and Cause Deleterious Effects. J. Clin. Cell. Immunol. 2013, 4, 134. [Google Scholar] [CrossRef]
  245. Weisbart, R.H.; Chan, G.; Jordaan, G.; Noble, P.W.; Liu, Y.; Glazer, P.M.; Nishimura, R.N.; Hansen, J.E. DNA-dependent targeting of cell nuclei by a lupus autoantibody. Sci. Rep. 2015, 5, 12022. [Google Scholar] [CrossRef] [PubMed]
  246. Choi, D.K.; Bae, J.; Shin, S.M.; Shin, J.Y.; Kim, S.; Kim, Y.S. A general strategy for generating intact, full-length IgG antibodies that penetrate into the cytosol of living cells. MAbs 2014, 6, 1402–1414. [Google Scholar] [CrossRef] [PubMed]
  247. Weisbart, R.H.; Gera, J.F.; Chan, G.; Hansen, J.E.; Li, E.; Cloninger, C.; Levine, A.J.; Nishimura, R.N. A cell-penetrating bispecific antibody for therapeutic regulation of intracellular targets. Mol. Cancer Ther. 2012, 11, 2169–2173. [Google Scholar] [CrossRef] [PubMed]
  248. Jang, J.Y.; Jeong, J.G.; Jun, H.R.; Lee, S.C.; Kim, J.S.; Kim, Y.S.; Kwon, M.H. A nucleic acid-hydrolyzing antibody penetrates into cells via caveolae-mediated endocytosis, localizes in the cytosol and exhibits cytotoxicity. Cell. Mol. Life Sci. 2009, 66, 1985–1997. [Google Scholar] [CrossRef] [PubMed]
  249. Rhodes, D.A.; Isenberg, D.A. TRIM21 and the Function of Antibodies inside Cells. Trends Immunol. 2017, 38, 916–926. [Google Scholar] [CrossRef] [PubMed]
  250. Noble, P.W.; Bernatsky, S.; Clarke, A.E.; Isenberg, D.A.; Ramsey-Goldman, R.; Hansen, J.E. DNA-damaging autoantibodies and cancer: The lupus butterfly theory. Nat. Rev. Rheumatol. 2016, 12, 429–434. [Google Scholar] [CrossRef]
  251. Guo, K.; Li, J.; Tang, J.P.; Tan, C.P.; Hong, C.W.; Al-Aidaroos, A.Q.; Varghese, L.; Huang, C.; Zeng, Q. Targeting intracellular oncoproteins with antibody therapy or vaccination. Sci. Transl. Med. 2011, 3, 99ra85. [Google Scholar] [CrossRef]
  252. Dao, T.; Yan, S.; Veomett, N.; Pankov, D.; Zhou, L.; Korontsvit, T.; Scott, A.; Whitten, J.; Maslak, P.; Casey, E.; et al. Targeting the intracellular WT1 oncogene product with a therapeutic human antibody. Sci. Transl. Med. 2013, 5, 176ra133. [Google Scholar] [CrossRef] [PubMed]
  253. Boutajangout, A.; Ingadottir, J.; Davies, P.; Sigurdsson, E.M. Passive immunization targeting pathological phospho-tau protein in a mouse model reduces functional decline and clears tau aggregates from the brain. J. Neurochem. 2011, 118, 658–667. [Google Scholar] [CrossRef] [PubMed]
  254. Chai, X.; Wu, S.; Murray, T.K.; Kinley, R.; Cella, C.V.; Sims, H.; Buckner, N.; Hanmer, J.; Davies, P.; O’Neill, M.J.; et al. Passive immunization with anti-Tau antibodies in two transgenic models: Reduction of Tau pathology and delay of disease progression. J. Biol. Chem. 2011, 286, 34457–34467. [Google Scholar] [CrossRef] [PubMed]
  255. d’Abramo, C.; Acker, C.M.; Jimenez, H.T.; Davies, P. Tau passive immunotherapy in mutant P301L mice: Antibody affinity versus specificity. PLoS ONE 2013, 8, e62402. [Google Scholar] [CrossRef]
  256. Gu, J.; Congdon, E.E.; Sigurdsson, E.M. Two novel tau antibodies targeting the 396/404 region are primarily taken up by neurons and reduce tau pathology. J. Biol. Chem. 2013, 288, 33081–33095. [Google Scholar] [CrossRef]
  257. Yanamandra, K.; Kfoury, N.; Jiang, H.; Mahan, T.E.; Ma, S.; Maloney, S.E.; Wozniak, D.F.; Diamond, M.I.; Holtzman, D.M. Anti-tau antibodies that block tau aggregate seeding in vitro markedly decrease pathology and improve cognition in vivo. Neuron 2013, 80, 402–414. [Google Scholar] [CrossRef] [PubMed]
  258. Castillo-Carranza, D.L.; Sengupta, U.; Guerrero-Munoz, M.J.; Lasagna-Reeves, C.A.; Gerson, J.E.; Singh, G.; Estes, D.M.; Barrett, A.D.; Dineley, K.T.; Jackson, G.R.; et al. Passive immunization with Tau oligomer monoclonal antibody reverses tauopathy phenotypes without affecting hyperphosphorylated neurofibrillary tangles. J. Neurosci. 2014, 34, 4260–4272. [Google Scholar] [CrossRef] [PubMed]
  259. Sankaranarayanan, S.; Barten, D.M.; Vana, L.; Devidze, N.; Yang, L.; Cadelina, G.; Hoque, N.; DeCarr, L.; Keenan, S.; Lin, A.; et al. Passive immunization with phospho-tau antibodies reduces tau pathology and functional deficits in two distinct mouse tauopathy models. PLoS ONE 2015, 10, e0125614. [Google Scholar] [CrossRef] [PubMed]
  260. Pozzi, S.; Codron, P.; Soucy, G.; Renaud, L.; Cordeau, P.J.; Dutta, K.; Bareil, C.; Julien, J.P. Monoclonal full-length antibody against TAR DNA binding protein 43 reduces related proteinopathy in neurons. JCI Insight 2020, 5, e140420. [Google Scholar] [CrossRef]
  261. Nguyen, L.; Montrasio, F.; Pattamatta, A.; Tusi, S.K.; Bardhi, O.; Meyer, K.D.; Hayes, L.; Nakamura, K.; Banez-Coronel, M.; Coyne, A.; et al. Antibody Therapy Targeting RAN Proteins Rescues C9 ALS/FTD Phenotypes in C9orf72 Mouse Model. Neuron 2020, 105, 645–662.e611. [Google Scholar] [CrossRef]
  262. Congdon, E.E.; Sigurdsson, E.M. Tau-targeting therapies for Alzheimer disease. Nat. Rev. Neurol. 2018, 14, 399–415. [Google Scholar] [CrossRef] [PubMed]
  263. Foss, S.; Bottermann, M.; Jonsson, A.; Sandlie, I.; James, L.C.; Andersen, J.T. TRIM21—From Intracellular Immunity to Therapy. Front. Immunol. 2019, 10, 2049. [Google Scholar] [CrossRef] [PubMed]
  264. McEwan, W.A.; Tam, J.C.H.; Watkinson, R.E.; Bidgood, S.R.; Mallery, D.L.; James, L.C. Intracellular antibody-bound pathogens stimulate immune signaling via the Fc receptor TRIM21. Nat. Immunol. 2013, 14, 327–336. [Google Scholar] [CrossRef] [PubMed]
  265. Randow, F.; MacMicking, J.D.; James, L.C. Cellular self-defense: How cell-autonomous immunity protects against pathogens. Science 2013, 340, 701–706. [Google Scholar] [CrossRef]
  266. Wang, R.; Lu, K.P.; Zhou, X.Z. Function and regulation of cis P-tau in the pathogenesis and treatment of conventional and nonconventional tauopathies. J. Neurochem. 2023, 166, 904–914. [Google Scholar] [CrossRef]
  267. Kondo, A.; Shahpasand, K.; Mannix, R.; Qiu, J.; Moncaster, J.; Chen, C.-H.; Yao, Y.; Lin, Y.-M.; Driver, J.A.; Sun, Y.; et al. Antibody against early driver of neurodegeneration cis P-tau blocks brain injury and tauopathy. Nature 2015, 523, 431–436. [Google Scholar] [CrossRef] [PubMed]
  268. Albayram, O.; Kondo, A.; Mannix, R.; Smith, C.; Tsai, C.-Y.; Li, C.; Herbert, M.K.; Qiu, J.; Monuteaux, M.; Driver, J.; et al. Cis P-tau is induced in clinical and preclinical brain injury and contributes to post-injury sequelae. Nat. Commun. 2017, 8, 1000. [Google Scholar] [CrossRef] [PubMed]
  269. Albayram, O.; MacIver, B.; Mathai, J.; Verstegen, A.; Baxley, S.; Qiu, C.; Bell, C.; Caldarone, B.J.; Zhou, X.Z.; Lu, K.P.; et al. Traumatic Brain Injury-related voiding dysfunction in mice is caused by damage to rostral pathways, altering inputs to the reflex pathways. Sci. Rep. 2019, 9, 8646. [Google Scholar] [CrossRef] [PubMed]
  270. Qiu, C.; Albayram, O.; Kondo, A.; Wang, B.; Kim, N.; Arai, K.; Tsai, C.Y.; Bassal, M.A.; Herbert, M.K.; Washida, K.; et al. Cis P-tau underlies vascular contribution to cognitive impairment and dementia and can be effectively targeted by immunotherapy in mice. Sci. Transl. Med. 2021, 13, eaaz7615. [Google Scholar] [CrossRef]
  271. Foster, K.; Manca, M.; McClure, K.; Koivula, P.; Trojanowski, J.Q.; Havas, D.; Chancellor, S.; Goldstein, L.; Brunden, K.R.; Kraus, A.; et al. Preclinical characterization and IND-enabling safety studies for PNT001, an antibody that recognizes cis-pT231 tau. Alzheimer’s Dement. 2023, 19, 4662–4674. [Google Scholar] [CrossRef]
  272. Luca, W.; Foster, K.; McClure, K.; Ahlijanian, M.K.; Jefson, M. A Phase 1 Single-Ascending-Dose Trial in Healthy Volunteers to Evaluate the Safety, Tolerability, Pharmacokinetics, and Immunogenicity of Intravenous PNT001, a Novel Mid-domain Tau Antibody Targeting cis-pT231 Tau. J. Prev. Alzheimers Dis. 2024, 11, 366–374. [Google Scholar] [CrossRef] [PubMed]
  273. McEwan, W.A.; Falcon, B.; Vaysburd, M.; Clift, D.; Oblak, A.L.; Ghetti, B.; Goedert, M.; James, L.C. Cytosolic Fc receptor TRIM21 inhibits seeded tau aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 574–579. [Google Scholar] [CrossRef] [PubMed]
  274. Mukadam, A.S.; Miller, L.V.C.; Smith, A.E.; Vaysburd, M.; Sakya, S.A.; Sanford, S.; Keeling, S.; Tuck, B.J.; Katsinelos, T.; Green, C.; et al. Cytosolic antibody receptor TRIM21 is required for effective tau immunotherapy in mouse models. Science 2023, 379, 1336–1341. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of Pin1 involvement in ubiquitin-mediated degradation. After Pin1 binds and catalyzes a conformational change to a phosphorylated substrate, the resulting protein may have either decreased or increased susceptibility to ubiquitin-mediated degradation, depending on the molecular characteristics of the new conformation. Specifically, the resulting configuration may have an altered susceptibility to other post-translational mechanisms and polyubiquitination, as well as an affinity with ubiquitin factors. Furthermore, because Pin1 negatively regulates many E3 ligases, it may indirectly play a role in the stability of the substrate of these ubiquitin factors as well.
Figure 1. Overview of Pin1 involvement in ubiquitin-mediated degradation. After Pin1 binds and catalyzes a conformational change to a phosphorylated substrate, the resulting protein may have either decreased or increased susceptibility to ubiquitin-mediated degradation, depending on the molecular characteristics of the new conformation. Specifically, the resulting configuration may have an altered susceptibility to other post-translational mechanisms and polyubiquitination, as well as an affinity with ubiquitin factors. Furthermore, because Pin1 negatively regulates many E3 ligases, it may indirectly play a role in the stability of the substrate of these ubiquitin factors as well.
Cells 13 00731 g001
Figure 2. Stereospecific antibodies selectively deplete extracellular and intracellular cis P-tau. Pin1 plays a physiological role in safeguarding neurons against pathogenic cis-P-tau accumulation and neurofibrillary tangle formation by catalyzing the cis–trans isomerization of tau. However, when Pin1 is depleted during brain injury, cis-P-tau can accumulate and result in tau-associated neurodegeneration. Upon treatment with stereospecific cis-P-tau antibodies, cis-P-tau antibodies may bind to the p-Ser/Thr residues of their antigen and capture it extracellularly or intracellularly. Regardless, the antibody or antibody–antigen complexes enter the cell via Fc-receptor endocytosis. Finally, the TRIM21 E3 ligase engages the Fc region of the antibody, facilitating the ubiquitination and subsequent degradation of the TRIM21–antibody–antigen complex while sparing trans-P-tau or unphosphorylated, unaltered tau to carry out their physiological functions.
Figure 2. Stereospecific antibodies selectively deplete extracellular and intracellular cis P-tau. Pin1 plays a physiological role in safeguarding neurons against pathogenic cis-P-tau accumulation and neurofibrillary tangle formation by catalyzing the cis–trans isomerization of tau. However, when Pin1 is depleted during brain injury, cis-P-tau can accumulate and result in tau-associated neurodegeneration. Upon treatment with stereospecific cis-P-tau antibodies, cis-P-tau antibodies may bind to the p-Ser/Thr residues of their antigen and capture it extracellularly or intracellularly. Regardless, the antibody or antibody–antigen complexes enter the cell via Fc-receptor endocytosis. Finally, the TRIM21 E3 ligase engages the Fc region of the antibody, facilitating the ubiquitination and subsequent degradation of the TRIM21–antibody–antigen complex while sparing trans-P-tau or unphosphorylated, unaltered tau to carry out their physiological functions.
Cells 13 00731 g002
Table 1. List of Pin1 substrates in which stability is altered upon Pin1 isomerization.
Table 1. List of Pin1 substrates in which stability is altered upon Pin1 isomerization.
Substrate DiseaseReferences
Increased Stability
β-catenin Cancer[22]
c-MycCancer[23,24,25,26]
Cyclin D1Cancer[27]
p53Cancer[28,29,30]
Bcl-2Cancer[31]
NF-kBInflammation[10]
p73Cancer[32]
BIMELHuntington’s [33]
Emi1Cancer[34]
HBxHepB/Cancer[35]
Her2Cancer[36,37]
Mcl-1Cancer[38]
Notch1Cancer[17,39]
AKTCancer[40]
Cep55 Cancer[41]
p27Cancer[42,43]
v-RelCancer[44]
NCID1Cancer[45]
TaxHTLV/Cancer[46,47]
COX-2Inflammation[48,49]
PPARγCancer/Metabolism[50]
NanogCancer[51]
Integrase ParticleHIV[52]
4-OctCancer[53]
Mutant p53Cancer[54]
ADAR2 ALS[55,56]
ERαCancer[57,58]
HIPK2Cancer[59]
p63Cancer[60]
RBBP8DNA Repair[61]
Sp1Cancer[62]
HIF-1αCancer[63,64,65]
ERαCancer[66]
SeparaseCancer[67]
PML-RARaCancer[68]
CRMP2AAlzheimer’s[69]
PERIODCircadian Rhythm[70]
BRD4Cancer[71]
p27Cancer[42]
Tissue FactorCancer[72,73]
TRIM59Cancer[74]
YAP/TAZCancer[75,76]
ACC1Cancer[77]
BRCA1-BARD1Cancer[78]
FAAP20DNA repair[79]
BRCA1Cancer[80]
HBcHepB/Cancer[81]
HP1αCancer[82]
Nrf2Cancer[83]
HIF-2αCancer[84]
TRIM32Cancer[85]
HBV CP HepB/Cancer[81]
Decreased Stability
TauAlzheimer’s[86,87,88]
CF2Cancer[89]
RARaCancer[90]
BTKCancer[91]
IRF3Cancer[92]
Cyclin ECancer[93,94,95,96]
DaxxCancer[97]
SMRTCancer[98]
FOXO4Cancer[99]
PMLCancer[100,101]
A3GHIV[102]
Smad2/3Cancer[103]
GRK2Cancer[104]
CDK10Cancer[105]
Fbw7Cancer[23,106]
PKCParkinson’s[107]
BoraCancer[108]
SUV39H1Cancer[109]
RUNX3Cancer[110]
KLF10Cancer[111]
Che-1Cancer[112]
RESTNeurodegeneration[113]
Mutant HTTHuntington’s[114,115]
c-JunCancer[116]
IRF7Inflammation[117]
PRDM16 Metabolism[20]
PSD-95Epilepsy[118]
STK3Cancer[75]
ATGLMetabolism[119]
SIK1Cancer[120]
pVHLCancer[11]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jeong, J.; Usman, M.; Li, Y.; Zhou, X.Z.; Lu, K.P. Pin1-Catalyzed Conformation Changes Regulate Protein Ubiquitination and Degradation. Cells 2024, 13, 731. https://doi.org/10.3390/cells13090731

AMA Style

Jeong J, Usman M, Li Y, Zhou XZ, Lu KP. Pin1-Catalyzed Conformation Changes Regulate Protein Ubiquitination and Degradation. Cells. 2024; 13(9):731. https://doi.org/10.3390/cells13090731

Chicago/Turabian Style

Jeong, Jessica, Muhammad Usman, Yitong Li, Xiao Zhen Zhou, and Kun Ping Lu. 2024. "Pin1-Catalyzed Conformation Changes Regulate Protein Ubiquitination and Degradation" Cells 13, no. 9: 731. https://doi.org/10.3390/cells13090731

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