Next Article in Journal
Generation of Differentiating and Long-Living Intestinal Organoids Reflecting the Cellular Diversity of Canine Intestine
Next Article in Special Issue
Roles of GSK-3 and β-Catenin in Antiviral Innate Immune Sensing of Nucleic Acids
Previous Article in Journal
CRISPR/Cas9-Mediated Genome Editing Reveals Oosp Family Genes are Dispensable for Female Fertility in Mice
Previous Article in Special Issue
GSK3β: A Master Player in Depressive Disorder Pathogenesis and Treatment Responsiveness
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

GSK3: A Kinase Balancing Promotion and Resolution of Inflammation

Institute of Clinical Chemistry, Hannover Medical School, 30625 Hannover, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
These authors contributed equally to this work.
Submission received: 12 March 2020 / Revised: 25 March 2020 / Accepted: 26 March 2020 / Published: 28 March 2020

Abstract

:
GSK3 has been implicated for years in the regulation of inflammation and addressed in a plethora of scientific reports using a variety of experimental (disease) models and approaches. However, the specific role of GSK3 in the inflammatory process is still not fully understood and controversially discussed. Following a detailed overview of structure, function, and various regulatory levels, this review focusses on the immunoregulatory functions of GSK3, including the current knowledge obtained from animal models. Its impact on pro-inflammatory cytokine/chemokine profiles, bacterial/viral infections, and the modulation of associated pro-inflammatory transcriptional and signaling pathways is discussed. Moreover, GSK3 contributes to the resolution of inflammation on multiple levels, e.g., via the regulation of pro-resolving mediators, the clearance of apoptotic immune cells, and tissue repair processes. The influence of GSK3 on the development of different forms of stimulation tolerance is also addressed. Collectively, the role of GSK3 as a kinase balancing the initiation/perpetuation and the amelioration/resolution of inflammation is highlighted.

1. Introduction

Glycogen synthase kinase (GSK) 3 is a serine/threonine kinase consisting of two highly similar paralogs, GSK3α and β. With more than 40 definitely identified targets and more than 500 proposed candidate substrates, GSK3 is involved in the regulation of a plethora of intracellular/molecular, organic/cellular, and (patho-)physiological events [1]. For instance, GSK3 has an influence on transcriptional regulation, alternative splicing [2,3], and mRNA stability [4]. On the protein level, GSK3 affects translation and protein synthesis [5,6], protein activity, localization, and degradation [7] (including its participation in the β-catenin destruction complex [8]). Metabolic processes, such as glucose metabolism [7], lipid deposition [9] and accumulation [10], and processes associated with mitochondrial activity (biogenesis, bioenergetics, permeability, and motility [11]), are influenced by GSK3. In consequence, relevant cellular functions are governed by GSK3. These include proliferation, differentiation, apoptosis, [1], adhesion, and migration [12]. Amongst others, GSK3 is of importance for physiological processes, such as embryonic and tissue development [1,7], tumor suppression [13], innate and adaptive immune responses [14], a variety of neuronal functions, and circadian rhythm regulation [12].
GSK3 also plays a role in the pathogenesis of common diseases, e.g., neurological/neurodegenerative diseases [13,15]; diabetes mellitus [16]; inflammatory diseases [17], such as rheumatoid arthritis [18]; and different types of cancer [19]. Moreover, GSK3 is also of interest in geriatrics [15,20]. Thus, the modulation of GSK3 (especially GSK3β) activity via natural compounds [21] or the design of pharmacologically applicable inhibitors [19] is still a promising target for various therapeutic approaches.
Since the general aspects of GSK3 functions have been extensively reviewed elsewhere (e.g., [6,7,13,22] and others) they shall not be the topic of this review. Following the description of the basic molecular and mechanistic features of GSK3, we will focus on the janiform nature of GSK3 within the regulation of both the promotion and the termination of inflammation.

2. Structure and Function of GSK3

2.1. Protein Structure, Enzymatic Activity, and Substrate Specificity

2.1.1. General Aspects

GSK3 protein was initially purified in 1980 from rabbit skeletal muscle and characterized as an enzyme activating adenosine triphosphate/magnesium-dependent protein phosphatase [23,24] and phosphorylating glycogen synthase (GS) [25,26], thus inhibiting its activity and negatively controlling the final step of glycogen synthesis [6]. In 1990, nucleotide sequence analysis of a rat brain cDNA library revealed the existence of two highly similar GSK3 paralogs termed α and β [27] and further analyses led to the identification of two GSK3β splice variants (GSK3β1 and 2; see Section 2.2.1.) [28]. Due to their remarkable homology (see Section 2.2.1.), most GSK3-targeting small molecule inhibitors address the ATP binding pocket of both GSK3α and β, rendering their selective inhibition a persistent challenge [12]. However, due to the presence of a single amino acid (aa) exchange between α and β in the ATP binding pocket within the kinase domain (GSK3α-Glu196 → GSK3β-Asp133), the development of paralog-selective inhibitors has been reported recently (GSK3α: BRD0705, GSK3β: BRD3731) [29]. Though highly similar, GSK3α and β are differentially regulated and expressed, exhibit tissue-specific expression profiles/levels (see Section 2.2.), have common as well as distinct substrates and cellular functions, and cannot replace each other completely [12]. Moreover, the loss of one paralog does not result in an enhanced expression or activity of the remaining gene/protein [30].

2.1.2. Structure and Molecular Function

GSK3 is a monomeric serine/threonine kinase belonging to the CMGC kinase group (named after the founding kinase families cyclin-dependent kinases, mitogen-activated protein kinases (MAPK), GSK, and CDC-like kinases) [31]. GSK3β contains a negative regulatory N-terminal domain, the kinase domain including both the ATP binding site and the enzymatically active site, as well as a C-terminal domain also possessing negative regulatory capacity (for details, see Section 2.2.3.). GSK3α, basically characterized by the same protein structure, contains an extended and glycine-rich N-terminal domain [32,33]. Within the active site, conserved Lys residues (GSK3α: Lys148 and 149; GSK3β: Lys85 and 86) are responsible for ATP binding and catalyzing the γ-phosphate transfer to the substrate; thus, local point mutations (GSK3α: Lys148Arg; GSK3β: Lys85Arg) have a dominant-negative effect on GSK3 enzymatic activity [34].
GSK3 generally targets the consensus aa sequence Ser/Thr1-X1-X2-X3-Ser/Thr2. Though not absolutely required, X1 is often a proline [35]. In dependency of prior phosphorylation (“priming”) of the C-terminal serine/threonine residue (Ser/Thr2), the primed substrate is able to associate with the positively charged GSK3 substrate binding domain, thus enabling the phosphorylation of the respective substrate at the N-terminal serine or threonine (Ser/Thr1) [12,13,36] with a massively increased efficiency (up to 1000-fold) in comparison to unprimed substrates [32,37]. Though the number of intervening residues is generally three in most substrates, deviating numbers have been described [13]. Moreover, in GSK3 substrates that are subsequently addressed by β-TrCP (“F-box containing E3 ligase β-transducin repeat-containing protein”) for ubiquitin-mediated degradation, a more specific GSK3 recognition sequence, i.e., Asp-Ser-Gly-X-X-Ser, is present [7]. While the initial priming phosphorylation has to be mediated by another kinase, GSK3 is able to perform multiple successive phosphorylation steps at substrates containing a series of adjacent consensus sequences, thus generating its own priming phosphorylation for further phosphorylation events [38]. Only a limited number of substrates (e.g., Jun, avian myelocytomatosis viral oncogene homolog (Myc), histone H1.5, microtubule affinity-regulating kinase 2, CCAAT/enhancer binding protein (C/EBP) β, Tau protein, and p21) may be bound and phosphorylated at specific sites without priming phosphorylation [7,12,39], potentially due to the presence of acidic residues in the C-terminal position [13]. However, this assumption is still controversially discussed [7] and an influence of protein conformation has also been proposed [12]. In addition, other motifs in these proteins may be addressed by GSK3 via the conservative mechanism, as shown for Jun and Tau following Jun N-terminal kinase (JNK)-mediated priming phosphorylation [35] and C/EBPβ after extracellular signal-regulated kinase (ERK) 2-mediated priming [40].

2.2. Regulation of GSK3

2.2.1. GSK3 Gene Structure and mRNA Expression

Neither the GSK3α nor the GSK3β promoter contains a classical TATA box and in contrast to GSK3α, GSK3β also misses an initiator element [41,42]. Both genes are regulated by a variety of common transcription factors (TFs), such as activator protein (AP-)1, specificity protein 1, cyclic adenosine monophosphate (cAMP-)responsive element binding protein (CREB), and myeloid zinc finger 1 [41,42,43]. The GSK3α promoter additionally provides binding sites for late SV40 factor and Yin and Yang 1 [42], whilst GSK3β enables the binding of the TF Tst-1 (“Tuberculin skin test reactivity, absence of”), Myb (“myoblastosis”), TFCP2, and AP-2 [41]. The GSK3β promoter further possesses several CCAAT boxes and inverted CCAAT elements [41] that are recognized by members of the C/EBP family, such as C/EBPα and β [44]. Another binding site for TF AP-4 is generated by the minor allele of a single nucleotide polymorphism (SNP; rs334,558) in the GSK3β promoter [45].
Both genes possess approximately 85% overall sequence identity, including 98% identity within the catalytic domain. The C-terminal section is more variable, showing only 36% homology [27,32]. In addition, GSK3α and β are widely conserved among a variety of different species [46], with the exception of birds, which proved to be naturally occurring GSK3α knock-out (KO) organisms [47]. The human GSK3α gene (GSK3A, chromosomal localization: 19q13.2) is expressed as a single mRNA comprising 11 exons with 2193 bases (gene bank accession number: NM_019884.3; average G/C content: 61%), including a 5′ untranslated region (UTR) of 137 bases (83% G/C), the protein coding sequence (1452 bases, including stop codon; 61% G/C), and a 3′UTR of 604 bases (58% G/C) [48]. Human GSK3β (GSK3B, 3q13.33) is encoded by a total of 12 exons, yielding three variants comprising 7134 (NM_002093.3, transcript variant 1), 7095 (NM_001146156.1, transcript variant 2), and 6982 bases (NM_001354596.1, transcript variant 3). Each variant (42% GC) includes a 5′UTR of 983 bases (58% G/C), the protein coding sequence (1302 bases for variants 1 and 3, 1263 bases for variant 2 lacking exon 9; 47% G/C), and a longer 3’UTR (4849 bases for variants 1 and 2, 4697 bases for variant 3; 37% G/C) [48,49]. Generation of the different variants depends on RNA helicases DEAD-box 5- and 17-controlled [50] alternative splicing [51] that may be influenced by the occurrence of intronic SNPs [45]. In other organisms, however, divergent numbers of variants have been described, e.g., five differentially expressed porcine GSK3β isoforms [52] and six transcripts in the goat [53].
GSK3α mRNA and protein as well as GSK3β mRNA are ubiquitously but variably expressed, mostly at medium to high levels, in all human tissues and organs. GSK3β protein expression, however, is slightly more restricted; can hardly be detected in human heart, skeletal, and smooth muscles as well as in lung, vagina, and soft tissues; and is most strongly expressed in the brain and neuronal tissue [54]. Moreover, in many tissues, a remarkable discrepancy between mRNA and protein levels can be observed [46,54], an effect that may be ascribed to differentially regulated and paralog-specific transcriptional and translational events [46]. However, in GSK3-expressing cells/tissues, its expression levels remain relatively stable and only a limited number of cellular conditions appear to exist in which GSK3 expression is significantly regulated [13], including murine brain development [55] and maturation of murine Th17 lymphocytes [56]. GSK3 mRNA amounts may be adjusted by mRNA stability-modifying proteins, such as human antigen R [57].

2.2.2. GSK3 Proteins—Paralogs, Isoforms, Expression, and Localization

Translation of the human GSK3 mRNAs results in the synthesis of one GSK3α (protein bank accession number: NP_063937.2) and two GSK3β protein isoforms, i.e., GSK3β1 (NP_001139628.1, derived from transcript variant 2) and GSK3β2 (NP_002084.2, derived from transcript variants 1 and 3). GSK3α is a 51 kDa protein consisting of 483 aa. Both GSK3β isoforms are slightly shorter (GSK3β1: 420 aa, approximately 47 kDa; GSK3β2: 433 aa, approximately 49 kDa) [12,39,58]. GSK3β1 is the ubiquitously expressed isoform, whereas the longer GSK3β2 variant (first identified in 2002 in the rat brain [28]) possesses an additional external loop of 13 aa located within the catalytic domain between residues 303 and 304 [12,59], exhibits differential substrate specificity and efficiency in comparison to GSK3β1 [60], and appears to be predominately expressed in neurons and involved in neuronal functions [12,61,62].
GSK3 paralogs show distinct patterns of intracellular distribution [12]. GSK3α, though in principle capable of nuclear translocation, is virtually exclusively located in the cytosol due to its fast export from the nucleus [63]. GSK3β is also mainly a cytoplasmic protein but nonetheless well detectable in both the nucleus and mitochondria [64] and at further membrane compartments, such as endosomes and lysosomes [65]. However, an aberrant accumulation of active GSK3β has been described in several cancer cells [19]. A 19 aa comprising nuclear localization sequence (NLS) has been described to be present in a basic loop (aa 85–123) within the kinase domain. To ensure cytoplasmic localization, the NLS is sequestered by cytosolic protein complexes [66]. Interestingly, the amount of active GSK3β is significantly higher in the extra-cytoplasmic areas (5- to 8-fold) [64]. Intracellular localization is controlled at different regulatory levels [12]. These include (de-)phosphorylation [67], association with proteins facilitating nuclear import (e.g., karyopherin β2 [68] or herpesvirus-derived latency-associated nuclear antigen (LANA) [69]) or export (e.g., GSK3-binding protein [70]), and partial digestion (see Section 2.2.3.). Recently, it has been shown that inactivation of the phosphatidylinositol-3-kinase (PI3K), protein kinase B (PKB/Akt), mechanistic target of rapamycin complex (mTORC) 1 axis results in the nuclear accumulation of both GSK3α and β. In contrast, active PI3K-Akt-mTORC1 favors the retention of GSK3β in the cytosol, including (partial) localization at endomembrane compartments [65].

2.2.3. Posttranslational Modifications

Phosphorylation

GSK3 shows a constitutively high basal activity in resting cells, which can be affected by post-translational modifications [12,13]. The presence of an activating phosphorylation at Tyr279 (GSK3α) and Tyr216 (GSK3β) is assumed to be necessary for proper formation of the catalytic domain during protein folding [39] and has been demonstrated to significantly facilitate substrate accessibility [71], to increase GSK3 protein stability [72], and to be essential for maximum enzymatic activity [73,74], thus increasing its catalytic capacity (approximately 5-fold, [75]). However, replacing this tyrosine by phenylalanine via in vitro mutagenesis had only little effects on GSK3 activity [34,75,76], indicating that Tyr279/216 phosphorylation is not an indispensable prerequisite for basal enzymatic activity. Moreover, these sites are prone to auto-phosphorylation [72,77] and are already phosphorylated following translation during heat shock protein 90 (Hsp90)-dependent protein folding, thus directly yielding active enzymes [13,78]. The tyrosine protein kinases src, fyn, and Pyk2 may further contribute to Tyr279/216 phosphorylation [79]. Though Tyr279/216 phosphorylation proved to be very stable [72], it can be dynamically regulated under certain conditions as demonstrated in models of neurodegeneration [7,80]. Decreased GSK3 activity is associated with reduced tyrosine phosphorylation under various conditions and thus, a regulatory role for phospho-tyrosine phosphatases has been proposed [13].
However, the major regulatory influence on GSK3 is mediated via inhibitory events, mainly in response to a variety of stimulating agents (e.g., insulin, serum, growth factors), leading to the fast reduction of enzymatic activity (30%–70% within 10 min) [12]. At the molecular level, this is predominantly mediated via inhibitory phosphorylation at different residues mainly located in the N-terminal region [79]. Among these, Ser21 and Ser9 are regarded as the most important regulatory sites in GSK3α and β [39,79] due to the function of the N-terminal protein domain as a pseudo-substrate [6]. This feature results from the interaction of phosphorylated Ser21/9 with three highly conserved basic residues [81] located within the priming phosphate binding site (GSK3α: Arg159, Arg243, and Lys268; GSK3β: Arg96, Arg180, and Lys205 [34,71,74,82]) blocking both the primed substrate binding and the catalytic domain [6]. Thus, Ser21/9 phosphorylation is sufficient for overriding Tyr279/216 phosphorylation-dependent GSK3 activation [80]. This mechanism also opens the possibility that phosphorylation of the few unprimed GSK3 substrates may not be negatively affected by Ser21/9 phosphorylation [39].
In GSK3β, activating effects of Ser147 phosphorylation [83] and inhibitory effects of phosphorylated Thr43 (acting as a priming phosphorylation for Ser9-dependent inactivation [84]) as well as Ser389 and Thr390 have also been observed [39,79]. In mice, an additional dual-specificity tyrosine phosphorylation-regulated kinase 1A-mediated inhibitory phosphorylation at Thr356 has been described [85], which still has to be established in humans. The respective sites are targeted by several kinases upon stimulation with multiple factors [39,79]. Ser21 and/or Ser9, for instance, can be phosphorylated by protein kinases A (PKA) in response to cAMP-inducing agents (e.g., forskolin or isoproterenol) and cAMP analogues [86], PKB/Akt (induced by tumor necrosis factor (TNF) [87], interferon (IFN) β [88], growth factors, insulin [89], and lipopolysaccharide (LPS) [90]), and various protein kinase C (PKC) isoforms (including α, βII, γ, δ, and η [91], e.g., induced by TNF [92], phorbol esters, and via the Wnt (“wingless-type MMTV integration site family member”) pathway [93]). Moreover, these residues may be triggered by p38-MAPK (in response to IL-13) [94], serum- and glucocorticoid-induced protein kinase [95], and both subfamilies of ribosomal protein S6 kinase (p90-RSK and p70-S6K, induced by growth factors) [39]. Interestingly, a Ser21/9 phosphorylation-independent mechanism of GSK3 inactivation in response to Wnt signaling has also been described [96], presumably involving GSK3 localization and restriction of substrate accessibility [7,97]. GSK3β-Ser147 phosphorylation is induced by Wnt signaling via PKCζ [83], Thr43 phosphorylation by growth factors via ERK [84], and Thr390 in response to cellular stress or Wnt proteins via p38 [98].
Among the known growth factors, insulin-like growth factor 1 [99], platelet-derived growth factor BB [100], hepatocyte growth factor [101], fibroblast growth factor 2 [102], epidermal growth factor, and transforming growth factor (TGF-)β [103] as well as stem cell factor [104] have been demonstrated to inhibit GSK3 activity. In turn, activation of Ser21/9-phosphorylated GSK3 can be mediated via protein phosphatases (PP) 1, 2A [79,105], and 2B [15].

Acetylation, Ribosylation, SUMOylation, Citrullination, Ubiquitination, and Methylation

Additional post-translational modifications may further modify GSK3 functions and activity [1,13]. Lysine acetylation has been shown to have a negative regulatory influence on GSK3 kinase activity in general. For instance, acetylation of GSK3α-Lys246 and GSK3β-Lys183 appears to hinder ATP binding [106]. Accordingly, deacetylases of the sirtuin (SIRT) family contribute to the activation of GSK3β, e.g., via deacetylation of Lys183 by SIRT2 [106] or Lys205 by SIRT1 [107] and 3 [108]. ADP-ribosyltransferase ARTD10-mediated mono-ADP-ribosylation also leads to an inhibition of GSK3β activity [109], whereas removal of ADP-ribose by mono-ADP-ribosylhydrolase macrodomain protein MacroD2 restores its activity [110]. GSK3β Lys292-SUMOylation appears to be required for its stability, kinase activity, and nuclear localization [111] and peptidyl-deiminase 4-mediated citrullination also supports its nuclear translocation [112]. Using high-throughput approaches, further post-translational modifications have been found in both GSK3α and GSK3β, including ubiquitination at multiple sites [113] and methylation of GSK3α-Arg16 [114] or GSK3β-Lys27 [115].

Proteolytic Cleavage and Degradation

Under certain conditions, GSK3 activity can be enhanced by proteolytic cleavage of its N- and/or C-terminus [116,117]. The inhibitory effects of Ser21/Ser9 phosphorylation, for instance, can be countered by matrix metalloprotease (MMP-)2 [118] or calpain-driven N-terminal cleavage of GSK3α and β, generating remarkably truncated but enzymatically active protein variants (MMP-2: GSK3β: 30 kDa; calpain: GSK3α: 42 and 30 kDa, GSK3β: 40 and 30 kDa) [116]. Further analyses revealed that residues Thr38-Thr39 and Ile384-Gln385 are the relevant calpain cleavage sites in GSK3β resulting in the formation of the fragments ΔN-GSK3β (aa 39–420), ΔC-GSK3β (aa 1–384), and ΔN/ΔC-GSK3β (aa 39–384, i.e., the most active form). The remaining (non-catalytic) C-terminus, however, appears to be essential for GSK3 enzymatic activity, since it was demonstrated in deletion experiments that the absence of aa 417–430 in GSK3α or aa 345–367 in GSK3β yields a virtually complete loss of both Tyr279/216 auto-phosphorylation and substrate (e.g., Tau) phosphorylation [34].
Truncation is facilitated by dephosphorylation of Ser9 (mainly via phosphatase 1/2A; both N- and C-terminal) and Ser389 (only C-terminal) [117]. C-terminal GSK3β truncation, in turn, enhances its nuclear translocation and association with PP2A, thus promoting its activation via Ser9 dephosphorylation [119]. In the case of GSK3α, the N-terminal region further appears to be responsible for its fast and efficient export from the nucleus, since N-terminal binding of specific calcium/calpain-sensitive interaction partner(s) or deletion of the N-terminus resulted in nuclear accumulation of GSK3α [63]. For GSK3β, it has been shown that the presence of the N-terminal part is necessary for the interaction with certain proteins, e.g., PKB/Akt, p53, or the 14–3-3ζ adapter protein [120].
Degradation of GSK3 proteins is predominantly mediated via polyubiquitination and subsequent proteasomal destruction in a Ser9 phosphorylation-dependent manner [95].

3. Role of GSK3 in Inflammation and the Resolution of Inflammation

In this chapter, the contribution of GSK3 to the steering of both the initiation/maintenance of inflammation and the resolution of inflammation will be highlighted.

3.1. GSK3 and Inflammation

3.1.1. Cytokine Expression

GSK3 acts a potent driver of inflammation, rendering GSK3 inhibitors a promising target of anti-inflammatory research [17,121]. It is well established that enzymatically active GSK3 acts as a crucial positive regulator of pro-inflammatory cytokines (e.g., TNF, interleukin (IL-)1β, IL-6 [15], IL-17, IL-18 [122], IL-23 [94], IL-12, IFN-γ [17]), chemokines (IL-8 [122], C-C motif chemokine ligand (CCL) 2 [17], 3, 4 [123], and 12, C-X-C motif chemokine ligand (CXCL) 1, 2, 5 [17], and 10 [124]), and further pro-inflammatory mediators (e.g., nitric oxide (NO) [125] or prostaglandin E2 [126]). Vice versa, anti-inflammatory cytokines, such as IL-2 [127], IL-10 [90], IL-22 [128], IL-33 [129], and IL-1 receptor antagonist [130], are negatively regulated by GSK3. Interestingly, GSK3 has also been linked to anti-inflammatory action leading to the termination of inflammatory events (see Section 3.2.). As a reflection of the impact of GSK3β on inflammatory diseases, the application of the GSK3β activity index (i.e., the ratio of total to Ser9-phosphorylated GSK3β) has been proposed as a new diagnostic and prediction tool [131,132].

3.1.2. Animal Models

The role of GSK3 has been assessed in numerous animal (especially mouse) KO models, including both conventional and conditional approaches [12,20]. In addition, in a variety of inflammatory animal disease models, the influence of GSK3 has been studied [121].

GSK3 KO Models

GSK3α KO mice were viable, fertile, had normal body mass [133], and no obvious skeletal defects [134], but showed increased water uptake and urine production [135], enhanced sensitivity towards insulin and glucose, reduced fat mass, and increased hepatic glycogen levels [133]. In an alternative GSK3α KO model, mice were also viable but characterized by higher body weight, heavier organs (especially brain, heart, and testis), male infertility, and slightly reduced lifespan [136]. Interestingly, despite a multitude of indications that GSK3 has an essential influence on inflammatory processes [17,121], GSK3α KO organisms are not characterized by significant signs of altered immune reactions, implying that GSK3β may be the more prominent paralog with respect to immune regulation.
In contrast to GSK3α, GSK3β KO proved to be lethal in the late phase of murine embryogenesis [12]. The respective mice are characterized by a variety of cellular and organ defects, including liver degeneration induced by TNF-hypersensitivity-related apoptosis of hepatocytes [137] and severe cardiac defects resulting from impaired cardiomyocyte differentiation [138]. Since the premature death of KO individuals impedes the analysis of GSK3β functions, several conditional, tissue-specific GSK3β KO models have been generated. Though not predominantly characterized by alterations of the immune system or in host defense, some aspects of conditional KO may also have an impact on inflammatory processes (e.g., proliferation, differentiation, apoptosis, metabolism, or TF activation). A myeloid cell-specific GSK3β KO, for instance, led to liver tissue/hepatocellular protection as well as diminished pro- (TNF, IL-6, CXCL10, neutrophil infiltration/activation) and enhanced anti-inflammatory responses (IL-10) in a murine model of ischemia-reperfusion injuries [139]. GSK3β KO in the renal proximal tubule proved to be protective against mortality and tubular injury due to rapid tissue regeneration as reflected by reduced apoptosis in the renal cortex and accelerated proliferation of renal proximal tubule cells in a mercury chloride-induced acute nephrotoxic injury model [140]. In murine experimental adriamycin nephropathy [141] and oxidative glomerular injury [142], podocyte-specific GSK3β KO significantly decreased podocyte loss and injury, reduced glomerular damage, attenuated proteinuria [141,142], and reduced glomerular reactive oxygen species (ROS) production, the latter presumably due to an increase in the antioxidant genes, inducing nuclear factor erythroid 2-related factor 2 [142,143]. The respective podocytes were characterized by increased glycogen accumulation [142] as well as preserved cytoskeleton integrity and focal adhesions, reduced mitochondria dysfunction, and diminished pro-inflammatory nuclear factor (NF-)κB activation [141]. Anti-apoptotic effects of GSK3β deficiency have also been observed in murine GSK3β KO oligodendrocytes that are protected from caspase-dependent (but not -independent) apoptosis [144]. Moreover, in heterozygous GSK3β+/- mice exhibiting Pam3CSK4-induced peritonitis, the extent of inflammation was significantly lower than in the wildtype and equivalent results have been obtained in chimeric mice possessing GSK3β KO hematopoietic cells [145]. In summary, these data indicate that GSK3β indeed plays an essential role in controlling cellular functions contributing to initiation or resolution of inflammation (see also Section 3.2.).

Animal Models of Inflammatory Diseases

The influence of GSK3 on inflammatory events has also been studied in a number of animal models focusing on inflammatory diseases. For instance, in a murine collagen-induced arthritis (CIA) model, treatment with the GSK3 inhibitor TDZD-8 led to significantly reduced joint inflammation and destruction as reflected by reduced edema formation and histological joint alterations/bone resorption, decreased numbers of circulating leukocytes and infiltrating neutrophils as well as lower levels of TNF, IL-6, macrophage inflammatory protein (MIP) 1α and 2, inducible NO synthase (iNOS), and cyclooxygenase-2 (COX-2) [146]. In another murine CIA model, GSK3 inhibitors reduced hind paw erythema and swelling, pannus formation, infiltration of macrophages and T-cells, and bone erosion. Inhibitor-treated CIA mice also showed decreased pro-inflammatory cytokine production, e.g., TNF, IL-1β, IL-6, and IFN-γ [147]. In a collagen antibody-induced arthritis model, GSK3 inhibition using LiCl also reduced joint swelling [145]. Likewise, treatment with LiCl resulted in an amelioration of peritonitis in a murine Pam3CSK4-induced peritonitis model [145]. In mice with peptidoglycan (PGN)-induced peritonitis, anti-inflammatory effects (increased IL-10, decreased TNF, IL-1β, IL-6) could be elicited by ephedrine hydrochloride (EH), a substance that inhibits GSK3β via the PI3K-Akt axis in cell culture experiments [148] (see Section 3.1.3.). In a rat model of ischemic stroke, the application of GSK3 inhibitor VIII significantly reduced infarct volume, brain swelling, and neutrophil infiltration as well as signs of apoptosis (caspase-3 activation) and inflammation (COX-2 amounts; number of astrocytes, monocytes, and macrophages) [149]. In comparison to WT mice, protein tyrosine phosphatase receptor type O (PTPRO) KO mice are characterized by significantly attenuated inflammation (i.e., decreased amounts of TNF, IL-1β, IL-6, IFN-γ, CCL2, 3, and CXCL10; reduced numbers of infiltrating immune cells in the liver) following the induction of fulminant hepatitis using Concanavalin A injection. In that case, PTPRO KO was associated with increased PI3K/Akt activation and significantly elevated GSK3β-Ser9 phosphorylation [150].
Studying ischemic injury in an ischemia-reperfusion mouse model revealed that the respective WT mice exhibit significantly induced GSK3β activation (due to reduced Ser9 phosphorylation) together with strong inflammatory responses as reflected by enhanced leukocyte rolling/adhesion, numbers of circulating neutrophils, and TNF production. The injection of GSK3 inhibitor SB216763 during ischemia prevented the occurrence of inflammatory symptoms [151]. (Neuro)-inflammation is also a symptom observed upon hypoxic-ischemic injury. In a murine postnatal hypoxic-ischemic injury model, increased GSK3β activity and TNF/IL-6 expression were detected, which could be reduced by SB216763 [152].
Following induction of diabetes using streptozotocin, GSK3β activation by decreasing Ser9 phosphorylation was associated with increasing signs of inflammation (e.g., TNF, plasminogen activator inhibitor 1, and intracellular adhesion molecule 1 expression, 3-nitrotyrosine accumulation) in the liver of diabetic mice and (even more prominent) diabetic mice with zinc deficiency [153]. Under the same conditions, comparable results have been obtained in the liver of normal mice treated with a zinc chelator [153]. In contrast, SB216763-treated diabetic WT mice and cardiac-specific metallothionein-overexpressing transgenic mice (which are protected from diabetes-associated cardiomyopathy) showed neither GSK3β activation nor significant symptoms of cardiac inflammation [154]. In an alternative streptozotocin-induced diabetes model, elevated levels of GSK3β were accompanied with increased TNF, IL-1β, and IL-6 levels in the hippocampus. Pro-inflammatory cytokine expression could be ameliorated by application of Boswellia serrate extract, a polyphenol-rich substance also reducing hippocampal GSK3β expression [155]. Decreased GSK3β-Ser9 phosphorylation, increased TNF, IL-6, COX-2, and iNOS expression, and the abolishment of these effects by SB216763 can also be observed in the hippocampi of rats with diabetes induced by a combination of a high-fat diet and low streptozotocin concentrations [156].
An Alzheimer’s disease (AD) mouse model based on GSK3β overexpression is characterized by severe brain inflammation, e.g., increased numbers of activated microglia and enhanced TNF, IFN-γ, MIP-1α, -3α, and CCL2 (but also IL-10) expression [157]. In another AD model, in which the mice exhibit GSK3β hyperactivation and neuroinflammation, the application of tauroursodeoxycholic acid (an endogenous hydrophilic bile acid) led to the activation of Akt, increased GSK3β-Ser9 phosphorylation, reduced TNF expression, and decreased microglia activation [158].
These studies strongly suggest that active GSK3(β) is a potent driver of inflammation in vivo, whereas its inactivation has a mitigating influence. In consequence, the treatment with GSK3β-Ser9 phosphorylation-inducing substances, including a variety of natural products [21], generally dampens signs of (exaggerated) inflammation and tissue damage.

3.1.3. Role of GSK3 During Bacterial Infections

During bacterial infections, GSK3 enzymatic activity may be modulated via toll-like receptors (TLR) and subsequent PI3K-Akt activation, effects implying an inhibition of GSK3 in response to bacteria and their compounds [159]. The impact of TLR signaling on GSK3 activity, however, is ambiguous, difficult to predict, and presumably dependent on the specific prevailing conditions (e.g., the cell type and timeframe of observation). Thus, GSK3-dependent pro- as well as anti-inflammatory responses have been reported following TLR activation.
For instance, Akt-mediated GSK3 inactivation following stimulation of TLR2, 4, 5, or 9 with appropriate agonists (lipoteichoic acid, LPS or synthetic E. coli lipid A, flagellin, and human CpG, respectively) significantly suppressed pro-inflammatory cytokine secretion and induced (TLR2-dependent) IL-10 production in human monocytes in a CREB- and CREB-binding protein (CBP)-dependent manner [90]. In LPS-treated murine macrophage-like RAW264.7 cells and primary murine macrophages, preceding TLR2 stimulation by recombinant leucine-responsive regulatory protein preincubation results in PI3K/Akt activation, GSK3β-Ser9 phosphorylation, reduced NF-κB activity/nuclear translocation, and suppression of pro-inflammatory IL-6 and -12 expression [160]. In LPS-challenged human monocytes, it was shown that Akt-dependent GSK3 inactivation may be supported by additional mTORC2-dependent activation of Akt as well as (mTORC1-dependent) activation of GSK3β-Ser9-targeting S6K [161]. The application of GSK3 inhibitors protected mice from endotoxin shock [90] and enhanced the survival of Burkholderia pseudomallei-infected mice [162]. The increase in IL-10 via CREB due to Akt-mediated GSK3β inactivation has also been demonstrated in murine macrophages in response to Francisella tularensis (FT) [163] and in murine and human macrophages in response to Leishmania donovani (LD) infection [164]. Furthermore, increased IL-10 and decreased IL-6 production due to Ser9-dependent GSK3β inactivation has been observed in PGN-treated primary murine peritoneal macrophages and RAW cells following EH application [148]. GSK3β-Ser9 phosphorylation has also been detected in FT-infected murine macrophages [163] and LPS- or LD infection-challenged RAW cells [165].
Interestingly, a concomitant application of the LD-derived TLR4 agonist β-1,4-galactose terminal glycoprotein (GP29) enhanced GSK3β activity, resulting in reduced CREB and increased NF-κB-p65 and AP-1-Jun/Fos phosphorylation, decreased IL-10 expression, and induced IL-12 and NO synthesis in LD-infected RAW cells [165]. Vaccination with GP29 promotes a protective immune effect in a murine visceral leishmaniasis model by restricting IL-10 and increasing the production of pro-inflammatory cytokines (TNF, IL-12, and IFN-γ), NO, and ROS [166]. Moreover, a GSK3-dependent upregulation of TNF and NO in response to Streptococcus infections via TLR2 has been observed in murine macrophages [125] and microglia [167].

3.1.4. Role of GSK3 During Viral Infections

GSK3 also appears to be involved in the innate anti-viral immune response [159], though reports focusing on the effect of GSK3 on viral replication appear inconsistent and in part contradictory. This might be interpreted as different mechanistic approaches developed by viruses to subordinate the cellular machinery of the host or to escape the anti-viral activity of infected cells.
For instance, GSK3β has been identified as one of the host factors required for influenza A virus entry [168]. In human immunodeficiency virus (HIV-)1-infected T- and monocytic cell lines, upregulated GSK3β expression has been observed in both the cytoplasm and (to a lesser extent) the nucleus [169]. It has also been demonstrated that coxsackievirus B3 infection upregulated GSK3β levels in a mouse model [170] and increased the activity of GSK3β in HeLa cells, resulting in reduced β-catenin amounts, infection-associated cytopathic effects, and apoptosis as well as increased viral progeny release, effects that could be blocked by GSK3 inhibition [171]. Furthermore, GSK3β inhibition considerably reduces viral replication in HIV-1-infected macrophages or cell lines [169,172] and Varicella zoster virus-infected human MeWo melanoma cells [173]. In hepatitis C virus (HCV)-treated human hepatocarcinoma Huh7.5 cells, GSK3 inhibitors prevented the release of HCV virions, while viral replication was not affected [174]. Another group found that GSK3β (but not GSK3α) inhibition reduced both replication and viral particle production of HCV but not hepatitis E virus in Huh7.5 cells [175]. Moreover, GSK3-mediated phosphorylation appears to be involved in the destabilization of PHD finger protein 13, a host factor that (amongst other functions) is involved in repressing HIV-1 [176] and human cytomegalovirus gene expression [177]. Direct GSK3α/β-dependent phosphorylation of the coronavirus (CoV) nucelocapsid (N) has been described to be of importance for viral expansion, since GSK3 inhibition reduces both CoV-N phosphorylation and CoV replication [178].
On the other hand, the Karposi’s sarcoma-associated herpes virus-derived protein LANA is able to interact with GSK3β to favor nuclear GSK3β localization, and to increase both β-catenin [69,179] and Myc concentrations [180], which implies a LANA-dependent inactivation of GSK3β. Hepatitis B virus (HBV) replication is associated with increased GSK3β-Ser9 phosphorylation [181,182], indicating a beneficial effect of GSK3β inactivation for HBV replication. Mechanistically, anti-viral GSK3β effects may be associated in certain cases with its ability to enhance the anti-viral capacity of the zinc-finger anti-viral protein (an inhibitor of the replication of certain viruses) by sequential phosphorylation [183].

3.1.5. GSK3 and Interferons

The expression of interferons, i.e., cytokines with anti-viral properties (amongst others), also affects GSK3 activity. For instance, stimulation with IFN-β leads to a Jak1-PI3K-Akt1-driven significant reduction of GSK3β activity, an accumulation of activated CREB in the nucleus, and the subsequent induction of anti-inflammatory IL-10 in human dendritic cells (DC) [88]. In murine DC, the LPS-driven increase in IL-12 and IL-23 can be limited, while IL-10 production can be intensified by IFN-β preincubation, an effect that is associated with enhancement of Akt and GSK3β-Ser9 phosphorylation [184].
IFN-γ, in contrast, increases GSK3 activity as reflected by hyper-phosphorylation of typical GSK3 targets, such as Tau protein [185]. IFN-γ-induced activation of GSK3 facilitates iNOS expression [186], decreases IL-10 production by suppressing CREB and AP-1 transactivation activity [145], and enhances LPS-induced pro-inflammatory IL-6 expression via signal transducer and activator of transcription (STAT) 3 [187]. In this context, an association of enzymatically active GSK3β with the IFN-γ receptor has been demonstrated [187]. Equivalently, IFN-γ-induced activation of GSK3β suppressed IL-10 production, resulting in a subsequent upregulation of TNF and NO synthesis [188]. Interestingly, IFN-γ-dependent activation of GSK3β appears to involve both phosphatase-mediated dephosphorylation at Ser9 and Pyk2-mediated phosphorylation at Tyr216 [188,189]. IFN-γ-induced activation of GSK3β can be counteracted by Ser9 phosphorylation-inducing agents as shown for the immunosuppressant mycophenolate [190].

3.1.6. Clinical Application of GSK3 Inhibitors

Because of its remarkable impact on a variety of physiological key processes [191], GSK3 inhibition has been early regarded as a promising approach within the treatment of serious common diseases (also including inflammation or inflammatory components [17]), such as neurological [192] and cancerous [193] diseases. Though the initial high hopes have not been fully realized, GSK3 inhibitors are still under clinical investigation [194].
In several preclinical and clinical trials, the efficacy and safety of pharmacological GSK3 inhibitors for different clinical purposes are or have been addressed [194]. The most prominent GSK3 inhibiting agent is lithium, a non-selective metal cation whose precise inhibitory mechanism is still not fully elucidated. Lithium has been clinically applied for decades to patients with mood and bipolar disorders [195]. Thus, it is a well-characterized prevalent drug exhibiting an acceptable level of adverse effects [196], though in some elderly patients, toxic effects have been described [195]. Currently, the effect of lithium on Parkinson’s disease (NCT04273932, phase I), cognitive impairment (NCT03185208, phase IV) [197], and fracture healing (NCT02999022, phase II) [198] is being assessed.
Other common inhibitors studied are ATP analogs acting as reversible ATP competitors [195], such as LY2090314, a potent and relatively GSK3-selective inhibitor. Following promising preclinical and phase I clinical results in terms of pharmacokinetics, metabolism, excretion [199], and efficacy (in combination with pemetrexed and carboplatin) against advanced solid tumors [200], a phase II study showed no significant clinical effects of LY2090314 as a single agent in acute myeloid leukemia treatment, thus limiting its benefit [201]. Other components, such as AZD1080, were analyzed in phase I trials [202] but discontinued in development [195]. Chiefly, problems with this type of drug may be attributed to two major issues, i.e., a common lack of specificity combined with a frequently occurring toxicity [195]. However, even ATP competitors are still regarded as potential pharmaceutical agents. 9-ING-41, for instance, is a malmeimide-derived GSK3 inhibitor that initially showed potent anti-proliferative activity [203] and capacity to overcome chemoresistance [204] in different cancer cell lines. At the moment, the safety and efficacy of 9-ING-41 are being addressed in patients with hematologic malignancies and solid tumors in a phase I/II study (NCT03678883) [197,205].
Tideglusib, a thiadiazolidindione, is an oral administrable, non-ATP competitive inhibitor characterized by an irreversible inhibition of GSK3 [206]. Its therapeutic impact on AD [207] and progressive supranuclear palsy (PSP) [208] was assessed in a series of studies. Tideglusib was generally well tolerated, showed encouraging results in an AD pilot study [209], and appeared to reduce cerebral atrophy in a PSP subgroup in a phase II clinical trial [210]. However, no significant improvement in the Tideglusib-treated AD or PSP groups was observed in phase II clinical trials [207,208]. Following the completion of a phase II study (NCT02858908) assessing the safety, pharmacokinetics, and efficacy of Tideglusib in patients with myotonic dystrophy in 2018, a clinical phase II/III trial (NCT03692312) was announced for 2020 [197].

3.1.7. Regulation of Transcriptional Systems and Associated Signaling Pathways

On the transcriptional level, GSK3-regulated effects are mediated via the activation or inactivation of prominent signaling pathways and TF, in particular NF-κB [137], AP-1 [211,212,213], and C/EBP [214], which are described below in detail. However, various further TF, such as STAT1 [189], 3, and 5 [215], or CREB [216], also contribute to the transmission of GSK3-dependent signaling.

NF-κB-associated signaling

NF-κB, a TF consisting of the subunits p105/p50 (i.e., NFKB1), p65 (RelA), and c-Rel (representing the canonical NF-κB pathway) as well as p100/p52 (NFKB2) and RelB (the non-canonical pathway) [217], is one of the most important transcriptional mediators transferring GSK3 activity into gene expression [218]. Initially, it has been demonstrated that TNF-induced NF-κB DNA binding activity and transactivation of NF-κB-dependent promoter constructs were significantly reduced in murine GSK3β KO fibroblasts, effects that could be reversed by transient transfection of rat GSK3β. Interestingly, under these circumstances, IκB degradation and nuclear p65 translocation were not affected [137]. GSK3 inhibition [219] or genetic deletion [220] also had no effect on IκBα degradation, IKK activation, or p65 translocation and it has been proposed that direct p65 phosphorylation may play a decisive role in that context (see below) [219]. In addition, during GSK3 inactivation, the increasing β-catenin levels appear to further enhance NF-κB suppression, since β-catenin has been shown to directly interact with protein complexes that include p65- and/or p50-containing NF-κB dimers and to reduce NF-κB DNA binding and subsequent functions (i.e., transactivation activity and target gene expression) in this way [221].
The constitutive DNA binding and transactivation activities of p65/p50 heterodimers observed in pancreatic cancer cells are positively regulated by both GSK3α and β, since GSK3 inhibition (including paralog-specific siRNA) results in reduced p65 phosphorylation, DNA binding, and NF-κB-dependent gene expression, while p50 homodimers seem to be unaffected. This mechanism appears to involve GSK3-dependent constitutive IKK activation and IκBα phosphorylation [222]. In TNF-stimulated WT endothelial cells, a rapid transient (i.e., 5–10 min following stimulation) activation of GSK3 has been observed, an effect associated with IKKβ activation, IκBα degradation, enhanced DNA binding activity of NF-κB, and the formation of a pro-inflammatory phenotype [223]. The influence of long-term TNF incubation on GSK3 activity, however, has not been comprehensively addressed yet (see also Section 3.2.4.). Other studies demonstrated that GSK3β activity contributes to the activation of several NF-κB-dependent genes following TNF stimulation by enhancing p65 DNA binding activity, while specific subsets of NF-κB-dependent genes (including IκBα) are not affected [220,224]. Vice versa, at the promoters of certain NF-κB-dependent genes that are suppressed under GSK3β inhibiting conditions, either decreased p65 or increased p50 amounts can be detected [224]. This suggests the existence of specific and differentially regulated NF-κB-dependent gene sets and a distinct role for GSK3β in the modulation of the NF-κB system(s). This may also be reflected by the identification of a set of growth factor-responsive genes that are suppressed in quiescent cells by GSK3-dependent inhibition of IKK and IκBα phosphorylation, whereas GSK3 inhibition (either by inhibitors or growth factor stimulation) leads to increased IKK activity, IκBα phosphorylation/degradation, nuclear p65/p50 translocation, and binding of p65 to the promoters of several (though not all) of these genes [225]. Thus, GSK3 activation appears to involve an activation of NF-κB and associated signaling molecules in general. However, since single studies report that overexpression of constitutively active GSK3β variants significantly reduced stimulus-induced IKK activation, IκBα degradation, NF-κB DNA binding activity, and/or NF-κB-dependent transcription [226,227], an excess of enzymatically active GSK3β also appears to be disadvantageous for the responsiveness of the NF-κB system.
Most of the NF-κB subunits also act as substrates of GSK3. GSK3-dependent phosphorylation has been described for p65 [219] at Thr254 [228], Ser276 [229,230], Ser468 [231,232] (or its murine equivalent Ser467 [230]) and Ser536 [222,233], p105 at Ser903 and 907 [234], p100 at Ser707 [235], and RelB at Ser552 (in the mouse; corresponds to human RelB-Ser573 [236]) [237]. Various functional consequences arise from GSK3-driven NF-κB phosphorylation. Phosphorylation of p65-Thr254 appears to be essential for the transactivation of certain genes as shown for type II collagen expression during chondrocyte differentiation [228]. An increase in GSK3β-Tyr216 phosphorylation and the Tyr216/Ser9 ratio was associated with enhanced p65-Ser276 phosphorylation, p65-CBP interaction, and pro-inflammatory cytokine expression (TNF, IL-1β) [229]. In unstimulated cells, phosphorylation of p65 at Ser468 inhibits its transactivation activity and reactivation may be performed by PP1-mediated dephosphorylation [231]. In stimulated cells, however, p65-Ser468 phosphorylation results in the increased expression of a subset of NF-κB-dependent pro-inflammatory genes, including monocyte chemoattractant protein-1 [230,232]. EGF-stimulated epithelial cells show an activation of the canonical NF-κB pathway as reflected by increased p65-Ser536 phosphorylation due to the activation of both IKKα and GSK3β [233].
The major effect of GSK3-dependent phosphorylation of p105 and p100 is the modulation of protein stability and precursor digestion. While the interaction with GSK3β stabilizes p105 in resting cells by reducing the rate of constitutive p105 to p50 degradation, GSK3β-mediated phosphorylation of p105 at Ser903 and Ser907 primes p105 for subsequent proteasomal degradation upon TNF stimulation [234]. In turn, formation of p50 from its precursor p105 by limited proteolysis is facilitated in the absence of GSK3 [234].
GSK3-dependent constitutive phosphorylation of p100 at Ser707 in the nucleus enables its association with the F-box protein Fbxw7α-containing ubiquitin ligase complex, resulting in its subsequent ubiquitination and proteasomal degradation. As a consequence, binding and inhibition of other NF-κB subunits (e.g., p52 and RelB) by p100 is abolished. Thus, the efficient clearance of nuclear p100 is a prerequisite for proper activation of the non-canonical NF-κB signaling pathway [235]. The proteasome-dependent generation of p52 from p100, however, appears to be performed in a GSK3-independent manner [235].
Following suitable stimulation, Ser552 phosphorylation of RelB by GSK3β leads to its efficient degradation, which can be prohibited using SB216763, siRNA, or enzymatically inactive GSK3β variants [237].
GSK3 also affects further factors of the NF-κB system. The stability of Bcl-3 (“B cell lymphoma 3”), an important interaction partner of p50 or p52 homodimers [238], is also negatively regulated by GSK3 via phosphorylation at Ser394 and Ser398, which induces subsequent ubiquitination and proteasomal degradation [239]. Moreover, binding and phosphorylation (especially at Ser8, 17, and 31) of NF-κB essential modifier (NEMO, aka IKKγ) by GSK3β is required for its stabilization and the proper mediation of NF-κB signaling. Artificial destabilization of NEMO by replacing the relevant GSK3 target residues by alanine is accompanied by increased Lys63-polyubiquitination of the remaining NEMO molecules, enhanced IKKα and β binding, and elevated constitutive IκBα degradation, effects that could be further intensified by GSK3 inhibition [240].

AP-1-Associated Signaling

The regulation of AP-1 is a crucial step in mediating the expression of a variety of GSK3-dependent genes, making AP-1 an important player within the GSK3-regulated transcriptional network [218]. In general, activation and activity of AP-1 and its subunits Jun, JunB, JunD, Fos, FosB, Fos-related antigen (Fra-)1, and Fra-2 are inhibited by active GSK3. Early, it was shown that GSK3 is able to phosphorylate the DNA binding domain within the C-terminus of Jun in vitro, thus reducing the DNA binding activity of AP-1 consisting of Jun homodimers [211], whereas PKC (i.e., PKCα, β1, β2, or γ)-mediated GSK3β inactivation reduces the negative regulatory Jun phosphorylation [91]. Once phosphorylated by GSK3, Jun family proteins may be ubiquitinated and subsequently degraded via the proteasomal pathway as shown for Jun [241] and JunB [242]. For JunB and D, a similar GSK3-driven regulation of DNA binding activity has been demonstrated [212,213]. However, since JunB/D and Fra-1/2 possess relatively weak transactivation domains, which may render them to AP-1 subunits with predominantly deactivating or repressing features, GSK3-driven regulation of AP-1 is often mediated via its activating subunits, such as Jun [218]. A good example for the regulatory interrelationship during inflammation is the repression of IL-10 via the inhibition of AP-1 (and CREB) following IFN-γ stimulation of primary human macrophages. In detail, IFN-γ induced the activation of GSK3, which was associated with a significant suppression of both basal and TLR2-induced fos mRNA and Jun protein levels, lower AP-1 DNA binding activity, and reduced expression of AP-1-dependent genes, such as IL-10 [145]. An alternative possibility to negatively regulate Jun/AP-1 activity is the suppression of JNK activating signal transduction as shown in drosophila [243].
Vice versa, the inhibition of GSK3 either using inhibitors [244] or in response to physiological stimulation, such as integrin receptor binding [245] or Fcε receptor I crosslinking [246], leads to an activation of AP-1. Under these circumstances, reduced phosphorylation and increased protein stability of Jun can be observed [244]. Moreover, Jun, Fra-1 [247], and Fos [248] have been shown to be induced in the presence of inactivated GSK3β or increased β-catenin levels. In diseases characterized by augmented amounts of Ser9-phosphorylated GSK3β levels, such as cholestatic liver disease, Jun levels are also increased [249]. However, in some cases, opposed effects have been observed. In murine BV-2 microglial cells, pretreatment with GSK3 inhibitor TWS119 reduced the LPS-dependent activating phosphorylation of JNK and Jun, AP-1 DNA binding activity, and AP-1-dependent reporter gene activation [250].

C/EBP-Associated Signaling

The C/EBP family (consisting of C/EBPα, β, γ, δ, ε, and ζ [44]) is also able to mediate GSK3-derived signals [214]. C/EBPα was early identified as a GSK3 substrate phosphorylated at Thr222 and Thr226 (corresponding to human Thr226/230 [236]), i.e., residues that are dephosphorylated by PP1 or 2a during GSK3 inhibition, e.g., in response to insulin [251]. Accordingly, in transgenic mice expressing inactivation-resistant GSK3α-Ser21Ala and GSK3β-Ser9Ala variants, C/EBPα-Thr222/226 phosphorylation was significantly enhanced, while total C/EBPα levels were comparable with the WT [252]. Furthermore, it has been described that GSK3 activity is able to induce conformational changes in C/EBPα [251]. However, since reduced Thr222/226 phosphorylation could not be detected in insulin-treated hepatocytes, this effect appears to be cell type specific [253]. Moreover, the functional role of these phospho-sites seems to be situation dependent, since no concrete regulatory influence has been observed by in vitro serine-to-alanine mutation in adipocytes [253], whereas these aa exchanges suppressed the induction of metallothionein 2A in murine hepatocytes [254].
An equivalent dephosphorylation following growth hormone-induced GSK3 inactivation has been observed for C/EBPβ and its isoforms liver-enriched activating protein (LAP, a TF possessing transactivating capacity) and liver-enriched inhibitory protein (LIP, an inhibitory factor lacking transactivation domains) in murine fibroblasts. In this case, interestingly, LAP DNA binding activity was increased, while LIP binding was decreased, at least at certain C/EBP binding sites, e.g., within the cfos promoter, an effect associated with increased promoter activity [255]. Thus, C/EBPβ may contribute to the activation of AP-1 under GSK3-inhibiting conditions (see Section 3.1.7.). In murine embryonic fibroblasts and macrophage-like cells, another phosphorylation site at Thr188 (human: Thr235 [236]) has been described that was required for the activation of certain genes, including C/EBPα, and is sensitive towards GSK3 inhibition [256,257]. Following stimulation of RAW cells with edema toxin (ET), the phosphorylation of C/EBPβ-Thr188 was enhanced and the association of both GSK3β and C/EBPβ with the tumor suppressor APC (“adenomatous polyposis coli”, a subunit of the β-catenin destruction complex [8]) was induced, suggesting that APC may facilitate GSK3-driven C/EBPβ-Thr188 phosphorylation [257]. Further, MAPK-phosphorylated Thr188 may act as a priming site for GSK3β-mediated phosphorylation of Ser184 and Ser179, i.e., phospho-residues favoring dimerization due to conformational changes and enhancing both DNA binding and transactivation activity [40,258]. Decreased C/EBPβ-Thr235 phosphorylation due to GSK3 inactivation was detected in TNF long-term-incubated monocytic cells in the presence of kenpaullone [259] as well as human monocyte-derived DC in response to platelet-activating factor [260]. In contrast, in IL-17-stimulated murine ST2 stromal cells, ERK-mediated priming phosphorylation at Thr188 and subsequent phosphorylation at Ser179 by GSK3 have been described as being repressive for certain C/EBP-dependent genes [261]. Equivalent results have been obtained in human umbilical vein endothelial cells (HUVEC) [262]. In differentiated murine preadipocytic 3T3-L1 cells, an O-linked β-N-acetylglucosamine modification at Ser180 or Ser181 prevented C/EBPβ-Thr188, -Ser184, and -Thr179 phosphorylation and decreased its DNA binding and transactivation activities [263]. In rat-derived osteoblasts, a corresponding GSK3β-dependent phosphorylation at C/EBPβ-Thr189, -Ser185, -Ser181, and -Ser177 was described. In this case, however, stimulation-induced and cGMP-dependent protein kinase (PKG)-dependent GSK3β inhibition was associated with C/EBPβ dephosphorylation, increased DNA binding activity, and enhanced C/EBPβ-dependent reporter gene expression [264], suggesting a differential regulation in the rat when compared to mice and humans. In murine microglial cells, LPS-induced nuclear translocation of C/EBPβ and δ was strongly enhanced in the presence of LiCl [265]. Moreover, murine C/EBPδ has been reported to be phosphorylated by GSK3β at Ser167 in response to pro-inflammatory cytokines, such as TNF and IL-1β, leading to enhanced C/EBPδ-dependent gene expression [266].
Under specific conditions, levels of C/EBP proteins appear to be (differentially) influenced by GSK3 activity. In prostate cancer cells, for instance, normally low C/EBPα protein (but not mRNA) levels massively increased in the presence of GSK3 inhibitors, while the higher C/EBPβ levels remained unchanged, suggesting that C/EBPα protein stability may be affected by GSK3 [267]. Inhibition of GSK3 activity by melatonin treatment led to downregulation of C/EBPβ and δ expression in human mesenchymal stem cells [268]. In RAW cells, C/EBPδ has been shown to be targeted for degradation by Fbxw7α following phosphorylation by GSK3β at Thr156 and (to a lesser extent) Ser160. GSK3β inhibition, in turn, either by inhibitors or LPS-induced Ser9 phosphorylation, enhanced C/EBPδ protein half-life [269]. In hepatocellular carcinoma cells, however, CEBPD promoter activation and C/EBPδ levels decreased in the presence of enhanced β-catenin amounts following GSK3 inactivation [270]. C/EBPζ, aka C/EBP homologous protein (acting as a pro-apoptotic TF), has been described to be upregulated in situations favoring GSK3α/β activation, such as endoplasmic reticulum (ER) stress [271]. Accordingly, expression of the constitutively active GSK3β-Ser9Ala variant induced C/EBPζ mRNA and protein levels in primary murine macrophages even under conditions antagonizing ER stress [272]

3.2. GSK3 and the Resolution of Inflammation

Negative modulation of GSK3 activity appears to be a crucial step during the regulatory events governing the resolution of inflammation, including the expression of pro-resolving cytokine profiles, clearance of apoptotic immune cells, and tissue repair/wound healing. However, under certain circumstances, GSK3 inhibition may also prohibit the termination of inflammation. In the following, the respective GSK3-associated mechanisms are discussed.

3.2.1. GSK3 and the Expression of Pro-Resolving Cytokine Profiles

Data from a study focusing on spontaneous resolution of inflammation in a murine trinitrobenzene sulfonic acid-induced colitis model revealed that colitis-induced secretion of IL-13 led to the inactivation (i.e., Ser9 phosphorylation) of GSK3β via the STAT6-driven activation of p38. The following increase in nuclear CREB resulted in decreased pro-inflammatory NF-κB-p65 DNA binding as well as reduced IL-17 and IL-23 but increased IL-10 expression, i.e., a modified cytokine expression profile favoring amelioration of inflammation [94]. In THP-1 cells challenged with varying doses of LPS, GSK3β-dependent cytokine expression profiles were differentially modified, suggesting a dose-dependent impact of LPS on the programming of innate immune cells and the resolution of inflammation. In this model, very low LPS doses yielded decreased Akt activity and increased Pyk2-mediated Tyr216 phosphorylation (i.e., activation) of GSK3β, resulting in elevated expression levels of pro-inflammatory IL-6. At the transcriptional level, an upregulation of TF forkhead box protein O1 (FoxO1) and a downregulation of CREB protein amounts contribute to the realization of this effect. Following high dose treatment, THP-1 cells showed increased Akt activity (implying GSK3 inhibition), upregulated CREB and downregulated FoxO1 levels (including activating CREB-Ser133 and inhibiting FoxO1-Ser256 phosphorylation), and elevated levels of both IL-6 and anti-inflammatory IL-33 [129].
Interestingly, resolution of inflammation also comprises effects that are rather regarded as pro-inflammatory. For instance, the proper production of ROS has been implicated to be of importance, since p47phox KO resulted in ROS deficiency-associated excessive acute inflammation in different mouse models (including p47phox KO mice and macrophages) following LPS challenge. Mechanistically, this can be attributed to decreased IL-10 levels, which are caused by significantly increased GSK3β activation due to reduced PI3K-Akt-dependent GSK3β-Ser9 phosphorylation in the absence of p47phox [273]. Thus, GSK3 inhibition generally appears to support the resolution of inflammatory processes. Consequently, the application of GSK3 inhibitors (e.g., SB216763, SB415286, TDZD-8 [274], indirubin-3′-monoxime [275], or lithium [145,276]) alleviates symptoms of systemic inflammation in murine and rat inflammatory disease models [121] (see Section 3.1.2.) and equivalent results have been obtained in human cell culture approaches [223].

3.2.2. GSK3 and the Clearance of Apoptotic Immune Cells

Clearance of apoptotic immune cells by macrophage-mediated efferocytosis is a key feature of resolving inflammation [277]. In PMA-differentiated human premonocytic THP-1 cells, treatment with the anti-inflammatory agent lipoxin A4 leads to increased phagocytosis of apoptotic polymorphonuclear neutrophils and lymphocytes in the dependence of PKCζ-mediated GSK3β-Ser9 phosphorylation [278]. Using primary human monocyte-derived macrophages, another study showed that during efferocytosis, GSK3β inhibition may be supported by miR-21 expression via the miR-21-induced silencing of phosphatase PTEN (“phosphatase and tensin homolog”), since this effect enables the activation of PI3K and GSK3-Ser9 phosphorylation [279].

3.2.3. GSK3 and Tissue Repair

In an in vitro airway injury and repair model, it was demonstrated that scratching-induced injury of a bronchial epithelial cell monolayer caused (PKC-dependent) GSK3β-Ser9 phosphorylation, accumulation and nuclear translocation of β-catenin, and increased cyclin D1 expression. These pro-proliferative events supported tissue repair, a process also contributing to the termination of inflammation [280]. An equivalent regulation of GSK3 has been observed in a wounding model using IEC-18 rat epithelial cells [281]. Consistently, wound healing was supported by overexpression of β-catenin or substances inducing GSK3β-Ser9 phosphorylation, such as lucidone [282], while GSK3β overexpression or PKC inhibition had the opposite effect [280]. Interestingly, the complete absence of GSK3β seems to have an adverse effect, since GSK3β KO murine embryonic fibroblasts exhibited reduced restitution in response to scrape wounding [281].

3.2.4. Inhibition of GSK3 and the Perpetuation of Inflammation

In certain situations, GSK3β inactivation has proven to counteract the termination of inflammation. The elimination of neutrophils, for instance, which represents another key module of fading inflammation [283], is significantly impaired in patients with severe injuries showing resistance towards intrinsic neutrophil apoptosis [284]. At the molecular level, reduced GSK3β enzymatic activity prevents the phosphorylation and subsequent ubiquitination/degradation of the anti-apoptotic protein Mcl-1 (“myeloid cell leukemia 1”). Thus, the stabilization and accumulation of Mcl-1 results in the suppression of pro-apoptotic events and a prolonged neutrophil lifespan [284]. Another example is the IL-17A-induced promotion of pulmonary fibrosis. Here, IL-17A stimulation results in PI3K-dependent inactivation of GSK3β, inhibition of GSK3β-Bcl-2 binding, reduced Bcl-2 phosphorylation and ubiquitination, and increased Bcl-2 levels in murine MLE-12 alveolar epithelial cells. Pathophysiologically, this leads to reduced autophagy and retarded resolution of inflammation, which enhances pulmonary fibrosis [285].
The development of different forms of tolerance towards the sustained stimulation with various activating agents is also regarded as an event contributing to the resolution of inflammation. In this context, the term tolerance describes a physiological phenomenon in which a pretreatment of organisms or specific cell types (e.g., monocyte/macrophages or hepatocytes) with a specific activating stimulus over a certain time leads to the resistance towards further stimulation with the same (or, in the case of cross-tolerance, another) substance. In its pure form, tolerance is mostly expressed as TNF, LPS, or TNF/LPS (cross-)tolerance [286] and reflected by the repression of immunologically relevant genes, such as IL-8 [287] or Il-6 [288]. TNF tolerance may occur in two forms, i.e., absolute tolerance (in which gene expression is blocked) and induction tolerance (in which gene expression is not further inducible). Low-dose preincubation predominantly results in absolute tolerance, whereas high-dose preincubation induces both absolute and induction tolerance [289]. Amongst other mechanisms, the development of low-dose TNF tolerance as well as TNF-induced cross-tolerance towards LPS depends on the activity of GSK3 [286], since GSK3 remained in an active state during TNF preincubation [288], while pharmacological inhibition or genetic deletion of GSK3 are able to reverse the formation of these forms of tolerance [288,289]. Mechanistically, GSK3 inhibition appears to counteract several molecular effects observed in low-dose tolerized cells, especially the increased phosphorylation of NF-κB-p65 at the inhibiting site Ser468 and the reduced phosphorylation at the activating site Ser536 [289]. The latter can be traced to direct association of C/EBPβ with p65 in tolerant cells [290]. It has been further shown that a GSK3 inhibitor can suppress the rapid resynthesis and nuclear accumulation of IκBα that occurs during LPS stimulation following TNF preincubation, suggesting that GSK3 is involved in the IκBα-mediated restriction of NF-κB signaling under these conditions [288]. In addition, GSK3 inhibition in TNF-tolerized cells also results in increased binding of p65 at the promoters of tolerizable genes (e.g., IL-6), reduced expression of A20 (a negative regulator of NF-κB signaling, e.g., during TNF incubation [291]), and enhanced chromatin accessibility [288]. Recently, it has further been demonstrated that numerous proteins differentially phosphorylated under conditions inducing high-dose TNF tolerance are potential GSK3 targets [259]. However, further analyses have to elucidate whether additional GSK3-dependent mechanisms are involved in the complex regulation of tolerance formation and the termination of pro-inflammatory signaling.

4. Concluding Remarks

As shown in this review, GSK3 is a versatile kinase targeting a variety of substrates. Thus, GSK3 controls a plethora of cellular functions at various mechanistic levels by orchestrating numerous molecular events. With respect to inflammation, GSK3 acts as a potent and multifunctional regulator of both pro- and anti-inflammatory effects. Among both paralogs, GSK3β appears to be the more influential kinase in terms of imunoregulatory action. Its specific effect, however, appears to depend on the respective cellular and molecular conditions. In general, active GSK3 involves a strong pro-inflammatory component by supporting the execution of multifaceted pro-inflammatory events. Conversely, GSK3 inhibition, either following physiological Ser9 phosphorylation or via pharmacological approaches, predominantly contributes to the amelioration of inflammation (as summarized in Table 1). The latter also includes the termination of pro-inflammatory signaling and the induction of (cross-)tolerance. Remarkably, several exceptions from this principle have been described. This janiform and to some extent contradictory nature is part of the mystery surrounding GSK3. During inflammation, modulation of GSK3(β) activity may affect multiple regulatory levels. These include enzymatic activity, signal transduction, transcription and translation, as well as localization and stability. Thus, due to its two-faced nature and its eclectic regulatory possibilities, GSK3 can be regarded as a key factor steering and balancing the initiation, progress, and resolution of inflammation. In consequence, a better understanding of its complex modulatory role during inflammatory processes may represent a promising opportunity to specify and modify different stages of inflammatory diseases.

Author Contributions

Conceptualization, K.B. and R.H.; Literature Research, L.H., M.D., and R.H.; Writing, Original Draft Preparation, R.H.; Writing, Review and Editing, L.H., M.D., K.B., and R.H. All authors have read and agreed to the published version of the manuscript.

Funding

L.H. was supported by the Hannover Biomedical Research School (HBRS) and the MD/PhD program Molecular Medicine. This research was funded by the German Research Foundation (DFG; grant SFB 566/B17 to K.B.) and the Deutsche Gesellschaft für Klinische Chemie und Laboratoriumsmedizin (DGKL; Stiftung für Pathobiochemie und Molekulare Diagnostik; grant 2012/20 to K.B.). We acknowledge support by the DFG and the Open Access Publication Fund of Hannover Medical School (MHH).

Acknowledgments

We apologize to all authors/working groups whose work has not been considered for this review due to space limitations or oversight.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Abbr.Terms
aaamino acid
ADAlzheimer’s Disease
AMPKAMP-activated protein kinase
APactivator protein
APCadenomatous polyposis coli
BclB cell lymphoma
BMbone marrow
β-TrCPF-box containing E3 ligase β-transducin repeat-containing protein
C/EBPCCAAT/enhancer binding protein
cAMPcyclic adenosine monophosphate
CBPCREB binding protein
CCLC-C-motif chemokine ligand
CIAcollagen-induced arthritis
CoVCoronavirus
CoV-NCoronavirus nucelocapsid
COXcyclooxygenase
CREBcAMP-responsive element binding protein
CXCLC-X-C-motif chemokine ligand
EHephedrine hydrochloride
ERendoplasmic reticulum
ERKextracellular signal-regulated kinase
ETedema toxin
Foxforkhead box protein
FraFos-related antigen
FTFrancisella tularensis
GP29β-1,4-galactose terminal glycoprotein
GSglycogen synthase
GSKglycogen synthase kinase
HBVHepatitis B virus
HCVHepatitis C virus
HIVHuman immunodeficiency virus
HUVEChuman umbilical vein endothelial cells
Hsp90heat shock protein 90
IFNinterferon
IKKIκB kinase
ILinterleukin
iNOSinducible NO synthase
IRFinterferon response factor
IκBinhibitor of κB
JNKJun N-terminal kinase
KOknock-out
LANAlatency-associated nuclear antigen
LAPliver-enriched activating protein
LDLeishmania donovani
LIPliver-enriched inhibitory protein
LPSlipopolysaccharide
MAPKmitogen-activated protein kinase
Mclmyeloid cell leukemia
MIPmacrophage inflammatory protein
MMPmatrix metalloprotease
mTORCmechanistic target of rapamycin complex
Mybmyoblastosis
Mycavian myelocytomatosis viral oncogene homolog (Myc)
NEMONF-κB essential modifier
NF-κBnuclear factor κB
NLSnuclear localization sequence
NOnitric oxide
P70-S6Kp70 ribosomal protein S6 kinase
P90-RSKp90 ribosomal protein S6 kinase
PGNpeptidoglycan
PI3Kphosphatidylinositol-3-kinase
PKAprotein kinase A
PKB/Aktprotein kinase B
PKCprotein kinase C
PKGcGMP-dependent protein kinase
PLINperilipin
PPprotein phosphatase
PSPprogressive supranuclear palsy
PTENphosphatase and tensin homolog
PTPROprotein tyrosine phosphatase receptor type O
ROSreactive oxygen species
SIRTsirtuin
STATsignal transducer and activator of transcription
TBKTANK-binding kinase
TFtranscription factor
TGFtransforming growth factor
TLRtoll-like receptor
TNFtumor necrosis factor
TstTuberculin skin test reactivity, absence of
UTRuntranslated region
Wntwingless-type MMTV integration site family member
WTwildtype

References

  1. Cormier, K.; Woodgett, J.R. Recent advances in understanding the cellular roles of GSK-3. F1000Research 2017, 6, 167. [Google Scholar] [CrossRef] [Green Version]
  2. Shinde, M.Y.; Sidoli, S.; Kulej, K.; Mallory, M.J.; Radens, C.; Reicherter, A.L.; Myers, R.L.; Barash, Y.; Lynch, K.W.; Garcia, B.A.; et al. Phosphoproteomics reveals that glycogen synthase kinase-3 phosphorylates multiple splicing factors and is associated with alternative splicing. J. Biol. Chem. 2017, 292, 18240–18255. [Google Scholar] [CrossRef] [Green Version]
  3. Liu, X.; Klein, P.S. Glycogen synthase kinase-3 and alternative splicing. Wiley Interdiscip. Rev. RNA 2018, 9, e1501. [Google Scholar] [CrossRef]
  4. Yokoo, H.; Nemoto, T.; Yanagita, T.; Satoh, S.; Yoshikawa, N.; Maruta, T.; Wada, A. Glycogen synthase kinase-3β: Homologous regulation of cell surface insulin receptor level via controlling insulin receptor mRNA stability in adrenal chromaffin cells. J. Neurochem. 2007, 103, 1883–1896. [Google Scholar] [CrossRef]
  5. Welsh, G.I.; Miyamoto, S.; Price, N.T.; Safer, B.; Proud, C.G. T-cell Activation Leads to Rapid Stimulation of Translation Initiation Factor eIF2B and Inactivation of Glycogen Synthase Kinase-3. J. Biol. Chem. 1996, 271, 11410–11413. [Google Scholar] [CrossRef] [Green Version]
  6. Cohen, P.; Frame, S. The renaissance of GSK3. Nat. Rev. Mol. Cell Biol. 2001, 2, 769–776. [Google Scholar] [CrossRef]
  7. Robertson, H.; Hayes, J.D.; Sutherland, C. A partnership with the proteasome; the destructive nature of GSK3. Biochem. Pharmacol. 2018, 147, 77–92. [Google Scholar] [CrossRef]
  8. Stamos, J.L.; Weis, W.I. The beta-catenin destruction complex. Cold Spring Harb Perspect. Biol. 2013, 5, a007898. [Google Scholar] [CrossRef]
  9. Gu, Y.; Gao, L.; Han, Q.; Li, A.; Yu, H.; Liu, D.; Pang, Q. GSK-3beta at the Crossroads in Regulating Protein Synthesis and Lipid Deposition in Zebrafish. Cells 2019, 8, 205. [Google Scholar]
  10. Wang, L.; Liu, X.; Zhan, S.; Guo, J.; Yang, S.; Zhong, T.; Li, L.; Zhang, H.; Wang, Y. Inhibition of GSK3beta Reduces Ectopic Lipid Accumulation and Induces Autophagy by the AMPK Pathway in Goat Muscle Satellite Cells. Cells 2019, 8, 1378. [Google Scholar] [CrossRef] [Green Version]
  11. Yang, K.; Chen, Z.; Gao, J.; Shi, W.; Li, L.; Jiang, S.; Hu, H.; Liu, Z.; Xu, D.; Wu, L. The Key Roles of GSK-3beta in Regulating Mitochondrial Activity. Cell Physiol. Biochem. 2017, 44, 1445–1459. [Google Scholar] [CrossRef]
  12. Kaidanovich-Beilin, O.; Woodgett, J.R. GSK-3: Functional Insights from Cell Biology and Animal Models. Front. Mol. Neurosci. 2011, 4, 40. [Google Scholar] [CrossRef] [Green Version]
  13. Beurel, E.; Grieco, S.F.; Jope, R.S. Glycogen synthase kinase-3 (GSK3): Regulation, actions, and diseases. Pharmacol. Ther. 2015, 148, 114–131. [Google Scholar] [CrossRef] [Green Version]
  14. Beurel, E.; Michalek, S.M.; Jope, R.S. Innate and adaptive immune responses regulated by glycogen synthase kinase-3 (GSK3). Trends Immunol. 2009, 31, 24–31. [Google Scholar]
  15. Duda, P.; Wiśniewski, J.; Wójtowicz, T.; Wójcicka, O.; Jaskiewicz, M.R.; Drulis-Fajdasz, D.; Rakus, D.; McCubrey, J.A.; Gizak, A. Targeting GSK3 signaling as a potential therapy of neurodegenerative diseases and aging. Expert Opin. Ther. Targets 2018, 22, 833–848. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Huang, N.Q.; Yan, F.; Jin, H.; Zhou, S.Y.; Shi, J.S.; Jin, F. Diabetes mellitus and Alzheimer’s disease: GSK-3beta as a potential link. Behav. Brain Res. 2018, 339, 57–65. [Google Scholar] [CrossRef]
  17. Jope, R.S.; Cheng, Y.; Lowell, J.; Worthen, R.; Sitbon, Y.H.; Beurel, E. Stressed and Inflamed, Can GSK3 Be Blamed? Trends Biochem. Sci. 2016, 42, 180–192. [Google Scholar] [CrossRef] [Green Version]
  18. Arioka, M.; Takahashi-Yanaga, F. Glycogen synthase kinase-3 inhibitor as a multi-targeting anti-rheumatoid drug. Biochem. Pharmacol. 2019, 165, 207–213. [Google Scholar]
  19. Walz, A.; Ugolkov, A.; Chandra, S.; Kozikowski, A.; Carneiro, B.A.; O’Halloran, T.V.; Giles, F.J.; Billadeau, D.D.; Mazar, A.P. Molecular Pathways: Revisiting Glycogen Synthase Kinase-3beta as a Target for the Treatment of Cancer. Clin. Cancer Res. 2017, 23, 1891–1897. [Google Scholar]
  20. Souder, D.C.; Anderson, R.M. An expanding GSK3 network: Implications for aging research. GeroScience 2019, 41, 369–382. [Google Scholar]
  21. McCubrey, J.A.; Lertpiriyapong, K.; Steelman, L.S.; Abrams, S.L.; Cocco, L.; Ratti, S.; Martelli, A.; Candido, S.; Libra, M.; Montalto, G.; et al. Regulation of GSK-3 activity by curcumin, berberine and resveratrol: Potential effects on multiple diseases. Adv. Biol. Regul. 2017, 65, 77–88. [Google Scholar] [CrossRef]
  22. McCubrey, J.A.; Steelman, L.S.; Bertrand, F.E.; Davis, N.M.; Sokolosky, M.; Abrams, S.L.; Montalto, G.; D’Assoro, A.B.; Libra, M.; Nicoletti, F.; et al. GSK-3 as potential target for therapeutic intervention in cancer. Oncotarget 2014, 5, 2881–2911. [Google Scholar] [CrossRef] [Green Version]
  23. Yang, S.D.; Vandenheede, J.R.; Goris, J.; Merlevede, W. ATP x Mg-dependent protein phosphatase from rabbit skeletal muscle. I. Purification of the enzyme and its regulation by the interaction with an activating protein factor. J. Biol. Chem. 1980, 255, 11759–11767. [Google Scholar]
  24. Vandenheede, J.R.; Yang, S.D.; Goris, J.; Merlevede, W. ATP x Mg-dependent protein phosphatase from rabbit skeletal muscle. II. Purification of the activating factor and its characterization as a bifunctional protein also displaying synthase kinase activity. J. Biol. Chem. 1980, 255, 11768–11774. [Google Scholar]
  25. Embi, N.; Rylatt, D.B.; Cohen, P. Glycogen synthase kinase-3 from rabbit skeletal muscle. Separation from cyclic-AMP-dependent protein kinase and phosphorylase kinase. JBIC J. Biol. Inorg. Chem. 1980, 107, 519–527. [Google Scholar]
  26. Rylatt, D.B.; Aitken, A.; Bilham, T.; Condon, G.D.; Embi, N.; Cohen, P. Glycogen synthase from rabbit skeletal muscle. Amino acid sequence at the sites phosphorylated by glycogen synthase kinase-3, and extension of the N-terminal sequence containing the site phosphorylated by phosphorylase kinase. JBIC J. Biol. Inorg. Chem. 1980, 107, 529–537. [Google Scholar]
  27. Woodgett, J.R. Molecular cloning and expression of glycogen synthase kinase-3/factor A. EMBO J. 1990, 9, 2431–2438. [Google Scholar] [CrossRef]
  28. Mukai, F.; Ishiguro, K.; Sano, Y.; Fujita, S.C. Alternative splicing isoform of tau protein kinase I/glycogen synthase kinase 3beta. J. Neurochem. 2002, 81, 1073–1083. [Google Scholar] [CrossRef]
  29. Wagner, F.F.; Benajiba, L.; Campbell, A.J.; Weïwer, M.; Sacher, J.R.; Gale, J.P.; Ross, L.; Puissant, A.; Alexe, G.; Conway, A.S.; et al. Exploiting an Asp-Glu “switch” in glycogen synthase kinase 3 to design paralog-selective inhibitors for use in acute myeloid leukemia. Sci. Transl. Med. 2018, 10, eaam8460. [Google Scholar] [CrossRef] [Green Version]
  30. Doble, B.W.; Patel, S.; Wood, G.A.; Kockeritz, L.K.; Woodgett, J.R. Functional redundancy of GSK-3alpha and GSK-3beta in Wnt/beta-catenin signaling shown by using an allelic series of embryonic stem cell lines. Dev. Cell 2007, 12, 957–971. [Google Scholar] [CrossRef] [Green Version]
  31. Varjosalo, M.; Keskitalo, S.; Van Drogen, A.; Nurkkala, H.; Vichalkovski, A.; Aebersold, R.; Gstaiger, M. The Protein Interaction Landscape of the Human CMGC Kinase Group. Cell Rep. 2013, 3, 1306–1320. [Google Scholar] [CrossRef] [Green Version]
  32. Doble, B.W.; Woodgett, J.R. GSK-3: Tricks of the trade for a multi-tasking kinase. J. Cell Sci. 2003, 116, 1175–1186. [Google Scholar] [CrossRef] [Green Version]
  33. Domoto, T.; Pyko, I.V.; Furuta, T.; Miyashita, K.; Uehara, M.; Shimasaki, T.; Nakada, M.; Minamoto, T. Glycogen synthase kinase-3beta is a pivotal mediator of cancer invasion and resistance to therapy. Cancer Sci. 2016, 107, 1363–1372. [Google Scholar] [CrossRef]
  34. Buescher, J.L.; Phiel, C. A Noncatalytic Domain of Glycogen Synthase Kinase-3 (GSK-3) Is Essential for Activity*. J. Biol. Chem. 2010, 285, 7957–7963. [Google Scholar] [CrossRef] [Green Version]
  35. Zeke, A.; Misheva, M.; Remenyi, A.; Bogoyevitch, M. JNK Signaling: Regulation and Functions Based on Complex Protein-Protein Partnerships. Microbiol. Mol. Biol. Rev. 2016, 80, 793–835. [Google Scholar] [CrossRef] [Green Version]
  36. Fiol, C.J.; Mahrenholz, A.M.; Wang, Y.; Roeske, R.W.; Roach, P.J. Formation of protein kinase recognition sites by covalent modification of the substrate. Molecular mechanism for the synergistic action of casein kinase II and glycogen synthase kinase 3. J. Biol. Chem. 1987, 262, 14042–14048. [Google Scholar]
  37. Thomas, G.M.; Frame, S.; Goedert, M.; Nathke, I.; Polakis, P.; Cohen, P. A GSK3-binding peptide from FRAT1 selectively inhibits the GSK3-catalysed phosphorylation of axin and beta-catenin. FEBS Lett. 1999, 458, 247–251. [Google Scholar] [CrossRef] [Green Version]
  38. Fiol, C.J.; Wang, A.; Roeske, R.W.; Roach, P.J. Ordered multisite protein phosphorylation. Analysis of glycogen synthase kinase 3 action using model peptide substrates. J. Biol. Chem. 1990, 265, 6061–6065. [Google Scholar]
  39. Sutherland, C. What Are the bona fide GSK3 Substrates? Int. J. Alzheimers Dis. 2011, 2011, 505607. [Google Scholar]
  40. Tang, Q.Q.; Gronborg, M.; Huang, H.; Kim, J.W.; Otto, T.C.; Pandey, A.; Lane, M.D. Sequential phosphorylation of CCAAT enhancer-binding protein beta by MAPK and glycogen synthase kinase 3beta is required for adipogenesis. Proc. Natl. Acad. Sci. USA 2005, 102, 9766–9771. [Google Scholar] [CrossRef] [Green Version]
  41. Lau, K.-F.; Miller, C.C.J.; Anderton, B.H.; Shaw, P.-C. Molecular Cloning and Characterization of the Human Glycogen Synthase Kinase-3β Promoter. Genomics 1999, 60, 121–128. [Google Scholar] [CrossRef]
  42. Lee, K.-F.; Chan, J.Y.-C.; Lau, K.-F.; Lee, W.-C.; Miller, C.C.J.; Anderton, B.H.; Shaw, P.-C. Molecular cloning and expression analysis of human glycogen synthase kinase-3α promoter. Mol. Brain Res. 2000, 84, 150–157. [Google Scholar] [CrossRef]
  43. Park, S.-A.; Lee, J.W.; Herbst, R.S.; Koo, J.S. GSK-3α Is a Novel Target of CREB and CREB-GSK-3α Signaling Participates in Cell Viability in Lung Cancer. PLoS ONE 2016, 11, e0153075. [Google Scholar] [CrossRef] [Green Version]
  44. Huber, R.; Pietsch, D.; Panterodt, T.; Brand, K. Regulation of C/EBPbeta and resulting functions in cells of the monocytic lineage. Cell Signal 2012, 24, 1287–1296. [Google Scholar] [CrossRef] [Green Version]
  45. Kwok, J.B.; Hallupp, M.; Loy, C.T.; Chan, D.K.; Woo, J.; Mellick, G.D.; Buchanan, D.D.; Silburn, P.A.; Halliday, G.M.; Schofield, P.R. GSK3B polymorphisms alter transcription and splicing in Parkinson’s disease. Ann. Neurol. 2005, 58, 829–839. [Google Scholar] [CrossRef]
  46. Ali, A.; Hoeflich, K.P.; Woodgett, J.R. Glycogen Synthase Kinase-3: Properties, Functions, and Regulation. Chem. Rev. 2001, 101, 2527–2540. [Google Scholar] [CrossRef]
  47. Alon, L.T.; Pietrokovski, S.; Barkan, S.; Avrahami, L.; Kaidanovich-Beilin, O.; Woodgett, J.R.; Barnea, A.; Eldar-Finkelman, H. Selective loss of glycogen synthase kinase-3α in birds reveals distinct roles for GSK-3 isozymes in tau phosphorylation. FEBS Lett. 2011, 585, 1158–1162. [Google Scholar] [CrossRef] [Green Version]
  48. The National Center for Biotechnology Information nucleotide data base. Available online: https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/nuccore/ (accessed on 11 March 2020).
  49. Atlas of Genetics and Cytogenetics in Oncology and Haematology. Available online: https://atlasgeneticsoncology.org/ (accessed on 11 March 2020).
  50. Samaan, S.; Tranchevent, L.-C.; Dardenne, E.; Espinoza, M.P.; Zonta, E.; Germann, S.; Gratadou, L.; Dutertre, M.; Auboeuf, D. The Ddx5 and Ddx17 RNA helicases are cornerstones in the complex regulatory array of steroid hormone-signaling pathways. Nucleic Acids Res. 2013, 42, 2197–2207. [Google Scholar] [CrossRef]
  51. Schaffer, B.; Wiedau-Pazos, M.; Geschwind, D. Gene structure and alternative splicing of glycogen synthase kinase 3 beta (GSK-3β) in neural and non-neural tissues. Gene 2003, 302, 73–81. [Google Scholar] [CrossRef]
  52. Wang, L.; Zuo, B.; Xu, D.; Ren, Z.; Zhang, H.; Li, X.; Lei, M.; Xiong, Y. Alternative splicing of the porcine glycogen synthase kinase 3beta (GSK-3beta) gene with differential expression patterns and regulatory functions. PLoS ONE 2012, 7, e40250. [Google Scholar]
  53. Hou, Y.; Wang, Y.; Wang, Y.; Zhong, T.; Li, L.; Zhang, H.; Wang, L. Multiple alternative splicing and differential expression pattern of the glycogen synthase kinase-3beta (GSK3beta) gene in goat (Capra hircus). PLoS ONE 2014, 9, e109555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. The Human Protein Atlas. Available online: https://www.proteinatlas.org (accessed on 11 March 2020).
  55. Beurel, E.; A Mines, M.; Song, L.; Jope, R.S. Glycogen synthase kinase-3 levels and phosphorylation undergo large fluctuations in mouse brain during development. Bipolar Disord. 2012, 14, 822–830. [Google Scholar] [CrossRef] [Green Version]
  56. Beurel, E.; Yeh, W.-I.; Michalek, S.M.; Harrington, L.E.; Jope, R.S. Glycogen synthase kinase-3 is an early determinant in the differentiation of pathogenic Th17 cells. J. Immunol. 2010, 186, 1391–1398. [Google Scholar] [CrossRef] [Green Version]
  57. Hoffman, O.; Burns, N.; Vadasz, I.; Eltzschig, H.K.; Edwards, M.G.; Vohwinkel, C.U. Detrimental ELAVL-1/HuR-dependent GSK3beta mRNA stabilization impairs resolution in acute respiratory distress syndrome. PLoS ONE 2017, 12, e0172116. [Google Scholar]
  58. National Center for Biotechnology Information protein data base. Available online: https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/protein (accessed on 11 March 2020).
  59. Hur, E.-M.; Zhou, F.-Q. GSK3 signalling in neural development. Nat. Rev. Neurosci. 2010, 11, 539–551. [Google Scholar] [CrossRef] [Green Version]
  60. Soutar, M.P.M.; Kim, W.-Y.; Williamson, R.; Peggie, M.; Hastie, C.J.; Mclauchlan, H.; Snider, W.D.; Gordon-Weeks, P.; Sutherland, C. Evidence that glycogen synthase kinase-3 isoforms have distinct substrate preference in the brain. J. Neurochem. 2010, 115, 974–983. [Google Scholar] [CrossRef]
  61. Wood-Kaczmar, A.; Kraus, M.; Ishiguro, K.; Philpott, K.L.; Gordon-Weeks, P.R. An alternatively spliced form of glycogen synthase kinase-3beta is targeted to growing neurites and growth cones. Mol. Cell. Neurosci. 2009, 42, 184–194. [Google Scholar] [CrossRef]
  62. Castano, Z.; Gordon-Weeks, P.R.; Kypta, R.M. The neuron-specific isoform of glycogen synthase kinase-3beta is required for axon growth. J. Neurochem. 2010, 113, 117–130. [Google Scholar] [CrossRef]
  63. Azoulay-Alfaguter, I.; Yaffe, Y.; Licht-Murava, A.; Urbanska, M.; Jaworski, J.; Pietrokovski, S.; Hirschberg, K.; Eldar-Finkelman, H. Distinct molecular regulation of glycogen synthase kinase-3alpha isozyme controlled by its N-terminal region: Functional role in calcium/calpain signaling. J. Biol. Chem. 2011, 286, 13470–13480. [Google Scholar] [CrossRef] [Green Version]
  64. Bijur, G.N.; Jope, R.S. Glycogen synthase kinase-3 beta is highly activated in nuclei and mitochondria. NeuroReport 2003, 14, 2415–2419. [Google Scholar] [CrossRef]
  65. Bautista, S.J.; Boras, I.; Vissa, A.; Mecica, N.; Yip, C.M.; Kim, P.K.; Antonescu, C.N. mTOR complex 1 controls the nuclear localization and function of glycogen synthase kinase 3beta. J. Biol. Chem. 2018, 293, 14723–14739. [Google Scholar] [CrossRef] [Green Version]
  66. Meares, G.P.; Jope, R.S. Resolution of the nuclear localization mechanism of glycogen synthase kinase-3: Functional effects in apoptosis. J. Biol. Chem. 2007, 282, 16989–17001. [Google Scholar] [CrossRef] [Green Version]
  67. Bechard, M.; Dalton, S. Subcellular Localization of Glycogen Synthase Kinase 3β Controls Embryonic Stem Cell Self-Renewal. Mol. Cell. Biol. 2009, 29, 2092–2104. [Google Scholar] [CrossRef] [Green Version]
  68. Shin, S.H.; Lee, E.J.; Chun, J.; Hyun, S.; Kim, Y.I.; Kang, S.S. The nuclear localization of glycogen synthase kinase 3beta is required its putative PY-nuclear localization sequences. Mol Cells 2012, 34, 375–382. [Google Scholar] [CrossRef]
  69. Fujimuro, M.; Hayward, S.D. The latency-associated nuclear antigen of Kaposi’s sarcoma-associated herpesvirus manipulates the activity of glycogen synthase kinase-3beta. J. Virol. 2003, 77, 8019–8030. [Google Scholar] [CrossRef] [Green Version]
  70. Franca-Koh, J.; Yeo, M.; Fraser, E.; Young, N.; Dale, T.C. The Regulation of Glycogen Synthase Kinase-3 Nuclear Export by Frat/GBP. J. Biol. Chem. 2002, 277, 43844–43848. [Google Scholar] [CrossRef] [Green Version]
  71. Dajani, R.; Fraser, E.; Roe, S.M.; Young, N.; Good, V.; Dale, T.C.; Pearl, L. Crystal structure of glycogen synthase kinase 3 beta: Structural basis for phosphate-primed substrate specificity and autoinhibition. Cell 2001, 105, 721–732. [Google Scholar] [CrossRef]
  72. Cole, A.; Frame, S.; Cohen, P. Further evidence that the tyrosine phosphorylation of glycogen synthase kinase-3 (GSK3) in mammalian cells is an autophosphorylation event. Biochem. J. 2004, 377, 249–255. [Google Scholar] [CrossRef] [Green Version]
  73. Hughes, K.; Nikolakaki, E.; Plyte, S.; Totty, N.; Woodgett, J.R. Modulation of the glycogen synthase kinase-3 family by tyrosine phosphorylation. EMBO J. 1993, 12, 803–808. [Google Scholar] [CrossRef]
  74. Frame, S.; Cohen, P.; Biondi, R.M. A Common Phosphate Binding Site Explains the Unique Substrate Specificity of GSK3 and Its Inactivation by Phosphorylation. Mol. Cell 2001, 7, 1321–1327. [Google Scholar] [CrossRef]
  75. Dajani, R.; Fraser, E.; Roe, S.M.; Yeo, M.; Good, V.M.; Thompson, V.; Dale, T.C.; Pearl, L.H. Structural basis for recruitment of glycogen synthase kinase 3beta to the axin-APC scaffold complex. EMBO J. 2003, 22, 494–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Itoh, K.; Tang, T.L.; Neel, B.G.; Sokol, S.Y. Specific modulation of ectodermal cell fates in Xenopus embryos by glycogen synthase kinase. Development 1995, 121, 3979–3988. [Google Scholar]
  77. Wang, Q.M.; Fiol, C.J.; A DePaoli-Roach, A.; Roach, P.J. Glycogen synthase kinase-3 beta is a dual specificity kinase differentially regulated by tyrosine and serine/threonine phosphorylation. J. Biol. Chem. 1994, 269, 14566–14574. [Google Scholar]
  78. Lochhead, P.A.; Kinstrie, R.; Sibbet, G.; Rawjee, T.; Morrice, N.; Cleghon, V. A Chaperone-Dependent GSK3β Transitional Intermediate Mediates Activation-Loop Autophosphorylation. Mol. Cell 2006, 24, 627–633. [Google Scholar] [CrossRef] [PubMed]
  79. McCubrey, J.A.; Fitzgerald, T.L.; Yang, L.V.; Lertpiriyapong, K.; Steelman, L.S.; Abrams, S.L.; Montalto, G.; Cervello, M.; Neri, L.M.; Cocco, L.; et al. Roles of GSK-3 and microRNAs on epithelial mesenchymal transition and cancer stem cells. Oncotarget 2016, 8, 14221–14250. [Google Scholar] [CrossRef]
  80. Bhat, R.V.; Shanley, J.; Correll, M.P.; Fieles, W.E.; Keith, R.A.; Scott, C.W.; Lee, C.-M. Regulation and localization of tyrosine216 phosphorylation of glycogen synthase kinase-3beta in cellular and animal models of neuronal degeneration. Proc. Natl. Acad. Sci. USA 2000, 97, 11074–11079. [Google Scholar] [CrossRef] [Green Version]
  81. Frame, S.; Cohen, P. GSK3 takes centre stage more than 20 years after its discovery. Biochem. J. 2001, 359, 1–16. [Google Scholar] [CrossRef]
  82. Ter Haar, E.; Coll, J.T.; A Austen, D.; Hsiao, H.M.; Swenson, L.; Jain, J. Structure of GSK3beta reveals a primed phosphorylation mechanism. Nat. Genet. 2001, 8, 593–596. [Google Scholar]
  83. Tejeda-Munoz, N.; Gonzalez-Aguilar, H.; Santoyo-Ramos, P.; Castaneda-Patlan, M.C.; Robles-Flores, M. Glycogen Synthase Kinase 3beta Is Positively Regulated by Protein Kinase Czeta-Mediated Phosphorylation Induced by Wnt Agonists. Mol. Cell. Biol. 2015, 36, 731–741. [Google Scholar] [CrossRef] [Green Version]
  84. Ding, Q.; Xia, W.; Liu, J.C.; Yang, J.Y.; Lee, D.F.; Xia, J.; Bartholomeusz, G.; Li, Y.; Pan, Y.; Li, Z.; et al. Erk associates with and primes GSK-3beta for its inactivation resulting in upregulation of beta-catenin. Mol. Cell 2005, 19, 159–170. [Google Scholar] [CrossRef]
  85. Song, W.J.; Song, E.A.; Jung, M.S.; Choi, S.H.; Baik, H.H.; Jin, B.K.; Kim, J.H.; Chung, S.H. Phosphorylation and inactivation of glycogen synthase kinase 3beta (GSK3beta) by dual-specificity tyrosine phosphorylation-regulated kinase 1A (Dyrk1A). J. Biol. Chem. 2015, 290, 2321–2333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Fang, X.; Yu, S.X.; Lu, Y.; Bast, R.C.; Woodgett, J.R.; Mills, G.B. Phosphorylation and inactivation of glycogen synthase kinase 3 by protein kinase A. Proc. Natl. Acad. Sci. USA 2000, 97, 11960–11965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Oguma, K.; Oshima, H.; Aoki, M.; Uchio, R.; Naka, K.; Nakamura, S.; Hirao, A.; Saya, H.; Taketo, M.M.; Oshima, M. Activated macrophages promote Wnt signalling through tumour necrosis factor-α in gastric tumour cells. EMBO J. 2008, 27, 1671–1681. [Google Scholar] [CrossRef] [Green Version]
  88. Wang, H.; Brown, J.; Garcia, C.A.; Tang, Y.; Benakanakere, M.R.; Greenway, T.; Alard, P.; Kinane, D.F.; Martin, M. The role of glycogen synthase kinase 3 in regulating IFN-beta-mediated IL-10 production. J. Immunol. 2011, 186, 675–684. [Google Scholar] [CrossRef] [PubMed]
  89. Cross, D.A.E.; Alessi, D.R.; Cohen, P.; Andjelkovich, M.; Hemmings, B.A. Inhibition of glycogen synthase kinase-3 by insulin mediated by protein kinase B. Nature 1995, 378, 785–789. [Google Scholar] [CrossRef] [PubMed]
  90. Martin, M.; Rehani, K.; Jope, R.S.; Michalek, S.M. Toll-like receptor-mediated cytokine production is differentially regulated by glycogen synthase kinase 3. Nat. Immunol. 2005, 6, 777–784. [Google Scholar] [CrossRef] [PubMed]
  91. Goode, N.; Hughes, K.; Woodgett, J.R.; Parker, P.J. Differential regulation of glycogen synthase kinase-3 beta by protein kinase C isotypes. J. Biol. Chem. 1992, 267, 16878–16882. [Google Scholar]
  92. Johnson, A. TNF-induced activation of pulmonary microvessel endothelial cells: A role for GSK3beta. Am. J. Physiol. Cell. Mol. Physiol. 2009, 296, L700–L709. [Google Scholar] [CrossRef]
  93. Cook, D.; Fry, M.; Hughes, K.; Sumathipala, R.; Woodgett, J.R.; Dale, T.C. Wingless inactivates glycogen synthase kinase-3 via an intracellular signalling pathway which involves a protein kinase C. EMBO J. 1996, 15, 4526–4536. [Google Scholar] [CrossRef]
  94. Fichtner-Feigl, S.; Kesselring, R.; Martin, M.; Obermeier, F.; Ruemmele, P.; Kitani, A.; Brunner, S.M.; Haimerl, M.; Geissler, E.K.; Strober, W.; et al. IL-13 orchestrates resolution of chronic intestinal inflammation via phosphorylation of glycogen synthase kinase-3beta. J. Immunol. 2014, 192, 3969–3980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Failor, K.L.; Desyatnikov, Y.; Finger, L.A.; Firestone, G.L. Glucocorticoid-Induced Degradation of Glycogen Synthase Kinase-3 Protein Is Triggered by Serum- and Glucocorticoid-Induced Protein Kinase and Akt Signaling and Controls ?-Catenin Dynamics and Tight Junction Formation in Mammary Epithelial Tumor Cells. Mol. Endocrinol. 2007, 21, 2403–2415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ding, V.W.; Chen, R.-H.; McCormick, F. Differential Regulation of Glycogen Synthase Kinase 3β by Insulin and Wnt Signaling. J. Biol. Chem. 2000, 275, 32475–32481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wu, D.; Pan, W. GSK3: A multifaceted kinase in Wnt signaling. Trends Biochem. Sci. 2010, 35, 161–168. [Google Scholar] [CrossRef] [Green Version]
  98. Thornton, T.M.; Pedraza-Alva, G.; Deng, B.; Wood, C.D.; Aronshtam, A.; Clements, J.L.; Sabio, G.; Davis, R.J.; Matthews, D.E.; Doble, B.W.; et al. Phosphorylation by p38 MAPK as an Alternative Pathway for GSK3 Inactivation. Science 2008, 320, 667–670. [Google Scholar] [CrossRef] [Green Version]
  99. Desbois-Mouthon, C.; Cadoret, A.; Eggelpoël, M.-J.B.-V.; Bertrand, F.; Cherqui, G.; Perret, C.; Capeau, J. Insulin and IGF-1 stimulate the β-catenin pathway through two signalling cascades involving GSK-3β inhibition and Ras activation. Oncogene 2001, 20, 252–259. [Google Scholar] [CrossRef] [Green Version]
  100. Chao, J.; Yang, L.; Yao, H.; Buch, S. Platelet-derived growth factor-BB restores HIV Tat -mediated impairment of neurogenesis: Role of GSK-3β/β-catenin. J. Neuroimmune Pharmacol. 2013, 9, 259–268. [Google Scholar] [CrossRef]
  101. Papkoff, J.; Aikawa, M. WNT-1 and HGF Regulate GSK3β Activity and β-Catenin Signaling in Mammary Epithelial Cells. Biochem. Biophys. Res. Commun. 1998, 247, 851–858. [Google Scholar] [CrossRef]
  102. Holnthoner, W.; Pillinger, M.; Gröger, M.; Wolff, K.; Ashton, A.W.; Albanese, C.; Neumeister, P.; Pestell, R.G.; Petzelbauer, P. Fibroblast Growth Factor-2 Induces Lef/Tcf-dependent Transcription in Human Endothelial Cells. J. Biol. Chem. 2002, 277, 45847–45853. [Google Scholar] [CrossRef] [Green Version]
  103. Cheon, S.S.; Nadesan, P.; Poon, R.; Alman, B. Growth factors regulate beta-catenin-mediated TCF-dependent transcriptional activation in fibroblasts during the proliferative phase of wound healing. Exp. Cell Res. 2004, 293, 267–274. [Google Scholar] [CrossRef]
  104. Wei, J.-J.; Song, C.-W.; Sun, L.-C.; Yuan, Y.; Li, N.; Yan, B.; Liao, S.-J.; Zhu, J.-H.; Wang, Q.; Zhang, G.-M.; et al. SCF and TLR4 ligand cooperate to augment the tumor-promoting potential of mast cells. Cancer Immunol. Immunother. 2011, 61, 303–312. [Google Scholar] [CrossRef] [PubMed]
  105. Hernández, F.; Langa, E.; Cuadros, R.; Ávila, J.; Villanueva, N. Regulation of GSK3 isoforms by phosphatases PP1 and PP2A. Mol. Cell. Biochem. 2010, 344, 211–215. [Google Scholar] [CrossRef] [PubMed]
  106. Sarikhani, M.; Mishra, S.; Maity, S.; Kotyada, C.; Wolfgeher, D.; Gupta, M.P.; Singh, M.; Sundaresan, N.R. SIRT2 deacetylase regulates the activity of GSK3 isoforms independent of inhibitory phosphorylation. eLife 2018, 7, e32952. [Google Scholar] [CrossRef] [PubMed]
  107. Monteserin-Garcia, J.; Al-Massadi, O.; Seoane, L.M.; Álvarez, C.V.; Shan, B.; Stalla, J.; Pereda, M.P.; Casanueva, F.F.; Stalla, G.K.; Theodoropoulou, M. Sirt1 inhibits the transcription factor CREB to regulate pituitary growth hormone synthesis. FASEB J. 2013, 27, 1561–1571. [Google Scholar] [CrossRef]
  108. Song, C.L.; Tang, H.; Ran, L.K.; Ko, B.C.; Zhang, Z.Z.; Chen, X.; Ren, J.H.; Tao, N.N.; Li, W.Y.; Huang, A.L.; et al. Sirtuin 3 inhibits hepatocellular carcinoma growth through the glycogen synthase kinase-3beta/BCL2-associated X protein-dependent apoptotic pathway. Oncogene 2016, 35, 631–641. [Google Scholar] [CrossRef]
  109. Feijs, K.L.; Kleine, H.; Braczynski, A.; Forst, A.H.; Herzog, N.; Verheugd, P.; Linzen, U.; Kremmer, E.; Lüscher, B. ARTD10 substrate identification on protein microarrays: Regulation of GSK3β by mono-ADP-ribosylation. Cell Commun. Signal. 2013, 11, 5. [Google Scholar] [CrossRef] [Green Version]
  110. Rosenthal, F.; Feijs, K.L.; Frugier, E.; Bonalli, M.; Forst, A.H.; Imhof, R.; Winkler, H.; Fischer, D.; Caflisch, A.; O Hassa, P.; et al. Macrodomain-containing proteins are new mono-ADP-ribosylhydrolases. Nat. Struct. Mol. Biol. 2013, 20, 502–507. [Google Scholar] [CrossRef]
  111. Eun Jeoung, L.; Sung Hee, H.; Jaesun, C.; Sung Hwa, S.; Kwang Hum, Y.; Min Kyoung, K.; Tae Yoon, P.; Sang Sun, K. Regulation of glycogen synthase kinase 3beta functions by modification of the small ubiquitin-like modifier. Open Biochem. J. 2008, 2, 67–76. [Google Scholar]
  112. Stadler, S.C.; Vincent, C.T.; Fedorov, V.D.; Patsialou, A.; Cherrington, B.D.; Wakshlag, J.J.; Mohanan, S.; Zee, B.M.; Zhang, X.; Garcia, B.A.; et al. Dysregulation of PAD4-mediated citrullination of nuclear GSK3beta activates TGF-beta signaling and induces epithelial-to-mesenchymal transition in breast cancer cells. Proc. Natl. Acad. Sci. USA 2013, 110, 11851–11856. [Google Scholar] [CrossRef] [Green Version]
  113. Udeshi, N.D.; Svinkina, T.; Mertins, P.; Kuhn, E.; Mani, D.R.; Qiao, J.W.; Carr, S.A. Refined preparation and use of anti-diglycine remnant (K-epsilon-GG) antibody enables routine quantification of 10,000s of ubiquitination sites in single proteomics experiments. Mol. Cell. Proteom. 2013, 12, 825–831. [Google Scholar] [CrossRef] [Green Version]
  114. Larsen, S.C.; Sylvestersen, K.B.; Mund, A.; Lyon, D.; Mullari, M.; Madsen, M.V.; Daniel, J.; Jensen, L.J.; Nielsen, M.L. Proteome-wide analysis of arginine monomethylation reveals widespread occurrence in human cells. Sci. Signal. 2016, 9, rs9. [Google Scholar] [CrossRef] [PubMed]
  115. Olsen, J.B.; Cao, X.-J.; Han, B.; Chen, L.H.; Horvath, A.; Richardson, T.I.; Campbell, R.M.; Garcia, B.A.; Nguyen, H. Quantitative Profiling of the Activity of Protein Lysine Methyltransferase SMYD2 Using SILAC-Based Proteomics. Mol. Cell. Proteom. 2016, 15, 892–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Goñi-Oliver, P.; Lucas, J.J.; Avila, J.; Hernández, F. N-terminal Cleavage of GSK-3 by Calpain. J. Biol. Chem. 2007, 282, 22406–22413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Ma, S.; Liu, S.; Huang, Q.; Xie, B.; Lai, B.; Wang, C.; Song, B.; Li, M. Site-specific phosphorylation protects glycogen synthase kinase-3beta from calpain-mediated truncation of its N and C termini. J. Biol. Chem. 2012, 287, 22521–22532. [Google Scholar] [CrossRef] [Green Version]
  118. Kandasamy, A.D.; Schulz, R. Glycogen synthase kinase-3beta is activated by matrix metalloproteinase-2 mediated proteolysis in cardiomyoblasts. Cardiovasc. Res. 2009, 83, 698–706. [Google Scholar] [CrossRef] [Green Version]
  119. Jin, N.; Wu, Y.; Xu, W.; Gong, C.X.; Iqbal, K.; Liu, F. C-terminal truncation of GSK-3beta enhances its dephosphorylation by PP2A. FEBS Lett. 2017, 591, 1053–1063. [Google Scholar] [CrossRef] [Green Version]
  120. Goni-Oliver, P.; Avila, J.; Hernandez, F. Calpain regulates N-terminal interaction of GSK-3beta with 14–3-3zeta, p53 and PKB but not with axin. Neurochem. Int. 2011, 59, 97–100. [Google Scholar] [CrossRef]
  121. Jope, R.S.; Yuskaitis, C.J.; Beurel, E. Glycogen Synthase Kinase-3 (GSK3): Inflammation, Diseases, and Therapeutics. Neurochem. Res. 2006, 32, 577–595. [Google Scholar] [CrossRef] [Green Version]
  122. Ding, Q.; Liu, G.; Zeng, Y.; Zhu, J.; Liu, Z.; Jiang, J.; Huang, J. Glycogen synthase kinase3beta inhibitor reduces LPSinduced acute lung injury in mice. Mol. Med. Rep. 2017, 16, 6715–6721. [Google Scholar] [CrossRef]
  123. Jing, H.; Yen, J.H.; Ganea, D. A novel signaling pathway mediates the inhibition of CCL3/4 expression by prostaglandin E2. J. Biol. Chem. 2004, 279, 55176–55186. [Google Scholar] [CrossRef] [Green Version]
  124. Park, C.; Lee, S.; Cho, I.-H.; Lee, H.K.; Kim, N.; Choi, S.-Y.; Oh, S.B.; Park, K.; Kim, J.S.; Lee, S.J. TLR3-mediated signal induces proinflammatory cytokine and chemokine gene expression in astrocytes: Differential signaling mechanisms of TLR3-induced IP-10 and IL-8 gene expression. Glia 2005, 53, 248–256. [Google Scholar] [CrossRef] [PubMed]
  125. Chang, Y.T.; Chen, C.L.; Lin, C.F.; Lu, S.L.; Cheng, M.H.; Kuo, C.F.; Lin, Y.S. Regulatory role of GSK-3 beta on NF- kappa B, nitric oxide, and TNF- alpha in group A streptococcal infection. Mediators Inflamm. 2013, 2013, 720689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Noma, T.; Takahashi-Yanaga, F.; Arioka, M.; Mori, Y.; Sasaguri, T. Inhibition of GSK-3 reduces prostaglandin E2 production by decreasing the expression levels of COX-2 and mPGES-1 in monocyte/macrophage lineage cells. Biochem. Pharmacol. 2016, 116, 120–129. [Google Scholar] [CrossRef] [PubMed]
  127. Hwang, E.S.; Hong, J.-H.; Glimcher, L.H. IL-2 production in developing Th1 cells is regulated by heterodimerization of RelA and T-bet and requires T-bet serine residue 508. J. Exp. Med. 2005, 202, 1289–1300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Lutay, N.; Håkansson, G.; Alaridah, N.; Hallgren, O.; Westergren-Thorsson, G.; Godaly, G. Mycobacteria Bypass Mucosal NF-kB Signalling to Induce an Epithelial Anti-Inflammatory IL-22 and IL-10 Response. PLoS ONE 2014, 9, 86466. [Google Scholar] [CrossRef]
  129. Morris, M.; Gilliam, E.A.; Button, J.; Li, L. Dynamic Modulation of Innate Immune Response by Varying Dosages of Lipopolysaccharide (LPS) in Human Monocytic Cells. J. Biol. Chem. 2014, 289, 21584–21590. [Google Scholar] [CrossRef] [Green Version]
  130. Rehani, K.; Wang, H.; Garcia, C.A.; Kinane, D.F.; Martin, M. Toll-like receptor-mediated production of IL-1Ra is negatively regulated by GSK3 via the MAPK ERK1/2. J. Immunol. 2009, 182, 547–553. [Google Scholar] [CrossRef]
  131. Hong, H.; Chen, F.; Qiao, Y.; Yan, Y.; Zhang, R.; Zhu, Z.; Li, H.; Fan, Y.; Xu, G. GSK-3beta activation index is a potential indicator for recurrent inflammation of chronic rhinosinusitis without nasal polyps. J. Cell. Mol. Med. 2017, 21, 3633–3640. [Google Scholar] [CrossRef] [Green Version]
  132. Gao, Y.; Mi, J.; Chen, F.; Liao, Z.; Feng, X.; Lv, M.; He, H.; Cao, Y.; Yan, Y.; Zhu, Z.; et al. Detection of GSK-3beta activation index in pediatric chronic tonsillitis is an indicator for chronic recurrent inflammation. Am. J. Otolaryngol. 2018, 39, 277–281. [Google Scholar] [CrossRef]
  133. Macaulay, K.; Doble, B.W.; Patel, S.; Hansotia, T.; Sinclair, E.M.; Drucker, D.J.; Nagy, A.; Woodgett, J.R. Glycogen Synthase Kinase 3α-Specific Regulation of Murine Hepatic Glycogen Metabolism. Cell Metab. 2007, 6, 329–337. [Google Scholar] [CrossRef] [Green Version]
  134. Barrell, W.B.; Szabo-Rogers, H.L.; Liu, K.J. Novel reporter alleles of GSK-3alpha and GSK-3beta. PLoS ONE 2012, 7, e50422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Nørregaard, R.; Tao, S.; Nilsson, L.; Woodgett, J.R.; Kakade, V.; Yu, A.S.L.; Howard, C.; Rao, R. Glycogen synthase kinase 3α regulates urine concentrating mechanism in mice. Am. J. Physiol. Physiol. 2015, 308, F650–F660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Maurin, H.; Lechat, B.; Dewachter, I.; Ris, L.; Louis, J.; Borghgraef, P.; Devijver, H.; Jaworski, T.; Van Leuven, F. Neurological characterization of mice deficient in GSK3α highlight pleiotropic physiological functions in cognition and pathological activity as Tau kinase. Mol. Brain 2013, 6, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hoeflich, K.P.; Luo, J.; Rubie, E.A.; Tsao, M.S.; Jin, O.; Woodgett, J.R. Requirement for glycogen synthase kinase-3beta in cell survival and NF-kappaB activation. Nature 2000, 406, 86–90. [Google Scholar] [CrossRef] [PubMed]
  138. Kerkela, R.; Kockeritz, L.; Macaulay, K.; Zhou, J.; Doble, B.W.; Beahm, C.; Greytak, S.; Woulfe, K.; Trivedi, C.M.; Woodgett, J.R.; et al. Deletion of GSK-3beta in mice leads to hypertrophic cardiomyopathy secondary to cardiomyoblast hyperproliferation. J. Clin. Invest. 2008, 118, 3609–3618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Zhou, H.; Wang, H.; Ni, M.; Yue, S.; Xia, Y.; Busuttil, R.W.; Kupiec-Weglinski, J.W.; Lü, L.; Wang, X.; Zhai, Y. Glycogen synthase kinase 3β promotes liver innate immune activation by restraining AMP-activated protein kinase activation. J. Hepatol. 2018, 69, 99–109. [Google Scholar] [CrossRef]
  140. Howard, C.; Tao, S.; Yang, H.C.; Fogo, A.B.; Woodgett, J.R.; Harris, R.C.; Rao, R. Specific deletion of glycogen synthase kinase-3beta in the renal proximal tubule protects against acute nephrotoxic injury in mice. Kidney Int. 2012, 82, 1000–1009. [Google Scholar] [CrossRef] [Green Version]
  141. Li, C.; Ge, Y.; Dworkin, L.; Peng, A.; Gong, R. The beta isoform of GSK3 mediates podocyte autonomous injury in proteinuric glomerulopathy. J. Pathol. 2016, 239, 23–35. [Google Scholar] [CrossRef] [Green Version]
  142. Li, C.; Ge, Y.; Peng, A.; Gong, R. The redox sensitive glycogen synthase kinase 3beta suppresses the self-protective antioxidant response in podocytes upon oxidative glomerular injury. Oncotarget 2015, 6, 39493–39506. [Google Scholar] [CrossRef] [Green Version]
  143. Zhou, S.; Wang, P.; Qiao, Y.; Ge, Y.; Wang, Y.; Quan, S.; Yao, R.; Zhuang, S.; Wang, L.J.; Du, Y.; et al. Genetic and Pharmacologic Targeting of Glycogen Synthase Kinase 3beta Reinforces the Nrf2 Antioxidant Defense against Podocytopathy. J. Am. Soc. Nephrol. 2016, 27, 2289–2308. [Google Scholar] [CrossRef] [Green Version]
  144. Xing, B.; Brink, L.E.; Maers, K.; Sullivan, M.L.; Bodnar, R.J.; Stolz, N.B.; Cambi, F. Conditional depletion of GSK3b protects oligodendrocytes from apoptosis and lessens demyelination in the acute cuprizone model. Glia 2018, 66, 1999–2012. [Google Scholar] [CrossRef] [PubMed]
  145. Hu, X.; Paik, P.K.; Chen, J.; Yarilina, A.; Kockeritz, L.; Lu, T.T.; Woodgett, J.R.; Ivashkiv, L.B. IFN-γ Suppresses IL-10 Production and Synergizes with TLR2 by Regulating GSK3 and CREB/AP-1 Proteins. Immunity 2006, 24, 563–574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Cuzzocrea, S.; Mazzon, E.; Di Paola, R.; Muia, C.; Crisafulli, C.; Dugo, L.; Collin, M.; Britti, D.; Caputi, A.P.; Thiemermann, C. Glycogen synthase kinase-3beta inhibition attenuates the degree of arthritis caused by type II collagen in the mouse. Clin. Immunol. 2006, 120, 57–67. [Google Scholar] [CrossRef] [PubMed]
  147. Kwon, Y.J.; Yoon, C.H.; Lee, S.W.; Park, Y.B.; Lee, S.K.; Park, M.C. Inhibition of glycogen synthase kinase-3beta suppresses inflammatory responses in rheumatoid arthritis fibroblast-like synoviocytes and collagen-induced arthritis. Joint Bone Spine 2014, 81, 240–246. [Google Scholar] [CrossRef]
  148. Zheng, Y.; Yang, Y.; Li, Y.; Xu, L.; Wang, Y.; Guo, Z.; Song, H.; Yang, M.; Luo, B.; Zheng, A.; et al. Ephedrine hydrochloride inhibits PGN-induced inflammatory responses by promoting IL-10 production and decreasing proinflammatory cytokine secretion via the PI3K/Akt/GSK3beta pathway. Cell. Mol. Immunol. 2013, 10, 330–337. [Google Scholar] [CrossRef] [Green Version]
  149. Koh, S.-H.; Yoo, A.R.; Chang, D.-I.; Hwang, S.J.; Kim, S.H. Inhibition of GSK-3 reduces infarct volume and improves neurobehavioral functions. Biochem. Biophys. Res. Commun. 2008, 371, 894–899. [Google Scholar] [CrossRef]
  150. Jiang, R.; Chen, D.; Hou, J.; Tan, Z.; Wang, Y.; Huang, X.; Wang, X.; Sun, B. Survival and inflammation promotion effect of PTPRO in fulminant hepatitis is associated with NF-kappaB activation. J. Immunol. 2014, 193, 5161–5170. [Google Scholar] [CrossRef]
  151. Wang, D.Z.; Jones, A.W.; Wang, W.; Wang, M.; Korthuis, R.J. Soluble guanylate cyclase activation during ischemic injury in mice protects against postischemic inflammation at the mitochondrial level. Am. J. Physiol. Liver Physiol. 2016, 310, G747–G756. [Google Scholar] [CrossRef] [Green Version]
  152. D’Angelo, B.; Ek, C.J.; Sun, Y.; Zhu, C.; Sandberg, M.; Mallard, C. GSK3beta inhibition protects the immature brain from hypoxic-ischaemic insult via reduced STAT3 signalling. Neuropharmacology 2016, 101, 13–23. [Google Scholar] [CrossRef] [Green Version]
  153. Zhang, C.; Lu, X.; Tan, Y.; Li, B.; Miao, X.; Jin, L.; Shi, X.; Zhang, X.; Miao, L.; Li, X.-K.; et al. Diabetes-Induced Hepatic Pathogenic Damage, Inflammation, Oxidative Stress, and Insulin Resistance Was Exacerbated in Zinc Deficient Mouse Model. PLoS ONE 2012, 7, e49257. [Google Scholar] [CrossRef] [Green Version]
  154. Wang, Y.; Feng, W.; Xue, W.; Tan, Y.; Hein, D.W.; Li, X.K.; Cai, L. Inactivation of GSK-3beta by metallothionein prevents diabetes-related changes in cardiac energy metabolism, inflammation, nitrosative damage, and remodeling. Diabetes 2009, 58, 1391–1402. [Google Scholar] [CrossRef] [Green Version]
  155. Gomaa, A.A.; Makboul, R.M.; Al-Mokhtar, M.A.; Nicola, M.A. Polyphenol-rich Boswellia serrata gum prevents cognitive impairment and insulin resistance of diabetic rats through inhibition of GSK3beta activity, oxidative stress and pro-inflammatory cytokines. Biomed. Pharmacother. 2019, 109, 281–292. [Google Scholar] [CrossRef]
  156. Datusalia, A.K.; Sharma, S.S. Amelioration of diabetes-induced cognitive deficits by GSK-3beta inhibition is attributed to modulation of neurotransmitters and neuroinflammation. Mol. Neurobiol. 2014, 50, 390–405. [Google Scholar] [CrossRef]
  157. Llorens-Martin, M.; Jurado-Arjona, J.; Fuster-Matanzo, A.; Hernandez, F.; Rabano, A.; Avila, J. Peripherally triggered and GSK-3beta-driven brain inflammation differentially skew adult hippocampal neurogenesis, behavioral pattern separation and microglial activation in response to ibuprofen. Transl. Psychiatry 2014, 4, e463. [Google Scholar] [CrossRef]
  158. Dionisio, P.A.; Amaral, J.D.; Ribeiro, M.F.; Lo, A.C.; D’Hooge, R.; Rodrigues, C.M. Amyloid-beta pathology is attenuated by tauroursodeoxycholic acid treatment in APP/PS1 mice after disease onset. Neurobiol. Aging 2015, 36, 228–240. [Google Scholar] [CrossRef] [Green Version]
  159. Ko, R.; Lee, S.Y. Glycogen synthase kinase 3beta in Toll-like receptor signaling. BMB Rep. 2016, 49, 305–310. [Google Scholar] [CrossRef] [Green Version]
  160. Liu, Y.; Li, J.-Y.; Chen, S.-T.; Huang, H.-R.; Cai, H. The rLrp of Mycobacterium tuberculosis inhibits proinflammatory cytokine production and downregulates APC function in mouse macrophages via a TLR2-mediated PI3K/Akt pathway activation-dependent mechanism. Cell. Mol. Immunol. 2015, 13, 729–746. [Google Scholar] [CrossRef] [Green Version]
  161. Wang, H.; Brown, J.; Gu, Z.; Garcia, C.A.; Liang, R.; Alard, P.; Beurel, E.; Jope, R.S.; Greenway, T.; Martin, M. Convergence of the mammalian target of rapamycin complex 1- and glycogen synthase kinase 3-beta-signaling pathways regulates the innate inflammatory response. J. Immunol. 2011, 186, 5217–5226. [Google Scholar] [CrossRef]
  162. Tay, T.F.; Maheran, M.; Too, S.L.; Hasidah, M.S.; Ismail, G.; Embi, N. Glycogen synthase kinase-3beta inhibition improved survivability of mice infected with Burkholderia pseudomallei. Trop. Biomed. 2012, 29, 551–567. [Google Scholar]
  163. Zhang, P.; Katz, J.; Michalek, S.M. Glycogen synthase kinase-3β (GSK3β) inhibition suppresses the inflammatory response to Francisella infection and protects against tularemia in mice. Mol. Immunol. 2008, 46, 677–687. [Google Scholar] [CrossRef] [Green Version]
  164. Nandan, D.; Camargo de Oliveira, C.; Moeenrezakhanlou, A.; Lopez, M.; Silverman, J.M.; Subek, J.; Reiner, N.E. Myeloid cell IL-10 production in response to leishmania involves inactivation of glycogen synthase kinase-3beta downstream of phosphatidylinositol-3 kinase. J. Immunol. 2012, 188, 367–378. [Google Scholar] [CrossRef] [Green Version]
  165. Paul, J.; Naskar, K.; Chowdhury, S.; Chakraborti, T.; De, T. TLR mediated GSK3beta activation suppresses CREB mediated IL-10 production to induce a protective immune response against murine visceral leishmaniasis. Biochimie 2014, 107 Pt B, 235–246. [Google Scholar] [CrossRef]
  166. Paul, J.; Karmakar, S.; De, T. TLR-mediated distinct IFN-gamma/IL-10 pattern induces protective immunity against murine visceral leishmaniasis. Eur. J. Immunol. 2012, 42, 2087–2099. [Google Scholar] [CrossRef]
  167. Cheng, Y.-L.; Wang, C.-Y.; Huang, W.-C.; Tsai, C.-C.; Chen, C.-L.; Shen, C.-F.; Chi, C.-Y.; Lin, C.-F. Staphylococcus aureus Induces Microglial Inflammation via a Glycogen Synthase Kinase 3β-Regulated Pathway. Infect. Immun. 2009, 77, 4002–4008. [Google Scholar] [CrossRef] [Green Version]
  168. König, R.; Stertz, S.; Zhou, Y.; Inoue, A.; Hoffmann, H.-H.; Bhattacharyya, S.; Alamares, J.G.; Tscherne, N.M.; Ortigoza, M.; Liang, Y.; et al. Human host factors required for influenza virus replication. Nature 2010, 463, 813–817. [Google Scholar]
  169. Guendel, I.; Iordanskiy, S.; Van Duyne, R.; Kehn-Hall, K.; Saifuddin, M.; Das, R.; Jaworski, E.; Sampey, G.C.; Senina, S.; Shultz, L. Novel neuroprotective GSK-3beta inhibitor restricts Tat-mediated HIV-1 replication. J. Virol. 2014, 88, 1189–1208. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, T.; Zhang, J.; Xiao, A.; Liu, W.; Shang, Y.; An, J. Melittin ameliorates CVB3-induced myocarditis via activation of the HDAC2-mediated GSK-3beta/Nrf2/ARE signaling pathway. Biochem. Biophys. Res. Commun. 2016, 480, 126–131. [Google Scholar] [CrossRef]
  171. Yuan, J.; Zhang, J.; Wong, B.W.; Si, X.; Wong, J.; Yang, D.; Luo, H. Inhibition of glycogen synthase kinase 3beta suppresses coxsackievirus-induced cytopathic effect and apoptosis via stabilization of beta-catenin. Cell Death Differ. 2005, 12, 1097–1106. [Google Scholar] [CrossRef] [Green Version]
  172. Kehn-Hall, K.; Guendel, I.; Carpio, L.; Skaltsounis, L.; Meijer, L.; Al-Harthi, L.; Steiner, J.P.; Nath, A.; Kutsch, O.; Kashanchi, F. Inhibition of Tat-mediated HIV-1 replication and neurotoxicity by novel GSK3-beta inhibitors. Virology 2011, 415, 56–68. [Google Scholar] [CrossRef] [Green Version]
  173. Rahaus, M.; Desloges, N.; Wolff, M.H. Varicella-zoster virus requires a functional PI3K/Akt/GSK-3alpha/beta signaling cascade for efficient replication. Cell Signal 2007, 19, 312–320. [Google Scholar] [CrossRef]
  174. Sarhan, M.A.; Abdel-Hakeem, M.S.; Mason, A.L.; Tyrrell, D.L.; Houghton, M. Glycogen synthase kinase 3beta inhibitors prevent hepatitis C virus release/assembly through perturbation of lipid metabolism. Sci. Rep. 2017, 7, 2495. [Google Scholar] [CrossRef]
  175. Saleh, M.; Ruschenbaum, S.; Welsch, C.; Zeuzem, S.; Moradpour, D.; Gouttenoire, J.; Lange, C.M. Glycogen Synthase Kinase 3beta Enhances Hepatitis C Virus Replication by Supporting miR-122. Front. Microbiol. 2018, 9, 2949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Hofmann, S.; Dehn, S.; Businger, R.; Bolduan, S.; Schneider, M.; Debyser, Z.; Brack-Werner, R.; Schindler, M. Dual role of the chromatin-binding factor PHF13 in the pre- and post-integration phases of HIV-1 replication. Open Biol. 2017, 7, 170115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Reichel, A.; Stilp, A.-C.; Scherer, M.; Reuter, N.; Lukassen, S.; Kasmapour, B.; Schreiner, S.; Cicin-Sain, L.; Winterpacht, A.; Stamminger, T. Chromatin-Remodeling Factor SPOC1 Acts as a Cellular Restriction Factor against Human Cytomegalovirus by Repressing the Major Immediate Early Promoter. J. Virol. 2018, 92, e00342-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Wu, C.-H.; Yeh, S.-H.; Tsay, Y.-G.; Shieh, Y.-H.; Kao, C.-L.; Chen, Y.-S.; Wang, S.-H.; Kuo, T.-J.; Chen, D.-S.; Chen, D.-S. Glycogen Synthase Kinase-3 Regulates the Phosphorylation of Severe Acute Respiratory Syndrome Coronavirus Nucleocapsid Protein and Viral Replication. J. Biol. Chem. 2008, 284, 5229–5239. [Google Scholar] [CrossRef] [Green Version]
  179. Verma, S.C.; Lan, K.; Robertson, E.S. Structure and Function of Latency-Associated Nuclear Antigen. Hantaviruses 2007, 312, 101–136. [Google Scholar]
  180. Bubman, D.; Guasparri, I.; Cesarman, E. Deregulation of c-Myc in primary effusion lymphoma by Kaposi’s sarcoma herpesvirus latency-associated nuclear antigen. Oncogene 2007, 26, 4979–4986. [Google Scholar] [CrossRef] [Green Version]
  181. Chin, R.; Earnest-Silveira, L.; Koeberlein, B.; Franz, S.; Zentgraf, H.; Dong, X.; Gowans, E.; Bock, C.-T.; Torresi, J. Modulation of MAPK pathways and cell cycle by replicating hepatitis B virus: Factors contributing to hepatocarcinogenesis. J. Hepatol. 2007, 47, 325–337. [Google Scholar] [CrossRef]
  182. Piracha, Z.Z.; Kwon, H.; Saeed, U.; Kim, J.; Jung, J.; Chwae, Y.J.; Park, S.; Shin, H.J.; Kim, K. Sirtuin 2 Isoform 1 Enhances Hepatitis B Virus RNA Transcription and DNA Synthesis through the AKT/GSK-3beta/beta-Catenin Signaling Pathway. J. Virol. 2018, 92, e00955-18. [Google Scholar] [CrossRef] [Green Version]
  183. Sun, L.; Lv, F.; Guo, X.; Gao, G. Glycogen synthase kinase 3beta (GSK3beta) modulates antiviral activity of zinc-finger antiviral protein (ZAP). J. Biol. Chem. 2012, 287, 22882–22888. [Google Scholar] [CrossRef] [Green Version]
  184. Yen, J.-H.; Kong, W.; Hooper, K.M.; Emig, F.; Rahbari, K.M.; Kuo, P.-C.; Scofield, B.A.; Ganea, D. Differential effects of IFN-β on IL-12, IL-23, and IL-10 expression in TLR-stimulated dendritic cells. J. Leukoc. Biol. 2015, 98, 689–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Li, A.; Ceballos-Diaz, C.; DiNunno, N.; Levites, Y.; Cruz, P.E.; Lewis, J.; Golde, T.E.; Chakrabarty, P. IFN-gamma promotes tau phosphorylation without affecting mature tangles. FASEB J. 2015, 29, 4384–4398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Tsai, C.C.; Huang, W.C.; Chen, C.L.; Hsieh, C.Y.; Lin, Y.S.; Chen, S.H.; Yang, K.C.; Lin, C.F. Glycogen synthase kinase-3 facilitates con a-induced IFN-gamma-- mediated immune hepatic injury. J. Immunol. 2011, 187, 3867–3877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Beurel, E.; Jope, R.S. Glycogen synthase kinase-3 promotes the synergistic action of interferon-γ on lipopolysaccharide-induced IL-6 production in RAW264.7 cells. Cell. Signal. 2009, 21, 978–985. [Google Scholar] [CrossRef] [Green Version]
  188. Lin, C.F.; Tsai, C.C.; Huang, W.C.; Wang, C.Y.; Tseng, H.C.; Wang, Y.; Kai, J.I.; Wang, S.W.; Cheng, Y.L. IFN-gamma synergizes with LPS to induce nitric oxide biosynthesis through glycogen synthase kinase-3-inhibited IL-10. J. Cell. Biochem. 2008, 105, 746–755. [Google Scholar] [CrossRef]
  189. Tsai, C.C.; Kai, J.I.; Huang, W.C.; Wang, C.Y.; Wang, Y.; Chen, C.L.; Fang, Y.T.; Lin, Y.S.; Anderson, R.; Chen, S.H.; et al. Glycogen synthase kinase-3beta facilitates IFN-gamma-induced STAT1 activation by regulating Src homology-2 domain-containing phosphatase 2. J. Immunol. 2009, 183, 856–864. [Google Scholar] [CrossRef] [Green Version]
  190. Ryu, S.; Lee, Y.; Hyun, M.Y.; Choi, S.Y.; Jeong, K.H.; Park, Y.M.; Kang, H.; Park, K.Y.; Armstrong, C.A.; Johnson, A.; et al. Mycophenolate antagonizes IFN-gamma-induced catagen-like changes via beta-catenin activation in human dermal papilla cells and hair follicles. Int. J. Mol. Sci. 2014, 15, 16800–16815. [Google Scholar] [CrossRef] [Green Version]
  191. Takahashi-Yanaga, F. Activator or inhibitor? GSK-3 as a new drug target. Biochem. Pharmacol. 2013, 86, 191–199. [Google Scholar] [CrossRef]
  192. Panza, F.; Solfrizzi, V.; Frisardi, V.; Imbimbo, B.P.; Capurso, C.; D’Introno, A.; Colacicco, A.M.; Seripa, D.; Vendemiale, G.; Capurso, A.; et al. Beyond the neurotransmitter-focused approach in treating Alzheimer’s disease: Drugs targeting beta-amyloid and tau protein. Aging Clin. Exp. Res. 2009, 21, 386–406. [Google Scholar] [CrossRef]
  193. Taylor, A.; Rudd, C.E. Small Molecule Inhibition of Glycogen Synthase Kinase-3 in Cancer Immunotherapy. Advances in Experimental Medicine and Biology 2019, 1164, 225–233. [Google Scholar]
  194. Bhat, R.V.; Andersson, U.; Andersson, S.; Knerr, L.; Bauer, U.; Sundgren-Andersson, A.K. The Conundrum of GSK3 Inhibitors: Is it the Dawn of a New Beginning? J. Alzheimer’s Dis. 2018, 64, S547–S554. [Google Scholar] [CrossRef] [PubMed]
  195. Eldar-Finkelman, H.; Martínez, A. GSK-3 Inhibitors: Preclinical and Clinical Focus on CNS. Front. Mol. Neurosci. 2011, 4, 32. [Google Scholar] [CrossRef] [Green Version]
  196. Duthie, A.; Van Aalten, L.; Macdonald, C.; McNeilly, A.; Gallagher, J.; Geddes, J.; Lovestone, S.; Sutherland, C. Recruitment, Retainment, and Biomarkers of Response; A Pilot Trial of Lithium in Humans With Mild Cognitive Impairment. Front. Mol. Neurosci. 2019, 12, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. U.S. National Library of Medicine Database of Clinical Trails. Available online: https://clinicaltrials.gov/ (accessed on 25 March 2020).
  198. Nam, D.; Balasuberamaniam, P.; Milner, K.; Kunz, M.; Vachhani, K.; Kiss, A.; Whyne, C. Lithium for Fracture Treatment (LiFT): A double-blind randomised control trial protocol. BMJ Open 2020, 10, e031545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Zamek-Gliszczynski, M.J.; Abraham, T.L.; Alberts, J.J.; Kulanthaivel, P.; Jackson, K.A.; Chow, K.H.; McCann, D.J.; Hu, H.; Anderson, S.; Furr, N.A.; et al. Pharmacokinetics, Metabolism, and Excretion of the Glycogen Synthase Kinase-3 Inhibitor LY2090314 in Rats, Dogs, and Humans: A Case Study in Rapid Clearance by Extensive Metabolism with Low Circulating Metabolite Exposure. Drug Metab. Dispos. 2013, 41, 714–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Gray, J.E.; Infante, J.R.; Brail, L.H.; Simon, G.R.; Cooksey, J.F.; Jones, S.F.; Farrington, D.L.; Yeo, A.; Jackson, K.A.; Chow, K.H.; et al. A first-in-human phase I dose-escalation, pharmacokinetic, and pharmacodynamic evaluation of intravenous LY2090314, a glycogen synthase kinase 3 inhibitor, administered in combination with pemetrexed and carboplatin. Investig. New Drugs 2015, 33, 1187–1196. [Google Scholar] [CrossRef] [PubMed]
  201. Rizzieri, D.A.; Cooley, S.; Odenike, O.; Moonan, L.; Chow, K.H.; Jackson, K.; Wang, X.; Brail, L.; Borthakur, G. An open-label phase 2 study of glycogen synthase kinase-3 inhibitor LY2090314 in patients with acute leukemia. Leuk. Lymphoma 2016, 57, 1800–1806. [Google Scholar] [CrossRef]
  202. Georgievska, B.; Sandin, J.; Doherty, J.; Mörtberg, A.; Neelissen, J.; Andersson, A.; Gruber, S.; Nilsson, Y.; Schött, P.; I Arvidsson, P.; et al. AZD1080, a novel GSK3 inhibitor, rescues synaptic plasticity deficits in rodent brain and exhibits peripheral target engagement in humans. J. Neurochem. 2013, 125, 446–456. [Google Scholar] [CrossRef]
  203. Pal, K.; Cao, Y.; Gaisina, I.; Bhattacharya, S.; Dutta, S.K.; Wang, E.; Gunosewoyo, H.; Kozikowski, A.P.; Billadeau, D.D.; Mukhopadhyay, D. Inhibition of GSK-3 induces differentiation and impaired glucose metabolism in renal cancer. Mol. Cancer Ther. 2013, 13, 285–296. [Google Scholar] [CrossRef] [Green Version]
  204. Ugolkov, A.; Gaisina, I.; Zhang, J.-S.; Billadeau, D.D.; White, K.; Kozikowski, A.; Jain, S.; Cristofanilli, M.; Giles, F.; O’Halloran, T.; et al. GSK-3 inhibition overcomes chemoresistance in human breast cancer. Cancer Lett. 2016, 380, 384–392. [Google Scholar] [CrossRef]
  205. Kuroki, H.; Anraku, T.; Kazama, A.; Bilim, V.; Tasaki, M.; Schmitt, D.; Mazar, A.P.; Giles, F.J.; Ugolkov, A.; Tomita, Y. 9-ING-41, a small molecule inhibitor of GSK-3beta, potentiates the effects of anticancer therapeutics in bladder cancer. Sci. Rep. 2019, 9, 19977–19979. [Google Scholar] [CrossRef] [PubMed]
  206. Dominguez, J.M.; Fuertes, A.; Orozco, L.; del Monte-Millan, M.; Delgado, E.; Medina, M. Evidence for irreversible inhibition of glycogen synthase kinase-3beta by tideglusib. J. Biol. Chem. 2012, 287, 893–904. [Google Scholar] [CrossRef] [Green Version]
  207. Lovestone, S.; Boada, M.; Dubois, B.; Hull, M.; Rinne, J.O.; Huppertz, H.J.; Calero, M.; Andres, M.V.; Gomez-Carrillo, B.; Leon, T.; et al. A phase II trial of tideglusib in Alzheimer’s disease. J. Alzheimers Dis. 2015, 45, 75–88. [Google Scholar] [CrossRef] [PubMed]
  208. Tolosa, E.; Litvan, I.; Höglinger, G.U.; Burn, D.; Lees, A.; Andrés, M.V.; Bsc, B.G.; Leon, T.; Del Ser, T.; TAUROS MRI Investigators; et al. A phase 2 trial of the GSK-3 inhibitor tideglusib in progressive supranuclear palsy. Mov. Disord. 2014, 29, 470–478. [Google Scholar] [CrossRef] [PubMed]
  209. Del Ser, T.; Steinwachs, K.C.; Gertz, H.J.; Andres, M.V.; Gomez-Carrillo, B.; Medina, M.; Vericat, J.A.; Redondo, P.; Fleet, D.; Leon, T. Treatment of Alzheimer’s disease with the GSK-3 inhibitor tideglusib: A pilot study. J. Alzheimers Dis. 2013, 33, 205–215. [Google Scholar] [CrossRef]
  210. Höglinger, G.U.; Huppertz, H.-J.; Wagenpfeil, S.; Andrés, M.V.; Belloch, V.; Leon, T.; Del Ser, T.; TAUROS MRI Investigators; Gmez, J.C.; Tijero, B.; et al. Tideglusib reduces progression of brain atrophy in progressive supranuclear palsy in a randomized trial. Mov. Disord. 2014, 29, 479–487. [Google Scholar]
  211. Boyle, W.J.; Smeal, T.; Defize, L.H.; Angel, P.; Woodgett, J.R.; Karin, M.; Hunter, T. Activation of protein kinase C decreases phosphorylation of c-Jun at sites that negatively regulate its DNA-binding activity. Cell 1991, 64, 573–584. [Google Scholar] [CrossRef]
  212. Nikolakaki, E.; Coffer, P.J.; Hemelsoet, R.; Woodgett, J.R.; Defize, L.H. Glycogen synthase kinase 3 phosphorylates Jun family members in vitro and negatively regulates their transactivating potential in intact cells. Oncogene 1993, 8, 833–840. [Google Scholar]
  213. De Groot, R.P.; Auwerx, J.; Bourouis, M.; Sassone-Corsi, P. Negative regulation of Jun/AP-1: Conserved function of glycogen synthase kinase 3 and the Drosophila kinase shaggy. Oncogene 1993, 8, 841–847. [Google Scholar]
  214. A Grimes, C.; Jope, R.S. The multifaceted roles of glycogen synthase kinase 3β in cellular signaling. Prog. Neurobiol. 2001, 65, 391–426. [Google Scholar] [CrossRef]
  215. Beurel, E.; Jope, R.S. Differential regulation of STAT family members by glycogen synthase kinase-3. J. Biol. Chem. 2008, 283, 21934–21944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Moens, U.; Fbsbioscience.Org. Multisite phosphorylation of the cAMP response element-binding protein CREB by a diversity of protein kinases. Front. Biosci. 2007, 12, 1814–1832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Sun, S.C. The non-canonical NF-kappaB pathway in immunity and inflammation. Nat. Rev. Immunol. 2017, 17, 545–558. [Google Scholar] [CrossRef] [PubMed]
  218. Tullai, J.W.; Graham, J.R.; Cooper, G.M. A GSK-3-mediated transcriptional network maintains repression of immediate early genes in quiescent cells. Cell Cycle 2011, 10, 3072–3077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Schwabe, R.F.; Brenner, D.A. Role of glycogen synthase kinase-3 in TNF-alpha-induced NF-kappaB activation and apoptosis in hepatocytes. Am. J. Physiol. Gastrointest Liver Physiol. 2002, 283, G204–G211. [Google Scholar] [CrossRef] [PubMed]
  220. Steinbrecher, K.A.; Wilson, W., 3rd; Cogswell, P.C.; Baldwin, A.S. Glycogen synthase kinase 3beta functions to specify gene-specific, NF-kappaB-dependent transcription. Mol. Cell. Biol. 2005, 25, 8444–8455. [Google Scholar] [CrossRef] [Green Version]
  221. Deng, J.; A Miller, S.; Wang, H.; Xia, W.; Wen, Y.; Zhou, B.P.; Li, Y.; Lin, S.-Y.; Hung, M.-C. beta-catenin interacts with and inhibits NF-kappa B in human colon and breast cancer. Cancer Cell 2002, 2, 323–334. [Google Scholar] [CrossRef] [Green Version]
  222. Wilson, W., 3rd; Baldwin, A.S. Maintenance of constitutive IkappaB kinase activity by glycogen synthase kinase-3alpha/beta in pancreatic cancer. Cancer Res. 2008, 68, 8156–8163. [Google Scholar] [CrossRef] [Green Version]
  223. Eto, M.; Kouroedov, A.; Cosentino, F.; Luüscher, T.F. Glycogen Synthase Kinase-3 Mediates Endothelial Cell Activation by Tumor Necrosis Factor-α. Circ. 2005, 112, 1316–1322. [Google Scholar] [CrossRef] [Green Version]
  224. Zhang, J.S.; Herreros-Villanueva, M.; Koenig, A.; Deng, Z.; de Narvajas, A.A.; Gomez, T.S.; Meng, X.; Bujanda, L.; Ellenrieder, V.; Li, X.K.; et al. Differential activity of GSK-3 isoforms regulates NF-kappaB and TRAIL- or TNFalpha induced apoptosis in pancreatic cancer cells. Cell Death Dis. 2014, 5, e1142. [Google Scholar] [CrossRef]
  225. Graham, J.R.; Tullai, J.W.; Cooper, G.M. GSK-3 represses growth factor-inducible genes by inhibiting NF-kappaB in quiescent cells. J. Biol. Chem. 2010, 285, 4472–4480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Sanchez, J.F.; Sniderhan, L.F.; Williamson, A.L.; Fan, S.; Chakraborty-Sett, S.; Maggirwar, S.B. Glycogen synthase kinase 3beta-mediated apoptosis of primary cortical astrocytes involves inhibition of nuclear factor kappaB signaling. Mol. Cell. Biol. 2003, 23, 4649–4662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Ma, Y.; Wang, M.; Li, N.; Wu, R.; Wang, X. Bleomycin-induced nuclear factor-kappaB activation in human bronchial epithelial cells involves the phosphorylation of glycogen synthase kinase 3beta. Toxicol. Lett. 2009, 187, 194–200. [Google Scholar] [CrossRef] [PubMed]
  228. Itoh, S.; Saito, T.; Hirata, M.; Ushita, M.; Ikeda, T.; Woodgett, J.R.; Algul, H.; Schmid, R.M.; Chung, U.I.; Kawaguchi, H. GSK-3alpha and GSK-3beta proteins are involved in early stages of chondrocyte differentiation with functional redundancy through RelA protein phosphorylation. J. Biol. Chem. 2012, 287, 29227–29236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Agarwal, D.; Dange, R.B.; Raizada, M.K.; Francis, J. Angiotensin II causes imbalance between pro- and anti-inflammatory cytokines by modulating GSK-3beta in neuronal culture. Br. J. Pharmacol. 2013, 169, 860–874. [Google Scholar] [CrossRef] [Green Version]
  230. Bao, H.; Ge, Y.; Peng, A.; Gong, R. Fine-tuning of NFkappaB by glycogen synthase kinase 3beta directs the fate of glomerular podocytes upon injury. Kidney Int. 2015, 87, 1176–1190. [Google Scholar] [CrossRef] [Green Version]
  231. Buss, H.; Dorrie, A.; Schmitz, M.L.; Frank, R.; Livingstone, M.; Resch, K.; Kracht, M. Phosphorylation of serine 468 by GSK-3beta negatively regulates basal p65 NF-kappaB activity. J. Biol. Chem. 2004, 279, 49571–49574. [Google Scholar] [CrossRef] [Green Version]
  232. Gong, R.; Rifai, A.; Ge, Y.; Chen, S.; Dworkin, L.D. Hepatocyte growth factor suppresses proinflammatory NFkappaB activation through GSK3beta inactivation in renal tubular epithelial cells. J. Biol. Chem. 2008, 283, 7401–7410. [Google Scholar] [CrossRef] [Green Version]
  233. Zhang, L.; Zhang, J.; Chen, Z.; Wang, L.; Wu, X.; Ou, M. Epidermal growth factor (EGF) triggers the malignancy of hemangioma cells via activation of NF-kappaB signals. Biomed. Pharmacother. 2016, 82, 133–140. [Google Scholar] [CrossRef]
  234. Demarchi, F.; Bertoli, C.; Sandy, P.; Schneider, C. Glycogen synthase kinase-3 beta regulates NF-kappa B1/p105 stability. J. Biol. Chem. 2003, 278, 39583–39590. [Google Scholar] [CrossRef] [Green Version]
  235. Busino, L.; Millman, S.E.; Scotto, L.; Kyratsous, C.A.; Basrur, V.; O’Connor, O.; Hoffmann, A.; Elenitoba-Johnson, K.; Pagano, M. Fbxw7α- and GSK3-mediated degradation of p100 is a pro-survival mechanism in multiple myeloma. Nature 2012, 14, 375–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. PhosphoSitePlus. Available online: https://www.phosphosite.org/ (accessed on 11 March 2020).
  237. Neumann, M.; Klar, S.; Wilisch-Neumann, A.; Hollenbach, E.; Kavuri, S.; Leverkus, M.; Kandolf, R.; Brunner-Weinzierl, M.C.; Klingel, K. Glycogen synthase kinase-3beta is a crucial mediator of signal-induced RelB degradation. Oncogene 2011, 30, 2485–2492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Ghosh, S.; Hayden, M. New regulators of NF-κB in inflammation. Nat. Rev. Immunol. 2008, 8, 837–848. [Google Scholar] [CrossRef] [PubMed]
  239. Viatour, P.; Dejardin, E.; Warnier, M.; Lair, F.; Claudio, E.; Bureau, F.; Marine, J.-C.; Merville, M.-P.; Maurer, U.; Green, D.; et al. GSK3-Mediated BCL-3 Phosphorylation Modulates Its Degradation and Its Oncogenicity. Mol. Cell 2004, 16, 35–45. [Google Scholar] [CrossRef] [PubMed]
  240. Medunjanin, S.; Schleithoff, L.; Fiegehenn, C.; Weinert, S.; Zuschratter, W.; Braun-Dullaeus, R.C. GSK-3beta controls NF-kappaB activity via IKKgamma/NEMO. Sci. Rep. 2016, 6, 38553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Wei, W.; Jin, J.; Schlisio, S.; Harper, J.W.; Kaelin, W.G. The v-Jun point mutation allows c-Jun to escape GSK3-dependent recognition and destruction by the Fbw7 ubiquitin ligase. Cancer Cell 2005, 8, 25–33. [Google Scholar] [CrossRef] [Green Version]
  242. Pérez-Benavente, B.; García, J.L.; Rodríguez, M.S.; Pineda-Lucena, A.; Piechaczyk, M.; De Mora, J.F.; Farràs, R. GSK3-SCFFBXW7 targets JunB for degradation in G2 to preserve chromatid cohesion before anaphase. Oncogene 2012, 32, 2189–2199. [Google Scholar] [CrossRef]
  243. Franciscovich, A.L.; Mortimer, A.D.; Freeman, A.A.; Gu, J.; Sanyal, S. Overexpression screen in Drosophila identifies neuronal roles of GSK-3 beta/shaggy as a regulator of AP-1-dependent developmental plasticity. Genetics 2008, 180, 2057–2071. [Google Scholar] [CrossRef] [Green Version]
  244. Tullai, J.W.; Tacheva, S.; Owens, L.J.; Graham, J.R.; Cooper, G.M. AP-1 Is a Component of the Transcriptional Network Regulated by GSK-3 in Quiescent Cells. PLoS ONE 2011, 6, 20150. [Google Scholar] [CrossRef] [Green Version]
  245. Troussard, A.A.; Tan, C.; Yoganathan, T.N.; Dedhar, S. Cell-Extracellular Matrix Interactions Stimulate the AP-1 Transcription Factor in an Integrin-Linked Kinase- and Glycogen Synthase Kinase 3-Dependent Manner. Mol. Cell. Biol. 1999, 19, 7420–7427. [Google Scholar] [CrossRef] [Green Version]
  246. Kitaura, J.; Asai, K.; Maeda-Yamamoto, M.; Kawakami, Y.; Kikkawa, U.; Kawakami, T. Akt-Dependent Cytokine Production in Mast Cells. J. Exp. Med. 2000, 192, 729–740. [Google Scholar] [CrossRef] [PubMed]
  247. Kikuchi, A. Roles of Axin in the Wnt signalling pathway. Cell. Signal. 1999, 11, 777–788. [Google Scholar] [CrossRef]
  248. Gonzales, M.; Bowden, G. The role of PI 3-kinase in the UVB-induced expression of c-fos. Oncogene 2002, 21, 2721–2728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Zhuang, S.; Hua, X.; He, K.; Zhou, T.; Zhang, J.; Wu, H.; Ma, X.; Xia, Q.; Zhang, J. Inhibition of GSK-3beta induces AP-1-mediated osteopontin expression to promote cholestatic liver fibrosis. FASEB J. 2018, 32, 4494–4503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Wang, M.J.; Huang, H.Y.; Chen, W.F.; Chang, H.F.; Kuo, J.S. Glycogen synthase kinase-3beta inactivation inhibits tumor necrosis factor-alpha production in microglia by modulating nuclear factor kappaB and MLK3/JNK signaling cascades. J. Neuroinflammation 2010, 7, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Ross, S.E.; Erickson, R.L.; Hemati, N.; MacDougald, O. Glycogen Synthase Kinase 3 Is an Insulin-Regulated C/EBP? Kinase. Mol. Cell. Biol. 1999, 19, 8433–8441. [Google Scholar] [CrossRef] [Green Version]
  252. Chen, H.; Fajol, A.; Hoene, M.; Zhang, B.; Schleicher, E.D.; Lin, Y.; Calaminus, C.; Pichler, B.J.; Weigert, C.; Häring, H.U.; et al. PI3K-resistant GSK3 controls adiponectin formation and protects from metabolic syndrome. Proc. Natl. Acad. Sci. USA 2016, 113, 5754–5759. [Google Scholar] [CrossRef] [Green Version]
  253. Liu, H.K.; Perrier, S.; Lipina, C.; Finlay, D.; McLauchlan, H.; Hastie, C.J.; Hundal, H.S.; Sutherland, C. Functional characterisation of the regulation of CAAT enhancer binding protein alpha by GSK-3 phosphorylation of Threonines 222/226. BMC Mol. Biol. 2006, 7, 14. [Google Scholar] [CrossRef] [Green Version]
  254. Datta, J.; Majumder, S.; Kutay, H.; Motiwala, T.; Frankel, W.; Costa, R.; Cha, H.C.; MacDougald, O.; Jacob, S.T.; Ghoshal, K. Metallothionein expression is suppressed in primary human hepatocellular carcinomas and is mediated through inactivation of CCAAT/enhancer binding protein alpha by phosphatidylinositol 3-kinase signaling cascade. Cancer Res. 2007, 67, 2736–2746. [Google Scholar] [CrossRef] [Green Version]
  255. Piwien-Pilipuk, G.; Van Mater, D.; Ross, S.E.; MacDougald, O.A.; Schwartz, J. Growth hormone regulates phosphorylation and function of CCAAT/enhancer-binding protein beta by modulating Akt and glycogen synthase kinase-3. J. Biol. Chem. 2001, 276, 19664–19671. [Google Scholar] [CrossRef] [Green Version]
  256. Park, B.H.; Qiang, L.; Farmer, S.R. Phosphorylation of C/EBPbeta at a consensus extracellular signal-regulated kinase/glycogen synthase kinase 3 site is required for the induction of adiponectin gene expression during the differentiation of mouse fibroblasts into adipocytes. Mol. Cell. Biol. 2004, 24, 8671–8680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Larabee, J.L.; Shakir, S.M.; Hightower, L.; Ballard, J.D. Adenomatous polyposis coli protein associates with C/EBP beta and increases Bacillus anthracis edema toxin-stimulated gene expression in macrophages. J. Biol. Chem. 2011, 286, 19364–19372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Kim, J.W.; Tang, Q.Q.; Li, X.; Lane, M.D. Effect of phosphorylation and S–S bond-induced dimerization on DNA binding and transcriptional activation by C/EBPbeta. Proc. Natl. Acad. Sci. USA 2007, 104, 1800–1804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Welz, B.; Bikker, R.; Junemann, J.; Christmann, M.; Neumann, K.; Weber, M.; Hoffmeister, L.; Preuß, K.; Pich, A.; Huber, R.; et al. Proteome and Phosphoproteome Analysis in TNF Long Term-Exposed Primary Human Monocytes. Int. J. Mol. Sci. 2019, 20, 1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Hamel-Côté, G.; Lapointe, F.; Gendron, D.; Rola-Pleszczynski, M.; Stankova, J. Regulation of platelet-activating factor-induced interleukin-8 expression by protein tyrosine phosphatase 1B. Cell Commun. Signal. 2019, 17, 21. [Google Scholar] [CrossRef] [Green Version]
  261. Shen, F.; Li, N.; Gade, P.; Kalvakolanu, D.V.; Weibley, T.; Doble, B.W.; Woodgett, J.R.; Wood, T.; Gaffen, S.L. IL-17 Receptor Signaling Inhibits C/EBP by Sequential Phosphorylation of the Regulatory 2 Domain. Sci. Signal. 2009, 2, ra8. [Google Scholar] [CrossRef] [Green Version]
  262. Maekawa, T.; Hosur, K.; Abe, T.; Kantarci, A.; Ziogas, A.; Wang, B.; Van Dyke, T.E.; Chavakis, T.; Hajishengallis, G. Antagonistic effects of IL-17 and D-resolvins on endothelial Del-1 expression through a GSK-3beta-C/EBPbeta pathway. Nat. Commun. 2015, 6, 8272. [Google Scholar] [CrossRef] [Green Version]
  263. Li, X.; Molina, H.; Huang, H.; Zhang, Y.Y.; Liu, M.; Qian, S.W.; Slawson, C.; Dias, W.B.; Pandey, A.; Hart, G.W.; et al. O-linked N-acetylglucosamine modification on CCAAT enhancer-binding protein beta: Role during adipocyte differentiation. J. Biol. Chem. 2009, 284, 19248–19254. [Google Scholar] [CrossRef] [Green Version]
  264. Zhao, X.; Zhuang, S.; Chen, Y.; Boss, G.R.; Pilz, R.B. Cyclic GMP-dependent protein kinase regulates CCAAT enhancer-binding protein beta functions through inhibition of glycogen synthase kinase-3. J. Biol. Chem. 2005, 280, 32683–32692. [Google Scholar] [CrossRef] [Green Version]
  265. Huang, W.-C.; Lin, Y.-S.; Wang, C.-Y.; Tsai, C.-C.; Tseng, H.-C.; Chen, C.-L.; Lu, P.-J.; Chen, P.-S.; Qian, L.; Hong, J.-S.; et al. Glycogen synthase kinase-3 negatively regulates anti-inflammatory interleukin-10 for lipopolysaccharide-induced iNOS/NO biosynthesis and RANTES production in microglial cells. Immunology 2009, 128, e275–e286. [Google Scholar] [CrossRef]
  266. Ko, C.Y.; Wang, W.L.; Wang, S.M.; Chu, Y.Y.; Chang, W.C.; Wang, J.M. Glycogen synthase kinase-3beta-mediated CCAAT/enhancer-binding protein delta phosphorylation in astrocytes promotes migration and activation of microglia/macrophages. Neurobiol. Aging 2014, 35, 24–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Zhu, Q.; Yang, J.; Han, S.; Liu, J.; Holzbeierlein, J.; Thrasher, J.B.; Li, B. Suppression of glycogen synthase kinase 3 activity reduces tumor growth of prostate cancer in vivo. Prostate 2010, 71, 835–845. [Google Scholar] [CrossRef] [PubMed]
  268. Rhee, Y.-H.; Ahn, J.-C. Melatonin attenuated adipogenesis through reduction of the CCAAT/enhancer binding protein beta by regulating the glycogen synthase 3 beta in human mesenchymal stem cells. J. Physiol. Biochem. 2016, 72, 145–155. [Google Scholar] [CrossRef]
  269. Balamurugan, K.; Sharan, S.; Klarmann, K.D.; Zhang, Y.; Coppola, V.; Summers, G.H.; Roger, T.; Morrison, D.K.; Keller, J.R.; Sterneck, E. FBXW7alpha attenuates inflammatory signalling by downregulating C/EBPdelta and its target gene Tlr4. Nat. Commun. 2013, 4, 1662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Li, M.; Zhou, W.; Yuan, R.; Chen, L.; Liu, T.; Huang, D.; Hao, L.; Xie, Y.; Shao, J. ROCK2 promotes HCC proliferation by CEBPD inhibition through phospho-GSK3beta/beta-catenin signaling. FEBS Lett. 2015, 589, 1018–1025. [Google Scholar] [CrossRef] [Green Version]
  271. Chen, G.; Bower, K.A.; Ma, C.; Fang, S.; Thiele, C.J.; Luo, J. Glycogen synthase kinase 3beta (GSK3beta) mediates 6-hydroxydopamine-induced neuronal death. FASEB J. 2004, 18, 1162–1164. [Google Scholar] [CrossRef] [PubMed]
  272. McAlpine, C.S.; Werstuck, G.H. Protein kinase R-like endoplasmic reticulum kinase and glycogen synthase kinase-3alpha/beta regulate foam cell formation. J. Lipid Res. 2014, 55, 2320–2333. [Google Scholar] [CrossRef] [Green Version]
  273. Deng, J.; Wang, X.; Qian, F.; Vogel, S.; Xiao, L.; Ranjan, R.; Park, H.; Karpurapu, M.; Ye, R.D.; Park, G.Y.; et al. Protective role of reactive oxygen species in endotoxin-induced lung inflammation through modulation of IL-10 expression. J. Immunol. 2012, 188, 5734–5740. [Google Scholar] [CrossRef] [Green Version]
  274. Dugo, L.; Collin, M.; Allen, D.A.; A Patel, N.S.; Bauer, I.; A Mervaala, E.M.; Louhelainen, M.; Foster, S.J.; Yaqoob, M.M.; Thiemermann, C. GSK-3beta inhibitors attenuate the organ injury/dysfunction caused by endotoxemia in the rat. Crit. Care Med. 2005, 33, 1903–1912. [Google Scholar] [CrossRef]
  275. Sathiya Priya, C.; Vidhya, R.; Kalpana, K.; Anuradha, C.V. Indirubin-3’-monoxime prevents aberrant activation of GSK-3beta/NF-kappaB and alleviates high fat-high fructose induced Abeta-aggregation, gliosis and apoptosis in mice brain. Int. Immunopharmacol. 2019, 70, 396–407. [Google Scholar] [CrossRef]
  276. Kerr, F.; Bjedov, I.; Sofola-Adesakin, O. Molecular Mechanisms of Lithium Action: Switching the Light on Multiple Targets for Dementia Using Animal Models. Front. Mol. Neurosci. 2018, 11, 297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Greenlee-Wacker, M.C. Clearance of apoptotic neutrophils and resolution of inflammation. Immunol. Rev. 2016, 273, 357–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Reville, K.; Crean, J.K.; Vivers, S.; Dransfield, I.; Godson, C. Lipoxin A4 redistributes myosin IIA and Cdc42 in macrophages: Implications for phagocytosis of apoptotic leukocytes. J. Immunol. 2006, 176, 1878–1888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Das, A.; Ganesh, K.; Khanna, S.; Sen, C.K.; Roy, S. Engulfment of apoptotic cells by macrophages: A role of microRNA-21 in the resolution of wound inflammation. J. Immunol. 2014, 192, 1120–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Zhu, M.; Tian, D.; Li, J.; Ma, Y.; Wang, Y.; Wu, R. Glycogen synthase kinase 3β and β-catenin are involved in the injury and repair of bronchial epithelial cells induced by scratching. Exp. Mol. Pathol. 2007, 83, 30–38. [Google Scholar] [CrossRef]
  281. Karrasch, T.; Spaeth, T.; Allard, B.; Jobin, C. PI3K-Dependent GSK3ß(Ser9)-Phosphorylation Is Implicated in the Intestinal Epithelial Cell Wound-Healing Response. PLoS ONE 2011, 6, e26340. [Google Scholar] [CrossRef]
  282. Yang, H.L.; Tsai, Y.C.; Korivi, M.; Chang, C.T.; Hseu, Y.C. Lucidone Promotes the Cutaneous Wound Healing Process via Activation of the PI3K/AKT, Wnt/beta-catenin and NF-kappaB Signaling Pathways. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 151–168. [Google Scholar] [CrossRef]
  283. Potey, P.M.; Rossi, A.G.; Lucas, C.; Dorward, D.A. Neutrophils in the initiation and resolution of acute pulmonary inflammation: Understanding biological function and therapeutic potential. J. Pathol. 2019, 247, 672–685. [Google Scholar] [CrossRef]
  284. Hornstein, T.; Lehmann, S.; Philipp, D.; Detmer, S.; Hoffmann, M.; Peter, C.; Wesselborg, S.; Unfried, K.; Windolf, J.; Flohe, S.; et al. Staurosporine resistance in inflammatory neutrophils is associated with the inhibition of caspase- and proteasome-mediated Mcl-1 degradation. J. Leukoc. Biol. 2016, 99, 163–174. [Google Scholar] [CrossRef]
  285. Liu, H.; Mi, S.; Li, Z.; Hua, F.; Hu, Z.-W. Interleukin 17A inhibits autophagy through activation of PIK3CA to interrupt the GSK3B-mediated degradation of BCL2 in lung epithelial cells. Autophagy 2013, 9, 730–742. [Google Scholar] [CrossRef] [Green Version]
  286. Huber, R.; Bikker, R.; Welz, B.; Christmann, M.; Brand, K. TNF Tolerance in Monocytes and Macrophages: Characteristics and Molecular Mechanisms. J. Immunol. Res. 2017, 2017, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Weber, M.; Sydlik, C.; Quirling, M.; Nothdurfter, C.; Zwergal, A.; Heiss, P.; Bell, S.; Neumeier, D.; Ziegler-Heitbrock, H.W.L.; Brand, K.; et al. Transcriptional Inhibition of Interleukin-8 Expression in Tumor Necrosis Factor-tolerant Cells. J. Biol. Chem. 2003, 278, 23586–23593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Park, S.H.; Park-Min, K.-H.; Chen, J.; Hu, X.; Ivashkiv, L.B. Tumor necrosis factor induces GSK3 kinase–mediated cross-tolerance to endotoxin in macrophages. Nat. Immunol. 2011, 12, 607–615. [Google Scholar] [CrossRef] [PubMed]
  289. Günther, J.; Vogt, N.; Hampel, K.; Bikker, R.; Page, S.; Müller, B.; Kandemir, J.; Kracht, M.; Dittrich-Breiholz, O.; Huber, R.; et al. Identification of Two Forms of TNF Tolerance in Human Monocytes: Differential Inhibition of NF-κB/AP-1– and PP1-Associated Signaling. J. Immunol. 2014, 192, 3143–3155. [Google Scholar] [CrossRef] [PubMed]
  290. Zwergal, A.; Quirling, M.; Saugel, B.; Huth, K.C.; Sydlik, C.; Poli, V.; Neumeier, D.; Ziegler-Heitbrock, H.W.L.; Brand, K. C/EBP beta blocks p65 phosphorylation and thereby NF-kappa B-mediated transcription in TNF-tolerant cells. J. Immunol. 2006, 177, 665–672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Bikker, R.; Christmann, M.; Preuß, K.; Welz, B.; Friesenhagen, J.; Dittrich-Breiholz, O.; Huber, R.; Brand, K.; Preuss, K.; Huber, R. TNF phase III signalling in tolerant cells is tightly controlled by A20 and CYLD. Cell. Signal. 2017, 37, 123–135. [Google Scholar] [CrossRef]
Table 1. Molecular events regulated by GSK3 that mediate pro- or anti-inflammatory effects.
Table 1. Molecular events regulated by GSK3 that mediate pro- or anti-inflammatory effects.
Predominantly Pro-Inflammatory Effects.
(Due to Active GSK3)
Predominantly Anti-Inflammatory Effects
(Due to GSK3 Inhibition)
NF-κB DNA binding activity ↑NF-κB DNA binding activity ↓
NF-κB transactivation activity ↑
(pro-inflammatory genes)
NF-κB transactivation activity ↓
(pro-inflammatory genes)
p65 phosphorylation (Thr254, Ser276, Ser536) ↑p65 phosphorylation (Ser468) ↑
p105 phosphorylation (Ser903, Ser907)* ↑p105 stability ↑
p100 phosphorylation (Ser707)* ↑p100 stability ↑
NEMO phosphorylation (Ser8, Ser17, Ser31) ↑-
RelB phosphorylation (Ser552)* ↑RelB stability ↑
Bcl-3 phosphorylation (Ser394, Ser398)* ↑Bcl-3 stability ↑
Jun/Fos phosphorylation ↑Jun/Fos phosphorylation ↓
AP-1 DNA binding activity ↓AP-1 DNA binding activity ↑
C/EBPβ-LIP DNA binding activity ↑
(anti-inflammatory genes)
C/EBPβ-LAP DNA binding activity ↑
(anti-inflammatory genes)
C/EBPβ transactivation activity ↓
(anti-inflammatory genes)
C/EBPβ transactivation activity ↑
(anti-inflammatory genes)
C/EBPβ phosphorylation
(Thr188, Ser184, Ser179) ↑
C/EBPβ phosphorylation
(Thr188, Ser184, Ser179) ↓
CREB activity ↓CREB activity ↑
* leads to subsequent degradation. ↑, increased; ↓, decreased.

Share and Cite

MDPI and ACS Style

Hoffmeister, L.; Diekmann, M.; Brand, K.; Huber, R. GSK3: A Kinase Balancing Promotion and Resolution of Inflammation. Cells 2020, 9, 820. https://0-doi-org.brum.beds.ac.uk/10.3390/cells9040820

AMA Style

Hoffmeister L, Diekmann M, Brand K, Huber R. GSK3: A Kinase Balancing Promotion and Resolution of Inflammation. Cells. 2020; 9(4):820. https://0-doi-org.brum.beds.ac.uk/10.3390/cells9040820

Chicago/Turabian Style

Hoffmeister, Leonie, Mareike Diekmann, Korbinian Brand, and René Huber. 2020. "GSK3: A Kinase Balancing Promotion and Resolution of Inflammation" Cells 9, no. 4: 820. https://0-doi-org.brum.beds.ac.uk/10.3390/cells9040820

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