Skip to main content

An overview of CCN4 (WISP1) role in human diseases

Abstract

CCN4 (cellular communication network factor 4), a highly conserved, secreted cysteine-rich matricellular protein is emerging as a key player in the development and progression of numerous disease pathologies, including cancer, fibrosis, metabolic and inflammatory disorders. Over the past two decades, extensive research on CCN4 and its family members uncovered their diverse cellular mechanisms and biological functions, including but not limited to cell proliferation, migration, invasion, angiogenesis, wound healing, repair, and apoptosis. Recent studies have demonstrated that aberrant CCN4 expression and/or associated downstream signaling is key to a vast array of pathophysiological etiology, suggesting that CCN4 could be utilized not only as a non-invasive diagnostic or prognostic marker, but also as a promising therapeutic target. The cognate receptor of CCN4 remains elusive till date, which limits understanding of the mechanistic insights on CCN4 driven disease pathologies. However, as therapeutic agents directed against CCN4 begin to make their way into the clinic, that may start to change. Also, the pathophysiological significance of CCN4 remains underexplored, hence further research is needed to shed more light on its disease and/or tissue specific functions to better understand its clinical translational benefit. This review highlights the compelling evidence of overlapping and/or diverse functional and mechanisms regulated by CCN4, in addition to addressing the challenges, study limitations and knowledge gaps on CCN4 biology and its therapeutic potential.

Introduction

Extracellular matrix (ECM) proteins, including but not limited to collagens, fibronectins, and elastin, provide structural stability and physical framework for cellular organization across all mammalian tissues and organs. In addition to physical scaffolding, ECM proteins play a crucial role in regulating cellular processes by interacting with cell surface receptors, cytokines, growth factors and other extracellular proteins. These regulatory functions are primarily driven by a subgroup of ECM proteins known as matricellular proteins which are non-structural in nature. Matricellular proteins, such as secreted protein acidic and rich in cysteine (SPARC)-protein family, thrombospondin, periostin, osteopontin and fibulins amongst others are multidomain proteins that are secreted in the ECM and are critical for day-to-day physiological processes, including cell proliferation, migration, and adhesion [1].

In addition to other known matricellular proteins, CCN family members are highly conserved, secreted multi-modular cysteine rich proteins, composed of six homologous proteins. The first three family members namely, (i) cysteine-rich protein 61 (Cyr61), (ii) connective tissue growth factor (CTGF) and (iii) nephroblastoma overexpressed (NOV) were discovered in early 1990’s which led to the acronym CCN based on their order of discovery. CCN is also abbreviated for “Cellular Communication Network” [2]. The other three members were discovered in the late 1990’s and were associated with Wnt-1 induced signaling pathway and hence were named, (iv) Wnt-1 induced secreted protein-1 (WISP1), (v) WISP2 and (vi) WISP3. Over the past two decades, extensive research on CCN family members uncovered their diverse cellular and biological functions, designating them with alternative names listed in Table 1. The International CCN Society proposed a unifying nomenclature for the CCN family members to avoid confusion due to numerous synonymic names interchangeably used in scientific literature [3]. In 2018, the HUGO (Human Genome Organization) Gene Nomenclature Committee officially approved and adapted the new names CCN1-6 for the family members [2, 3]. For this review, these proteins are named CCN1-CCN6, in accordance with the official nomenclature recommendations.

Table 1 Alternative names of CCN family of proteins

Based on PubMed(R) database, there are over 3000 publications on CCN protein family, out of which surprisingly more than 2900 studies primarily focus on CCN2 and only about 300 publications focus on CCN4, highlighting the dearth of literature and disparity towards other CCN members. Despite the overwhelming attention given to CCN protein family, relatively little is known about the role of CCN4 both in human health and diseases. Notably, the role of CCN4 has been previously studied and reviewed by researchers in a disease specific context [4, 5]. However, here, we aim to comprehensively review the growing body of literature on the diverse functions of CCN4 and its role in a vast array of pathophysiological conditions, including cancer, fibrosis, inflammatory conditions (i.e., arthritis) and metabolic disorders, including obesity and diabetes.

CCN4 overview

CCN4 is the fourth member of the CCN family, commonly known as WISP1. The earliest study published on CCN4 dates back to the late 1990s where Pennica and colleagues first identified CCN4 in mouse mammary epithelial cells and its role in tumorigenesis [6]. WISP1 gene on chromosome 8q24.22 in humans encodes the 367 amino acid CCN4 protein with a predicted molecular mass of approximately 40 kDa. The murine and human CCN4 cDNA length is 1766 and 2830 bases, respectively, comprising of four introns and five exons [6, 7]. Structurally, all CCN family members are characterized by four conserved cysteine rich domains, except CCN5 which lacks domain 4. The N-terminus of CCN proteins consists of a signal peptide sequence, essential for secretory proteins, which is followed by four structural domains named based on the sequence homology to (i) insulin-like growth factor binding protein like domain (IGFBP; domain 1), (ii) the von Willebrand factor C repeat (VWC; domain 2), (iii) thrombospondin-homology type 1 repeat (TSP1; domain 3) and (iv) the C-terminal Cysteine knot containing domain (CT; domain 4) as shown in Fig. 1. The protein sequence contains 38 conserved cysteine residues, distributed across domain 1 (12 cysteine residues), domain 2 (10 cysteine residues), domain 3 (6 cysteine residues), and domain 4 (10 cysteine residues). CCN4 shares approximately 39%, 37% and 42% amino acid sequence homology to CCN2, CCN5 and CCN6, respectively [8, 9].

Fig. 1
figure 1

Multi-modular structure of CCN family. All the CCN matricellular proteins consist of a signal peptide (SP), insulin-like growth factor binding protein like domain (IGFBP; domain 1), von Willebrand factor C repeat (VWC; domain 2), thrombospondin-homology type 1 repeat (TSP1; domain 3) and C-terminal Cysteine knot containing domain (CT; domain 4) encoded by exon 1 to exon 5 respectively. Domain 2 and domain 4 are connected by a variable hinge region which is highly susceptible to cleavage by proteolytic enzymes such as, Kallikrein-related peptidases 12 (KLK12), a disintegrin and metalloproteinase domain-containing protein 28 (ADAM28) and matrix metalloproteases (MMPs). Domain specific total cysteine residues and binding partners are also listed in the table. The image was created with BioRender.com. IGFBP: insulin-like growth factor binding protein; VWC: von Willebrand factor C repeat; TSP1: thrombospondin-homology type 1 repeat; CT: C-terminal Cysteine knot containing domain; IGF: insulin-like growth factors; BMP4: bone morphogenic protein 4; LRP1: Low density lipoprotein receptor-related protein 1; HSPGs: heparin sulphate proteoglycans; TGF-β: transforming growth factor beta

CCN family proteins have been shown to physically bind and interact with a plethora of multi-ligand receptors and proteins, ranging from ECM proteins (such as fibronectin [10, 11], vitronectin [12], perlecan [13], integrins [14,15,16], growth factors (such as fibroblast growth factors (FGFs) [17,18,19], vascular endothelial growth factor (VEGF) [20], bone morphogenetic protein (BMP) [21] and transforming growth factor (TGF-β) [21], proteoglycans (such as heparan sulfate proteoglycans (HSPG) [22,23,24], aggrecan, decorin and biglycan [25] and low-density lipoprotein receptor-related proteins (LRP) [26, 27], cation-independent mannose-6-phosphate [28], Neurogenic locus notch homolog protein (Notch) [29], and receptor activator of nuclear factor kappa B (NF-κB) (RANK) [9, 30, 31]. Despite abundance of binding partners available for CCN family of proteins, CCN4 has only been shown to interact with integrins [32, 33] and small leucine rich proteoglycans, such as decorin and biglycan [25], and domain specific binding sites that facilitate these interactions remain unknown, owing to the bias towards other CCN members, particularly CCN1 and CCN2.

Post-translational modifications (PTMs) play a crucial role in structural and functional characteristics of a protein. With respect to the CCN family, four potential N-glycosylation sites have been identified in CCN4 [6]. Similar glycosylation patterns have also been observed in CCN2, particularly at asparagine 28 and 225 leading to either 36 kDa or 38 kDa molecular weight that appears as a double band upon immunoblotting [34, 35]. Furthermore, researchers have also demonstrated that the glycosylated versus non-glycosylated ratio of CCN3 profoundly influences the cell proliferative, migrative and invasive properties of chondrosarcoma cell line Jeg3 [36]. Differences in the glycosylation patterns between normal and cancerous cells have also been reported, highlighting its functional significance in disease pathologies [37]. Besides glycosylation, O-fucosylation (O-linked fucose modification) has been recently identified in the TSP domain 3 of CCN2 protein [38]. Overall, these PTMs could be exploited to modulate CCN4 function; yet the PTM analysis in CCN family remains in its infancy, and more research is needed to shed more light on its clinical significance.

As previously mentioned, CCN genes contain four introns and five exons. Each of the exon codes for each domain in the structure. The N-terminus signal peptide is encoded by exon 1, whereas domain 1–4 are encoded by exon 2–5, respectively. From the evolutionary perspective, the order of CCN domains remains highly conserved and are connected through a linker. The linker connecting domain 1–2 (3aa) and domain 3–4 (9aa) in CCN4 structure are relatively short compared to domain 2–3 hinge (27aa) which makes it susceptible to proteolytic cleavage by protease enzymes, such as matrix metalloproteases (MMPs) generating truncated versions of the protein. The central variable linker has been shown to be targeted by MMP1, MMP3, MMP 7, MMP9, MMP13, MMP14 along with other peptidases, such as disintegrin and metalloproteinase domain-containing protein 28 (ADAM28), Kallikrein-related peptidases 12 (KLK12), plasmin and elastase [39,40,41,42]. Experimentally, the protease mediated degradome pattern has been extensively elucidated for CCN1, CCN2, CCN3 and CCN5 [39, 43,44,45]. While the cleavage sites for CCN4 remain undiscovered, researchers have identified three truncated versions in biological samples as shown in Fig. 2. In addition, the CCN4 mRNA is also subjected to alternative splicing, which can result in truncated variants of proteins, lacking one or more domains [46]. In 2001, Tanaka and colleagues reported a novel truncated CCN4 variant (Mol. Wt. 30 kDa), commonly known as WISP1v lacking domain 2 or VWR module as a product of alternative splicing in human scirrhous gastric carcinoma tissue [47]. Stable transfection of this variant increased gastric carcinoma cell migration up to five-fold as compared to full length CCN4 in a co-culture Boyden Chamber. Later in 2003, they also detected CCN4 truncated variant, WISP1v in human invasive cholangiocarcinoma tissues [48]. Further, in 2008, Inksonand colleagues at the National Institutes of Health (NIH) detected a similar CCN4 spliced variant lacking exon 3 in human bone marrow stromal cells (hBMSC) using RT-PCR, which they referred to as WISP1va [49]. In 2004, another truncated variant of CCN4, lacking domain 2, 3 and 4, also known as WISP1Δex3-4 was reported in 4 different human hepatocellular carcinoma cell lines (HuH-6, HuH-7, VGH and HepG2) in conjugation with full length and WISP1v. WISP1Δex3-4 variant only contains IGFBP/ domain 1 due to the alternative splicing of mRNA. The transcript contains exon 1, 2 and 5, however the conjugation of exon 2 and 5 causes a frame shift at residue 117, resulting in a premature stop of the reading frame and hence only translating to domain 1 [50]. Similarly, WISP1Δex3-4 truncated variant was also detected in chondrocytes derived from human chondrosarcoma cell line (HCS-2/8) and rabbit growth cartilage [51]. Although the functional significance of full length CCN4 has been extensively demonstrated, the domain specific activity remains unknown till date despite the detection of CCN4 truncated versions in human tissues.

Fig. 2
figure 2

Functional effects of CCN4 Truncated variants. The image was created with BioRender.com

Meanwhile, emerging evidence in the literature suggests that CCN4 can in turn modulate MMP expression. CCN4 treatment for 24 h in human chondrosarcoma cell line, JJ012 increased MMP2 expression in cell lysate and supernatant, detected at both protein and transcript level utilizing western blot and qPCR [33]. Further, CCN4 has also been shown to promote cell motility by upregulating MMP2 and MMP9 expression in human osteosarcoma cell line, U2OS as pretreatment with selective MMP2 and MMP9 inhibitors abrogated CCN4 induced wound healing and migration [32]. In addition, stimulation of murine macrophages (RAW 264.7), primary human chondrocytes and synovial cells with CCN4 (1.0 µg/ml) for 24 h upregulates MMP3 and MMP9 expression [52]. Also, CCN4 drives MMP2 and MMP9 expression in murine primary renal tubular epithelial cells as shRNA mediated CCN4 knockdown decreased MMP2 and MMP9 [53]. Finally, CCN4 has also been shown to upregulate the expression of MMP1, MMP2, MMP3, MMP9 and MMP14 in vein smooth muscle cells via β-catenin mediated pathways. CCN4 mediated MMP9 induction is partly due to activator protein-1 (AP-1) [54].

The control exercised by CCN4 on the MMPs can drive cell motility via two distinct, though parallel mechanisms. The observations derived from literature unveil that CCNs can indirectly influence cell motility by upregulating MMPs, which drives ECM degradation, facilitating cell migration and invasion [55]. Another direct mechanism by which CCN4 can influence cellular processes could be through MMP-dependent self-regulation, where CCN4-induced MMP can in turn act on CCN4 itself to generate different truncated versions of the protein. Depending on the tissue expression profile of MMP subtypes and their corresponding CCN4 cleavage sites, these truncated variants can possess similar and/or unique functional signature as compared to the full length CCN4 protein. Further, alterations in CCN4 expression profile have been observed in a plethora of diseases, ranging from cancer [56], liver fibrosis [57], idiopathic pulmonary fibrosis (IPF) [58,59,60], obesity and type 2 diabetes mellitus [61, 62], amongst others. The CCN-MMP interplay can continue to regulate one another, leading to a vicious positive-feedback cycle which can potentially aggravate the underlying condition by generating truncated CCN4 which could be responsible for diverse cellular functions. Taken together, CCN4 mediated regulation of MMPs is a highly complex, expression dependent, cell type and tissue-specific mechanism which yet remains underexamined and warrants further investigation.

In humans, CCN4 expression has been confirmed in various organs, such as lung, heart, kidney, pancreas, placenta, brain, small intestine, ovaries, and skeletal muscle, amongst others [6]. Within the organs, the expression is cell type specific. For example, CCN4 is mainly expressed in fibroblasts (lung [63], liver [57], heart [64]), epithelial cells (lung) [60], cardiomyocytes (heart) [65], neurons [66], microglia (brains) [67], chondrocytes [33], osteoblast (bone) [32] and many more. The diverse expression profile endows CCN4 protein with pleotropic functions in a tissue and cell specific manner. CCN4 plays a crucial role in cell proliferation, migration, adhesion, wound healing and repair in embryogenesis, fibrosis, tumorigenesis, osteoarthritis, etc. [68]. In addition, differential expression of CCN4 has been attributed to numerous diseases, which can be utilized as a prognostic biomarker.

CCN4 in cancer

Cancer represents a global health issue with increasing cases every year [69]. According to the national cancer society, cancer is the second leading cause of death, with approximately 1 in 3 people suffering in the US. Recent technological advances in cellular and molecular biology have opened endless avenues for the development of targeted anti-cancer treatment, ranging from stem-cell therapy, gene therapy to targeted precision therapy [70]. Differential gene expression and novel biomarker discovery efforts help identify promising druggable targets with potential therapeutic benefits in the clinic. CCN4 is one such recently identified protein which plays a crucial role in inflammation and tumorigenesis [71].

Over two decades, numerous studies have investigated the role of CCN4 in tumor microenvironment, however the ambiguous expression profile and paradoxical functional outcomes of CCN4 in various cancers makes it a challenging yet controversial target. Generally, CCN4 dysregulation and atypical expression profile have been linked in a range of pathophysiological conditions, such as fibrosis, diabetes, obesity etc. [57, 72]. Similar aberrant expression profile has been observed in various cancer types as well. For instance, CCN4 is upregulated in a vast array of cancer tissue specimens compared to healthy controls, such as lung cancer [73], ovarian cancer [74], colon cancer [75], gastric cancer [76], breast cancer [77] and esophageal squamous cell carcinoma [78], among others. CCN4 is also characterized as an oncogene, that promotes tumor progression by positively regulating pro-oncogenic cellular functions, like cell proliferation, migration, and invasion. In addition, higher CCN4 protein levels were associated with low survival rate in cancer patients [79,80,81]. In contrast, others have reported tumor suppressive nature of CCN4, promoting cellular apoptosis whilst inhibiting cell growth, migration, invasion, and metastasis. Many researchers have also reported CCN4 downregulation in breast cancer [82, 83], liver cancer [84] and skin cancer [85]. The discrepancies could also be attributed to other confounding variables, such as patient family-history, age, gender, co-morbidities, treatment regimen, cancer-type, cancer-stage, tumor size etc., however in-depth analysis is required before establishing any correlation. Rather than simply distinguishing all diseased patients as either high or low expressers, some studies also segregate the diseased cohort into two groups, CCN4 low and high expressing patients based on the expression profile [79,80,81]. The spatial and temporal tumor heterogeneity leads to a complex dynamic cellular state which can also govern the distinct CCN4 expression patterns. A summary of the functional and mechanistic effects of CCN4 on diverse tumor types, along with the expression patterns is provided in Table 2.

Table 2 Expression and functional effects of CCN4 in cancers

Tumor microenvironment is composed of cancer cells, stromal cells, such as immune cells, fibroblasts, endothelial cells in conjunction with ECM, all of which are constantly communicating and influencing each other via cytokines, chemokines and inflammatory signaling molecules [86]. Although numerous cell-based and animal experimental models have been utilized to study the effect of CCN4 in cancer, majority of the studies are conducted in an isolated single cancer cell line model, which fails to capture the intricate crosstalk amongst different cell types within the tumor microenvironment. Given that CCN4 is a secreted protein, better understanding of both autocrine and paracrine effects is crucial to decipher the functional disparity and clinical significance of CCN4 to develop novel pharmacological interventions targeting CCN4 in cancer [87]. Furthermore, mechanistic data is derived from either loss-of function approach, including shRNA, siRNA or CRISPR mediated transcriptional repression or protein overexpression approach using viral vector. Every experimental approach has its own caveats, here both approaches may fail to capture the true function of endogenous CCN4 due to the off-target effects and lack of specificity leading to partial loss-of function or by attaining supraphysiological protein levels [88, 89]. Some studies also utilize recombinant CCN4 protein to understand its biology, which has its own drawbacks. Besides the time-consuming and costly process, the multi-modular structure of CCN4 is susceptible to proteolytic cleavage generating functionally active or in-active truncated variants. Additionally, species to species variations in post-translational modification patterns could also influence the biological activity of the protein [90]. Hence, caution must be exercised while interpreting the functional consequences of full-length CCN4 protein.

Molecular pathways governing CCN4 mediated cellular oncogenic responses remains undiscovered, due to the lack of one CCN-specific cognate receptor. However, within the limited available literature, integrins have been identified as the major contributors in driving CCN4-dependent biological responses. Researchers have shown that CCN4 binds and interacts with integrins, such as, αVβ1, αVβ3, αVβ5, α4β1, α5β1 and α6β1 in myriad cancerous conditions to drive cell proliferation, migration, invasion etc. making them one of the most extensively documented functionally significant receptors for CCN4 till date [32, 33, 37, 74, 87, 91,92,93,94,95]. In addition, CCN4 polymorphism directly influences the risk of developing tumor [96, 97], disease progression [98], response to chemotherapy toxicity [99] and CCN4 expression via epigenetic modulation such as DNA methylation [100].

CCN4 in fibrosis

Fibroblasts are highly plastic cells with mesenchymal origin found throughout the body to provide structural integrity and basic framework for cells and tissues. Fibroblasts are elongated stellate shaped cells that are most commonly present in the stroma. They are responsible for maintaining, synthesizing, and organizing ECM proteins, such as collagen, fibronectin (FN1), laminins etc. and therefore, play a key role in wound healing and tissue repair [128]. Mechanical or chemical stimuli from the site of injury can activate the otherwise quiescent tissue resident fibroblasts and initiate their transformation into myofibroblasts. Under physiological homeostasis, once the wound healing and repair has been completed, myofibroblasts undergo apoptosis to prevent excessive ECM deposition. However, chronic persistent injury and insult, can dysregulate and disrupt the body’s natural restorative process, leading to excessive ECM deposition, tissue scarring and architectural remodeling, loss of tissue elasticity and function, resulting in fibrosis [129]. Fibrosis is the most common pathological outcome in chronic inflammatory conditions related to lungs (Idiopathic pulmonary fibrosis; IPF), skin (scleroderma), kidney diseases, liver, and heart (cardiac fibrosis) [130].

Fibrosis is a chronic, highly progressive, and irreversible condition that is the leading cause of organ dysfunction and death. Although, two FDA (Food and Drug Administration) approved drugs, Nintedanib and Pirfenidone provide symptomatic relief in IPF, there are currently no drugs that address the underlying cause to cure fibrosis [131]. Emerging evidence indicates that CCN4 protein is pro-fibrotic in nature and modulates fibroblast proliferation. CCN4 is highly upregulated in both pre-clinical bleomycin model of pulmonary fibrosis, paraquat induced model [132, 133] and in clinical IPF patients compared to non-diseased control. Furthermore, Klee and colleagues demonstrated that CCN4 is downstream of TGF-β and TNF-α, which are the master regulators of fibrosis and inflammation. CCN4 promotes the proliferation of human lung fibroblast in an IL-6 dependent manner as siRNA mediated knockdown or antibody-mediated neutralization of CCN4 abrogates the effect [63]. The CCN4 upregulation could be partly due to the downstream effects of TGF-β in conjunction with the downregulation of microRNAs (miRNAs), particularly miR-92a, which has been shown to modulate CCN4 expression. miR-92a expression is inversely corelated with CCN4 expression in IPF patient lung specimens [134]. In addition, miR-101 and miR-181a-5p regulate CCN4 expression in cystic fibrosis [135]. Airway epithelial cells not only act as the first line of defense against environmental threats but also serve as a dynamic junction to relay the extracellular signal to other immune cells that underlay smooth muscle cells, fibroblasts and myofibroblasts. Chronic epithelial insult and dysfunction have been attributed to the pathogenesis of asthma and IPF [136,137,138]. Heise and colleagues demonstrated that mechanical stress and stretch can induce CCN4 expression in primary mouse type II alveolar epithelial cells (AT-II cells) and drive epithelial to mesenchymal transition (EMT). Further, CCN4 neutralizing antibody significantly abrogated stretch induced EMT, emphasizing the critical role of CCN4 in EMT [139]. Similar findings were validated by another group, where treatment with recombinant CCN4 (1 µg/ml) promoted cell proliferation and EMT in primary mouse AT-II cells. In addition, stimulation with CCN4 (1 µg/ml) for 6 to 12 h upregulates fibrotic genes, such as Col1A1, Col1A2 and FN1 in mouse and human fibroblasts and the effects were attenuated in the presence of a CCN4 neutralizing antibody in the bleomycin model [60]. CCN4 mediated cell adhesion in airway epithelial cells (A549) is partly mediated by integrins, as αVβ5, αVβ3 or αVβ1 neutralizing antibodies partially blocked the effect [140]. Irradiation has been shown to upregulate CCN4 expression in human lung fibroblasts with implications in radiation-induced lung injury in cancer patients [141]. Nintedanib, a small molecule receptor tyrosine kinase inhibitor, approved for IPF has also been shown to regulate Wnt/β-catenin pathway and prevent myofibroblast activation by inhibiting CCN4 in mouse lung myofibroblast cell line, Mlg [142]. Furthermore, secreted CCN4 was significantly decreased upon treatment with Nintedanib (1 µM) in ex-vivo 3D-human lung tissue. While CCN4 levels remained unaffected upon treatment with Pirfenidone (500 µM) detected by ELISA (Enzyme-linked immunosorbent assay) [143]. However, in another study, both Nintedanib (0.3 µM) and Pirfenidone (1 mM) reduced in precision cut rat-lung slices [144]. CCN4 can also facilitate inflammatory response in fibrosis. In addition to IL6 and CCL2 production [63, 145]. CCN4 also mediates the release of the proinflammatory cytokine TNF-α from macrophages (RAW264.7) in an integrin αVβ3 dependent manner and regulates TLR4 signaling in an acute lung injury pre-clinical model [146].

Apart from pulmonary fibrosis, CCN4 has also been involved in liver fibrosis and inhibition of CCN4 can reverse liver fibrosis [147,148,149]. Stimulation with pro-fibrotic/ pleotropic cytokines, such as TGF-β and TNF-α increased CCN4 induction in-vitro in hepatic stellate cell lines (LX-2 and HSC-T6/ HSC). In addition, recombinant CCN4 drives LX-2 cell proliferation in a dose-dependent manner. CCN4 protein expression was also significantly upregulated in-vivo in carbon tetrachloride (CCl4)-induced liver fibrosis model [148] and CCN4 antibody significantly decreased pro-fibrotic protein expression (collagen, α-smooth muscle actin (αSMA), TGF-β1), reduced liver necrosis, NF-κB activation and pro-inflammatory cytokine production, such as IL-6, CCL-2 and TNF-α [149]. Huang and colleagues utilized RNA sequencing analysis to identify a set of differentially expressed genes in ex-vivo precision-cut lung tissue slice to design a robust biomarker panel to assess antifibrotic effects of various interventions. CCN4 was amongst other genes and secretory proteins in the panel and was used as a reliable end point parameter to evaluate the efficacy and anti-fibrotic activity of the compounds [144]. Interestingly, as the molecular mechanisms driving the initiation and progression of fibrosis remain poorly understood, recent study identified a novel pathway dissecting the role of CCN4 in the progression of liver fibrosis and not initiation of the disease as CCN4 knockout animals were protected against liver-fibrosis progression in pre-clinical CCl4-liver fibrosis and choline-deficient, L-amino-acid-defined, along with the high-fat diet (CDA-HFD)-induced NASH models. Furthermore, functional analysis confirmed that CCN4 mediated fibrogenesis and myofibroblast motility is partly driven by integrin (αV, α11) dependent myocardin-related transcription factor (MRTF) activation, that drives MRTF-downstream cytoskeletal gene targets, such as αSMA, myosin light chain 9, filamin A, etc., in primary HSCs. Although the precise mechanism of fibrosis remains elusive, evidence from current literature indicate that CCN4 could serve as a potential therapeutic target for the treatment of liver fibrosis and small or large molecule therapeutic modalities that inhibit CCN4 can elicit protective effects in liver injury and fibrosis.

CCN4 has a crucial role in skin biology, wound healing, and repair. CCN4 protein expression was upregulated 4–7 days post cutaneous wounding and facilitated wound healing as the extent of wound closure was significantly delayed in CCN4-knockout mice due to the downregulation of ECM proteins, such as Col1A1 and FN1 [150]. Immunohistochemistry analysis of the incision reveals that CCN4 is also abundantly expressed in inflammatory cells, such as neutrophils. CCN4 is crucial for wound healing as it drives proliferation and migration of both human and mouse dermal fibroblasts through integrin α5β1 as selective siRNA mediated CCN4-knockdown resulted in the loss of function. Furthermore, stimulation with 100 ng/ml CCN4 induced activation and phosphorylation of ERK and c-Jun N-terminal kinase (JNK), crucial for cell proliferation, an effect which was blocked in the presence of selective small molecule MAPK inhibitor, PD98059 and αVβ1-antibody [150]. CCN4 not only binds with integrins but can also interact with cell surface small-leucine rich proteoglycans such as, decorin and biglycan on human dermal fibroblasts [25], although the downstream mechanistic pathway engagement from the latter remains unknown.

Besides the role of CCN4 in lung, liver, and skin fibrosis, CCN4 has also been shown to be implicated in cardiac remodeling and fibrosis associated with cardiomyopathies. Similar to what others have shown in diverse pathological conditions, CCN4 is substantially upregulated post-myocardial infarction (MI) and ischemic injury [64, 65, 151, 152]. CCN4 modulates cardiac remodeling by positively influencing cardiomyocyte hypertrophy in an Akt-dependent manner and stimulation with recombinant CCN4 induces cardiac fibroblast proliferation and enhances ECM protein deposition, particularly collagen [65]. As previously mentioned, pro-inflammatory cytokines and chemokines are closely intertwined with CCN4 biology, and both have been shown to positively regulate each other. Along the same lines, stimulation with either TNF-α and/or IL-1β significantly induced CCN4 protein expression both in left-ventricular myocardium post-MI in-vivo and in rat cardiac myocytes in-vitro [65]. These findings were further strengthened in another study, where TNF- α induced CCN4 upregulation was shown to be dependent on ERK1/2 mediated CREB phosphorylation at Ser133 as pretreatment with small molecule inhibitors such as PD98059 (ERK1/2) and U0126 (MEK) failed to induce CREB-phosphorylation [64]. As TNF-α is known to activate a vast array of downstream signaling molecules, the authors also ruled out the involvement of JNK- and NF-κB activation for CCN4 upregulation and concluded that TNF-α mediated responses were strictly dependent on MEK1-ERK1/2-CREB signaling in cardiac fibroblasts [64]. In addition to CCN4, biglycan, the potential binding partner of CCN4 is also upregulated in cardiac fibroblasts up to threefold post-MI in-vivo suggesting intracellular CCN4 signal amplification [65].

Renal fibrosis is one of the most common pathological hallmarks in chronic kidney diseases (CKD). A recent study found that CCN4 levels were highly upregulated in preclinical unilateral ureteral obstruction (UUO) renal fibrosis model in animals and clinically in serum and kidney tissue biopsy samples from CKD patients [153,154,155]. Mechanistically, antibody mediated neutralization or siRNA mediation knockdown of CCN4 provides protection against renal fibrosis by attenuating fibrotic markers, such as Collagen, FN1 and αSMA deposition both in tubular epithelial cells (NRK52E cell line) and in-vivo in mouse models. Interestingly, the role of autophagy in fibrosis remains controversial as there are opposing results on whether it promotes or inhibits fibrogenesis. However, previous investigations have reported enhanced autophagic markers in proximal tubular cells, pharmacological inhibition of which reversed renal fibrosis. Similarly, CCN4 inhibition significantly reduced autophagy in UUO renal fibrosis model, suggesting that CCN4 also exercise control over pathways governing programmed cell death, modulating the development of renal fibrosis [156]. Serum CCN4 levels were also found to be elevated in vast array of CKDs (chronic kidney disease), such as diabetic nephropathy, IgA nephropathy and primary focal segmental glomerular sclerosis [153, 157]. CCN4 has also been implicated to drive migration, invasion and EMT in primary renal tubular epithelial cells in uremia associated with end-stage renal failure [53]. In addition, the growing body of literature on non-coding RNAs, particularly miRNA and circular RNA (circRNA) and its prominent role in the pathophysiology of a wide array of disease highlights them as potent gene regulators. Regarding CCN4 gene regulators, miR-92a, miR-101 and miR-181a-5p have been identified to inversely modulate CCN4 expression in lungs. A recent study discovered two novel non-coding RNAs that target CCN4 to modulate renal fibrosis. The results demonstrate that circRNA-33702 is overexpressed in UUU-renal fibrosis models and possess profibrotic role by aggravating collagen and FN1 expression [158]. Conversely, miR-29b-3p negatively modulates CCN4 expression, in conjunction with other ECM proteins in mouse proximal tubule cell line (BUMPT cells). Since circRNA-33702 and miR-29b-3p have opposing effects on CCN4 expression and colocalization, luciferase analysis revealed that circRNA-33702 directly binds miR-29b-3p to upregulate CCN4 expression and consequently promote renal fibrosis [158]. The anti-fibrotic effects of miR-29b-3p on cardiac and liver fibrosis have also been demonstrated by others [159,160,161,162]. Apart from miR-29b, long non-coding RNA, Gm12840 and miR-677-5p also target CCN4/Akt signaling pathway to modulate fibroblast activation in ischemia–reperfusion induced renal fibrosis [163]. Overall, non-coding RNAs possess exciting potential as novel therapeutic targets for the treatment of fibrosis, however more investigation in this area is required to comprehensively understand the crosstalk amongst non-coding RNAs in a tissue-specific context to target CCN4.

In addition, all the CCN family members are functionally interconnected with a high degree of crosstalk by compensatory or opposing mechanisms. Emerging evidence points towards the anti-fibrotic effects of CCN3 protein and one of the possible anti-fibrotic mechanisms involves downregulation of profibrotic CCN4 protein. Overexpression of CCN3 in the skin fibroblast cell line NIH3T3 significantly downregulated CCN4 expression and hence conferred protection against fibrosis, although the exact mechanisms by which CCN3 modulates CCN4 remains unknown. It is speculated that due to its presence in the nucleus, CCN3 may behave as a transcription factor and can directly inhibit CCN4 gene transcription. Another mechanistic explanation could be due to direct sequestration of CCN4 by protein–protein interaction, preventing the initiation of CCN4 mediated pro-fibrotic pathway [164]. Similar findings were also reported by others where CCN3 was identified as an endogenous inhibitor of pro-fibrotic CCN family members. A deeper understanding of the interactome of CCN family members is required for utilizing the antagonistic approach to develop anti-fibrotic therapeutics [165, 166]. Taken together, these results show that CCN4 is highly upregulated in fibrotic tissues, such as lungs, heart, liver and skin and can directly influence fibrogenesis, partly by integrin dependent mechanisms to promote fibroblast proliferation and migration, suggesting that CCN4 can not only serve as a diagnostic biomarker but also be exploited as a novel therapeutic target for the treatment of fibrosis. In addition, targeting CCN4 may eliminate the need for multiple tissue specific therapies in multi-organ fibrosis, given its role in the fibrogenesis of major organs, such as lung, liver, heart, kidney, and skin which encompasses a majority of all the fibrosis cases.

CCN4 in obesity and diabetes

Obesity is defined as excessive accumulation of fat in the adipose tissue throughout the body due to imbalance in the energy intake and expenditure, leading to various cardiovascular and metabolic disorders. Numerous environmental, genetic and lifestyle related factors also contribute to the increased body mass index (BMI) and weight gain [167]. Previously considered to be inactive, adipose tissue is a highly dynamic metabolically active endocrine organ, that produces a wide variety of cell-signaling molecules, namely adipocytokines or adipokines such as leptin, adiponectin, resistin, TNF-α, and IL-6, amongst others [168]. These adipokines are crucial for biochemical and metabolic homeostasis, however increased adiposity mediated adipokine dysregulation is the major culprit involved in the pathogenesis of metabolic syndrome, such as insulin resistance, diabetes mellitus, atherosclerosis, etc. [169]. CCN4 has recently been identified as a novel adipokine in humans, adding a unique functional aspect to its diverse biological repertoire [170]. Interestingly, recent studies have shown that CCN4 is also expressed and secreted by human adipocytes endowing it with the title of novel adipokine. Amongst other CCN protein members, CCN3 is also a fairly recently discovered adipokine linked to obesity [171]. Discovery of CCN4 and CCN3 as novel metabolic regulators opens new avenues for the treatment and management of obesity and associated co-morbidities.

Over the last decade, emerging evidence directly correlates systemic CCN4 levels with obesity, inflammation, and insulin-resistance. Large human cohort studies with obese and/or glucose tolerant patients revealed that circulating CCN4 is positively correlated with percent fat mass, leptin, triglyceride levels, adiposity, and BMI [172, 173]. Another study shows that serum CCN4 levels and CCN4 mRNA expression in visceral adipose tissue were significantly higher in obese men compared to non-obese men, independent of their glycemic status [174]. CCN4 also leads to insulin resistance by impairing insulin signaling in hepatocytes and primary human skeletal muscle cells. In a dose-dependent manner, CCN4 significantly abrogated insulin-mediated phosphorylation of insulin receptor (IRβ)-Tyr1150/1151, along with decreased Akt-Ser473/Thr308, GSK3β-Ser9 phosphorylation at the lowest dose of 0.1 µg/l in both human skeletal muscle cells and murine hepatocyte cell line AML12. Insulin receptor substrate 1 (IRS1) is a key cytoplasmic adaptor protein crucial for signal transmission downstream of the receptor and treatment with 0.1 µg/l and 1 µg/l CCN4 decreased IRS1 protein expression by 50%, suggesting the direct inhibitory effect of CCN4 on insulin cascade in human skeletal muscle cells [174]. Preincubation with 0.1 µg/l and 1 µg/l of CCN4 for 24 h significantly abrogated insulin-dependent glycogen synthesis in human primary myotubes [174]. To further validate the mechanism, Woo et al. demonstrated the effect of CCN4 knockdown on insulin resistance and glucose in skeletal muscle cells of HFD-mice [175]. CCN4 knockdown significantly abrogated the inhibitory effects on insulin signaling by restoring Akt and IRS1 phosphorylation. Mechanistically, the authors also showed that CCN4 mediated insulin resistance and inflammation in murine skeletal muscle cells (C2C12 cells) and hepatocytes respectively is driven via Toll-like receptor-4 (TLR4) as TLR4 knockdown significantly abrogated CCN4-mediated JNK phosphorylation, NF-κB translocation, insulin resistance and triglyceride accumulation in hepatocytes and C2C12 cells [175]. siRNA mediated knockdown of NF-κB and JNK prevents CCN4-mediated insulin resistance, highlighting a novel mechanism for CCN4-driven impaired insulin sensitivity. In addition, siRNA-mediated CCN4 knockdown significantly ameliorates hepatic steatosis, lipogenesis and insulin resistance in HFD-fed mice suggesting that CCN4 requires TLR4 activation to drive inflammation and insulin resistance [175].

Contrary to the previously described inhibitory effect of CCN4 on insulin signaling, another study reports that it drives insulin-producing pancreatic beta (β)-cell proliferation via Akt modulation in both mouse and human cells [176]. Cell proliferation markers such as antigen kiel 67 (Ki67) and phospho-histone H3 (pHH3) were significantly reduced in CCN4 knockout mice (CCN−/−) as compared to wild-type (CCN+/+). Further, CCN−/− mice treated with recombinant CCN4 exhibited twofold higher β-cell proliferation as compared to saline. Adenovirus-mediated systemic overexpression of CCN4 in streptozotocin-induced diabetes model, significantly increased plasma insulin levels by augmenting total β-cell mass and insulin positive area, however failed to reverse hyperglycemia [176]. CCN4 has also been implicated in the development and regeneration of pancreas [177, 178]. More recently, researchers have also identified that CCN4 is highly expressed upon treatment with high concentration of glucose (30 mM) in human kidney proximal tubular cells and in renal tissue of streptozotocin-induced diabetic nephropathy (DN) mouse model [179]. Functionally, CCN4 overexpression drives cell proliferation, migration, EMT and fibrosis and these effects were partially rescued via silencing N6-adenosine methyltransferase (METTL3), which decreases DN development by decreasing CCN4 expression in-vitro [179]. Given the current body of conflicting literature on the effect of CCN4 on glucose homeostasis and rising rate of metabolic disorders, it is extremely important to decipher the functional and mechanistic consequences of CCN4 due to its profound physiological relevance.

Two independent studies have also shown that CCN4 serum level in pregnant women with gestational diabetes mellitus (GDM) is significantly higher as compared to healthy non-GDM pregnant women [180, 181]. In addition, circulating CCN4 level in obese pregnant women with GDM is positively correlated with numerous clinical metabolic parameters such as systolic blood pressure, fasting blood glucose and aspartate aminotransferase (AST), highlighting the crucial role of CCN4 in the pathophysiology of GDM [180]. Overall, CCN4 can serve as a strong independent risk predictor and diagnostic marker and possess immense therapeutic potential in maternal-neonatal health and obstetric research. However, maternal, and neonatal safety drug assessment becomes crucial to assess the impact of possible CCN4 interventions on the fetal growth and development as CCN4 is expressed in osteoblasts and their progenitor cells during skeletogenic processes in embryonic development [182].

Obesity is also characterized as a chronic low-grade systemic inflammation due to the proinflammatory cytokine release from adipocytes and macrophages [183]. Studies show that macrophage infiltration and accumulation in adipose tissue was significantly higher in obese HFD mice as compared to normal mice [184]. Furthermore, phenotypical polarization of the infiltrated macrophages was observed in obese individuals with predominant pro-inflammatory M1 macrophages as compared to lean individuals with more anti-inflammatory M2 macrophages. The increased M1 population in obese adipose tissue overexpresses pro-inflammatory genes such as IL-6 and TNFα and lower anti-inflammatory cytokines such as IL-10, contributing to the persistent low-grade systemic inflammation and insulin resistance [184, 185]. In addition, CCN4 protein expression is upregulated in visceral and subcutaneous adipose tissue in glucose-tolerant patients and is positively correlated with the markers of obesity, inflammation, and insulin resistance [170, 172, 174]. Serum CCN4 levels were also significantly elevated in obese children and adolescents with direct positive correlation to IL-18, adiponectin, and leptin [61]. Interestingly, these effects were completely reversed upon weight loss as adipose tissue CCN4 expression was significantly decreased after weight reduction, suggesting that adipocytes are the major source of circulating CCN4 [170]. These findings were further validated in another single-center randomized trial with breast cancer survivor females, where a 12-week exercise regime significantly decreased waist circumference and body fat composition accompanied by reduced serum β-catenin and CCN4 levels [186]. Furthermore, studies have shown that stimulation of macrophages with CCN4 can significantly increase pro-inflammatory cytokines such as IL-6, TNF-α and IL1B at both mRNA and protein level [170, 187]. However, CCN4 stimulation had no significant pro-inflammatory effect on adipocytes suggesting that adipocyte-derived CCN4 does not elicit autocrine-response but rather have paracrine inflammatory effects on nearby macrophages. Additionally, stimulation of macrophages with CCN4 significantly increased the expression of pro-inflammatory M1 specific markers, such as CCR7 and COX2, whereas the expression of anti-inflammatory M2 specific markers such as CD36, CD163, MRC1 and COX1 were either markedly decreased or remained unchanged. This suggests that ‘M2 to M1’ phonotypical switch is driven at least in part due to adipocyte derived CCN4, along with other unknown mechanisms [170]. CCN4 alone does not initiate the release of inflammatory cytokines from adipocytes, however CCN4 imparts protective effects on LPS-treated adipocytes (3T3-L1) by preventing cell apoptosis and injury [188].

Murahovschi and colleagues have reported increased CCN4 expression and release during adipocyte differentiation, however no effect of CCN4 on adipocyte differentiation [170]. Yet another conflicting report suggests that CCN4 expression significantly decreases during adipocyte differentiation from preadipocytes to mature adipocytes and negatively regulates adipogenesis by physically interacting and redirecting transcriptional factor peroxisome proliferator-activated receptor gamma (PPARγ) to proteasomal degradation, which serves as a master regulator of adipocyte differentiation [189]. Potentially CCN4 does not promote new adipocyte formation, however, it can maintain and protect the pre-existing adipocytes that contribute heavily to the circulating CCN4 levels. CCN4 positively self-regulates itself by protecting the source, i.e., adipocytes, which can further aggravate systemic CCN4 levels, worsening the condition.

Increased adiposity and overexpressed CCN4 in adipocytes of obese individuals significantly contribute to the pro-inflammatory cytokine release by stimulating adipose tissue resident macrophages in a paracrine fashion and drives macrophage polarization with can further worsen the condition. Similarly, supraphysiological serum CCN4 levels also lead to insulin resistance by impairing insulin signaling. Taken together, all the evidence indicates that CCN4 is a central player and key contributor towards aggravation and perpetuation of the inflammatory response and insulin desensitization in obesity. Targeting CCN4 could have great therapeutic potential for metabolic disorders that would benefit numerous patients across the globe.

CCN4 in musculoskeletal system-osteoarthritis (OA) and rheumatoid arthritis (RA)

As per CDC, Arthritis is the leading cause of disability affecting nearly 1 out of 4 adults in the US. Arthritis means ‘disease of the joints’ and is usually characterized by chronic inflammation, pain, stiffness, loss of mobility and function due to progressive damage to the joint, bone and cartilage [190]. OA and RA are the most prevalent joint decaying diseases, with diverse etiologies but overlapping clinical hallmarks [191]. While there is no cure for debilitating OA and RA, treatment paradigms focus on symptomatic relief using a combination of pain management and physical therapy. A better functional and mechanistic understanding of the disease will assist in the discovery of novel diagnostic and therapeutic biomarkers for developing novel treatment strategies [190]. Given the diverse pharmacological effects of CCN4 in the human body, emerging evidence indicates the involvement of the Wnt-pathway in joint diseases.

The significance of CCN4 in the pathophysiology of OA and RA has been abundantly demonstrated by numerous researchers over the last decades. Substantial evidence implicates deleterious effects of CCN4 in the development of musculoskeletal disorders. Differential gene expression and transcriptomics analysis revealed significantly higher CCN4 expression in the cartilage of OA patients as compared to healthy controls [52, 100, 192,193,194,195,196]. Spatial expression profile of CCN4 reveals moderate to weak expression in the superficial layer, matrix and synovial perivascular cells of the knee and hip of RA and OA patients [192]. Another study revealed notable CCN4 upregulation in both synovium and cartilage specimens from OA-human patients and collagenase-induced OA mouse model [52, 195]. CCN4 augmentation has been shown to stimulate chondrocytes, synovial cells, and macrophages, to induce the expression of matrix-degrading proteolytic enzymes such as MMPs that impart deleterious effects on the joint tissue of OA and RA patients. Adenovirus-mediated CCN4 overexpression in knee joints of naïve mice significantly damaged cartilage by inducing MMP13, MMP9, ADAMTS-4 and ADAMTS-5 in synovium and cartilage, exacerbating the condition [52]. In addition, treatment with recombinant human CCN4 increases MMP1, MMP2, MMP3, MMP9 and MMP13 mRNA expression in human OA synovial specimens [197]. Interestingly, CCN4 expression is directly corelated with OA severity and was densely expressed in the most damaged and degraded areas of the joint, confirming its detrimental effects, and highlighting the key role of CCN4 in OA and RA progression [198]. These findings were also confirmed in another study that demonstrated a direct role of CCN4 in the pathogenesis of OA utilizing CCN4-knockdown approach in three different experimental models of OA, that is, intra-articular collagenase induced (CIOA), anterior cruciate ligament transection (ACLT) and destabilization of the medial meniscus (DMM) model. Cartilage degradation was significantly decreased in CCN4/ mice in all three OA-models compared to WT (wild type). These effects were attributed to decreased expression of protease, MMP3, MMP9, ADAMTS-4 and ADAMTS-5 in the synovium of CCN4/ mice, suggesting CCN4 is one of the key culprits in OA-pathogenesis, [197]. Interestingly, miR-128-3p expression is notably decreased in human OA tissue, and overexpression of miR-128-3p significantly decreased CCN4 expression. Furthermore, CCN4 mediates chondrocyte apoptosis, inflammation, and ECM degradation via PI3K/Akt/ NF-κB pathway, an effect that was inhibited by miR-128-3p providing protection against the harmful effects of CCN4, emerging as a novel therapeutic target for OA [196]. Besides low miR-128-3p expression, elevated levels of TGF-β in OA drive CCN4 expression in chondrocytes [199]. CCN4 can also contribute to OA pathology by skewing TGF-β signaling in chondrocytes from protective/ non-hypertrophic ALK-5/Smad 2/3 pathway towards damaging/ hypertrophic ALK-1/ Smad 1/5/8 pathway [200].

As previously described CCN4 binds to certain integrins, predominantly αVβ5, αVβ3 and αVβ1 to mediate its functional effects in fibrosis and cancer. Similarly, researchers have also shown that CCN4 engages integrins expressed on chondrocytes and synovial fibroblasts in OA. Contrary to the previously reported degenerative effects of CCN4 on chondrocyte matrix, another study shows that CCN4 displays a protective effect on primary human OA articular chondrocytes by inhibiting senescence and apoptosis. This effect was blocked in the presence of either αVβ3 antibody or a potent small molecule PI3K inhibitor (LY294002), suggesting that CCN4 mediated protective effects are αVβ3/PI3K/Akt dependent [198]. In another study, pretreatment with αVβ5 integrin blocking antibody, but not αVβ3 or α5β1 significantly reduces CCN4 induced concentration and time dependent increase in IL-6 production in OA synovial fibroblasts. Furthermore, CCN4-dependent IL-6 production was also obliterated in the presence of PI3K (Wortmannin and LY294002), Akt (Akti) and NF-κB (TPCK and PCTC) inhibitors suggesting that CCN4 drives pro-inflammatory cytokine IL-6 release via αVβ5/PI3K/Akt/NF-κB pathway in OA [201]. Together with CCN4, the expression of integrin αV and α5 subunit was significantly higher in human OA cartilage as compared to controls and assists CCN4-dependent chondrocyte differentiation [199]. Chondrocytes are terminally differentiated cells, possessing poor self-restoring capacity leading to longer recovery time after the avascular cartilage injury. Chondrocyte dedifferentiation causes multiple phenotypical changes, that accelerates hypertrophy, matrix calcification, degradation, and fibrosis. Dedifferentiated chondrocyte markers were notably upregulated in OA cartilage, suggesting that in conjunction with other mediators, CCN4 also promotes the destructive dedifferentiation process, aggravating disease progression [202]. Contrary to the general consensus, CCN4 also promotes chondrocyte proliferation, independent of integrins [199]. Treatment with CCN4 have been shown to also drive primary human OA chondrocyte migration, however the migratory chondrocytes were not phenotypically characterized to conclude whether they are non-differentiated which could facilitate repair or dedifferentiated which can cause damage [203].

Stimulation of human-osteoblast-like cells with recombinant human CCN4 dose-dependently increases mitogenic activity assessed by BrdU incorporation and osteoblastic differentiation measured by alkaline phosphatase activity [49]. Another study reported positive influence of CCN4 on bone formation. CCN4 overexpression both in-vivo and in-vitro in osteogenic hBMSC drives osteogenesis, increased bone volume and thickness by increasing bone morphogenic protein 2 (BMP-2) expression and activity in an α5β1 integrin dependent manner [204]. Furthermore, stimulation of mesenchymal stem cells (MSC) with recombinant CCN4 stimulates proliferation in a dose-dependent manner via BMP-3 induction, as CCN4-siRNA mediated knockdown significantly reduced the mitogenic effects of BMP-3 [205]. CCN4 also promotes recruitment, adhesion, and migration of monocytes by dose-dependently increasing VCAM-1 expression in osteoarthritic synovial lining, an effect that was abolished in the presence of an α6β1 or αVβ5 integrin neutralizing antibody. CCN4 also increases activation of protein kinase C (PKCδ), JNK, AP-1 and Syk (Spleen tyrosine kinase) proteins, all of which are necessary for CCN4-mediated VCAM-1 upregulation [206]. Based on literature evidence, CCN4 elicits diverse and sometimes opposite functional effects depending on the cell-type and the spatial expression pattern within the musculoskeletal system. Another study conducted in 304 postmenopausal Japanese women indicates that genetic variations such as single nucleotide polymorphism in CCN4 gene locus has been linked with spinal osteoarthritis determined by radiographical observations such as disc space narrowing, endplate sclerosis and osteophyte formation, again highlighting its therapeutic utility as a novel diagnostic biomarker [207].

Conclusion and future directions

From the literature summarized herein, aberrant CCN4 expression is highly correlated with adverse clinical outcomes and demonstrates the significant role of CCN4 in diverse pathophysiological conditions, such as cancer, fibrosis, metabolic disorders, and arthritis. Altered CCN4 expression may not be the only factor controlling vast array of cellular functions, such as cell proliferation, migration, invasion, and wound healing, but it could be modulating and working in conjunction with other mitogenic signaling molecules, growth factors and inflammatory mediators in driving pathogenesis, as shown in Fig. 3.

Fig. 3
figure 3

Functional effects of CCN4 in different diseases. CCN4 drives many cellular processes such as cell-proliferation, migration, invasion, epithelial to mesenchymal transition (EMT), apoptosis, wound healing, repair, and angiogenesis. Some of these functional effects are overlapping amongst a diverse array of pathological conditions like cancer, fibrosis, obesity, and inflammatory diseases. The image was created with BioRender.com

As more translational studies unearth the CCN4-dependent molecular mechanisms, it is highly likely that CCN4 could be involved in the progression of many more undiscovered co-morbid pathologies. However, the slow growing research on CCN family is largely due to some major challenges in the field that remains to be addressed in future research. Besides the full-length CCN4 protein, identifying the domain specific function is crucial to understand the contribution of the variable linker in CCN biology. Development of domain specific detection tools for human and murine biological tissues or fluids is crucial to understand the tissue specific degradome/cleavage pattern of CCN4 in clinical models. In conjugation with protein–protein interaction studies, researchers could possibly identify functionally active domain(s) for monoclonal antibody mediated targeted therapy. Furthermore, an indirect approach could also be utilized by targeting the key proteolytic enzyme to prevent CCN4 cleavage and subsequently the formation of functionally active single or multi-domain structures. The proportion and expression kinetics of these multi-modular truncated CCN4 variants could reveal the functional redundancy and/or diverse effects in various diseases. Novel tools and reagents are also required to carefully assess the tightly regulated spaciotemporal expression and half-life of these individual or multi-domain structures which yet remains imperative to uncover its holistic biological significance. In addition, CCN4 signaling in diverse cell/tissue specific context still remains understudied. Multi-omics-based approaches at both transcript (RNA seq) and protein level (Proteomics) could be utilized to unravel other CCN4-dependent downstream targets and/or pathways. Also, immunoassays for multi-analyte profiling could further shed light on CCN4-dependend inflammatory secretome-signature. Furthermore, homo- or hetero-multimerization within CCN family members adds to the preexisting complexities. In addition, identifying the cognate receptor/s for CCN family members is important.

Another major challenge is to study crosstalk between CCN family of proteins. One of the major gaps in the current research is how other CCN members interact and modulate CCN4 in a synergistic or antagonistic manner in a pathological context. Some CCN family members have opposing effects with respect to CCN4. A bispecific targeting approach could also be utilized to target antithetical or synergistic CCN members to attain better therapeutic outcomes. However, given the high degree of structural and sequential homology, it is of paramount importance to first understand why some CCN members have opposite functional effects in a disease specific context. Minor changes in the amino acid sequence can bring about huge variations in the overall protein folding and 3-dimensional structure, affecting the surface charge and substrate binding potency, which could be utilized for targeted strategies. Given that CCN4 drives diverse cellular processes in a tissue-specific context, a comprehensive analysis encompassing the molecular network between all the CCN family members remains indispensable. Taken together, the results from the literature reviewed here suggest that CCN4 plays a critical role in the development and progression of diverse pathologies and is emerging as a promising candidate with therapeutic potential yet untapped.

Availability of data and materials

Not applicable.

Abbreviations

ADAM28:

Disintegrin and metalloproteinase domain-containing protein 28

Akt:

Protein kinase B

AP-1:

Activator protein-1

AT-II:

Type II alveolar epithelial cells

BMI:

Body mass index

BMP:

Bone morphogenetic protein

CCL:

Chemokine C–C motif ligand

CCN4:

Cellular communication network factor 4

CD:

Cluster of differentiation

CKD:

Chronic kidney diseases

CT:

C-terminal cysteine knot

CTGF:

Connective tissue growth factor

CXCL:

Chemokine C-X-C motif ligand

Cyr6:

Cysteine-rich protein 61

DN:

Diabetic nephropathy

ECM:

Extracellular matrix

EGFR:

Epidermal growth factor receptor

EMT:

Epithelial to mesenchymal transition

ERK:

Extracellular signal-regulated kinase

FAK:

Focal adhesion kinase

FGF:

Fibroblast growth factor

FN1:

Fibronectin

GLUT1:

Glucose transporter 1

HIF-1α:

Hypoxia inducible factor 1 alpha

ICAM-1:

Intracellular adhesion kinase 1

IFN-γ:

Interferon gamma

IGFBP:

Insulin-like growth factor binding protein like domain

IL-6:

Interlukin-6

IPF:

Idiopathic pulmonary fibrosis

IRS:

Insulin receptor substrate 1

JNK:

C-Jun N-terminal kinase

LRP:

Low-density lipoprotein receptor-related proteins

MEK:

Mitogen-activated protein kinase kinase

miRNA:

MicroRNA

MMP:

Matrix metalloproteases

MRTF:

Myocardin-related transcription factor

mTOR:

Mechanistic target of rapamycin

NF-κB:

Nuclear factor kappa B

NOV:

Nephroblastoma overexpressed

OA:

Osteoarthritis

PI3K:

Phosphoinositol 3-kinase

PKC:

Protein kinase C

PPARγ:

Peroxisome proliferator-activated receptor gamma

PTM:

Post-translational modification

RA:

Rheumatoid arthritis

TGF-β:

Transforming growth factor: β

TLR:

Toll-like receptor

TSP1:

Thrombospondin-homology type 1 repeat

VCAM-1:

Vascular cell adhesion protein 1

VEGF:

Vascular endothelial growth factor

VWC:

von Willebrand factor C repeat

WISP1:

Wnt-1 induced secreted protein-1

αSMA:

α-Smooth muscle actin

References

  1. Bornstein P. Matricellular proteins: an overview. J Cell Commun Signal. 2009;3(3–4):163–5.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Perbal B, Tweedie S, Bruford E. The official unified nomenclature adopted by the HGNC calls for the use of the acronyms, CCN1–6, and discontinuation in the use of CYR61, CTGF, NOV and WISP 1–3 respectively. J Cell Commun Signal. 2018;12(4):625–9.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Brigstock DR, Goldschmeding R, Katsube KI, Lam SCT, Lau LF, Lyons K, et al. Proposal for a unified CCN nomenclature. J Clin Pathol Mol Pathol. 2003;56(2):127–8.

    Article  CAS  Google Scholar 

  4. Gurbuz I, Chiquet-Ehrismann R. CCN4/WISP1 (WNT1 inducible signaling pathway protein 1): A focus on its role in cancer. Int J Biochem Cell Biol. 2015;62:142–6. https://doi.org/10.1016/j.biocel.2015.03.007.

    Article  CAS  PubMed  Google Scholar 

  5. Sun C, Zhang H, Liu X. Emerging role of CCN family proteins in fibrosis. J Cell Physiol. 2021;236(6):4195–206.

    Article  CAS  PubMed  Google Scholar 

  6. Pennica D, Swanson TA, Welsh JW, Roy MA, Lawrence DA, Lee J, et al. WISP genes are members of the connective tissue growth factor family that are up-regulated in Wnt-1-transformed cells and aberrantly expressed in human colon tumors. Proc Natl Acad Sci U S A. 1998;95(25):14717–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Berschneider B, Königshoff M. WNT1 inducible signaling pathway protein 1 (WISP1): A novel mediator linking development and disease. Int J Biochem Cell Biol. 2011;43(3):306–9. https://doi.org/10.1016/j.biocel.2010.11.013.

    Article  CAS  PubMed  Google Scholar 

  8. Walker JM. CCN proteins. In: Takigawa M, editor. Springer protocols, vol. XIV. 2nd ed. Berlin: Springer; 2017. p. 430. https://doi.org/10.1007/978-1-0716-2744-0.

    Chapter  Google Scholar 

  9. Jun J-I, Lau LF. Taking aim at the extracellular matrix: CCN proteins as emerging therapeutic targets. Nat Rev Drug Discov. 2011;10(12):945–63. https://doi.org/10.1038/nrd3599.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Yoshida K, Munakata H. Connective tissue growth factor binds to fibronectin through the type I repeat modules and enhances the affinity of fibronectin to fibrin. Biochim Biophys Acta Gen Subj. 2007;1770(4):672–80.

    Article  CAS  Google Scholar 

  11. Hoshijima M, Hattori T, Inoue M, Araki D, Hanagata H, Miyauchi A, et al. CT domain of CCN2/CTGF directly interacts with fibronectin and enhances cell adhesion of chondrocytes through integrin α5β1. FEBS Lett. 2006;580(5):1376–82.

    Article  CAS  PubMed  Google Scholar 

  12. Francischetti IMB, Kotsyfakis M, Andersen JF, Lukszo J. Cyr61/CCN1 displays high-affinity binding to the somatomedin B 1-44 domain of vitronectin. PLoS ONE. 2010;5(2):1–12.

    Article  Google Scholar 

  13. Nishida T, Kubota S, Fukunaga T, Kondo S, Yosimichi G, Nakanishi T, et al. CTGF/Hcs24, hypertrophic chondrocyte-specific gene product, interacts with perlecan in regulating the proliferation and differentiation of chondrocytes. J Cell Physiol. 2003;196(2):265–75.

    Article  CAS  PubMed  Google Scholar 

  14. Chen N, Leu SJ, Todorović V, Lam SCT, Lau LF. Identification of a novel integrin αvβ3 binding site in CCN1 (CYR61) critical for pro-angiogenic activities in vascular endothelial cells. J Biol Chem. 2004;279(42):44166–76.

    Article  CAS  PubMed  Google Scholar 

  15. Kireeva ML, Lam SC, Lau LF, Biol LFMC. Adhesion of human umbilical vein endothelial cells to the immediate-early gene product Cyr61 is mediated through integrin αvβ3. J Biol Chem. 1998;273(5):3090–6. https://doi.org/10.1074/jbc.273.5.3090.

    Article  CAS  PubMed  Google Scholar 

  16. Lin CG, Leu S-J, Chen N, Tebeau CM, Lin S-X, Yeung C-Y, et al. CCN3 ( NOV ) is a novel angiogenic regulator of the CCN protein family. J Biol Chem. 2003;278(26):24200–8. https://doi.org/10.1074/jbc.M302028200.

    Article  CAS  PubMed  Google Scholar 

  17. Abd El Kader T, Kubota S, Anno K, Tanaka S, Nishida T, Furumatsu T, et al. Direct interaction between CCN family protein 2 and fibroblast growth factor 1. J Cell Commun Signal. 2014;8(2):157–63.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Aoyama E, Kubota S, Takigawa M. CCN2/CTGF binds to fibroblast growth factor receptor 2 and modulates its signaling. FEBS Lett. 2012;586(24):4270–5. https://doi.org/10.1016/j.febslet.2012.10.038.

    Article  CAS  PubMed  Google Scholar 

  19. Nishida T, Kubota S, Aoyama E, Janune D, Maeda A, Takigawa M. Effect of CCN2 on FGF2-induced proliferation and MMP9 and MMP13 productions by chondrocytes. Endocrinology. 2011;152(11):4232–41.

    Article  CAS  PubMed  Google Scholar 

  20. Inoki I, Shiomi T, Hashimoto G, Enomoto H, Nakamura H, Makino K-I, et al. Connective tissue growth factor binds vascular endothelial growth factor (VEGF) and inhibits VEGF-induced angiogenesis. FASEB J. 2002;16(2):219–21.

    Article  CAS  PubMed  Google Scholar 

  21. Abreu JG, Ketpura NI, Reversade B, De Robertis EM. Connective-tissue growth factor (CTGF) modulates cell signalling by bmp and TGF-β. Nat Cell Biol. 2002;4(8):599–604.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Bai T, Chen CC, Lau LF. The matricellular protein CCN1 activates a pro-inflammatory genetic program in murine macrophages. J Immunol. 2010;184(6):1–7.

    Article  Google Scholar 

  23. Chen N, Chen CC, Lau LF. Adhesion of human skin fibroblasts to Cyr61 is mediated through integrin α6β1 and cell surface heparan sulfate proteoglycans. J Biol Chem. 2000;275(32):24953–61. https://doi.org/10.1074/jbc.M003040200.

    Article  CAS  PubMed  Google Scholar 

  24. Kireeva ML, Latinkić BV, Kolesnikova TV, Chen CC, Yang GP, Abler AS, et al. Cyr61 and Fisp12 are both ECM-associated signaling molecules: activities, metabolism, and localization during development. Exp Cell Res. 1997;233(1):63–77.

    Article  CAS  PubMed  Google Scholar 

  25. Desnoyers L, Arnott D, Pennica D. WISP-1 binds to decorin and biglycan. J Biol Chem. 2001;276(50):47599–607. https://doi.org/10.1074/jbc.M108339200.

    Article  CAS  PubMed  Google Scholar 

  26. Juric V, Chen CC, Lau LF. TNFα-induced apoptosis enabled by CCN1/CYR61: pathways of reactive oxygen species generation and cytochrome C release. PLoS ONE. 2012;7(2):1–10.

    Article  Google Scholar 

  27. Mercurio S, Latinkic B, Itasaki N, Krumlauf R, Smith JC. Connective-tissue growth factor modulates WNT signalling and interacts with the WNT receptor complex. Development. 2004;131(9):2137–47.

    Article  CAS  PubMed  Google Scholar 

  28. Blalock TD, Gibson DJ, Duncan MR, Tuli SS, Grotendorst GR, Schultz GS. A connective tissue growth factor signaling receptor in corneal fibroblasts. Investig Ophthalmol Vis Sci. 2012;53(7):3387–94.

    Article  CAS  Google Scholar 

  29. Sakamoto K, Yamaguchi S, Ando R, Miyawaki A, Kabasawa Y, Takagi M, et al. The nephroblastoma overexpressed gene (NOV/ccn3) protein associates with Notch1 extracellular domain and inhibits myoblast differentiation via Notch signaling pathway. J Biol Chem. 2002;277(33):29399–405. https://doi.org/10.1074/jbc.M203727200.

    Article  CAS  PubMed  Google Scholar 

  30. Lau LF. Cell surface receptors for CCN proteins. J Cell Commun Signal. 2016;10:121–7. https://doi.org/10.1007/s12079-016-0324-z.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Aoyama E, Kubota S, Khattab HM, Nishida T, Takigawa M. CCN2 enhances RANKL-induced osteoclast differentiation via direct binding to RANK and OPG. Bone. 2015;73:242–8. https://doi.org/10.1016/j.bone.2014.12.058.

    Article  CAS  PubMed  Google Scholar 

  32. Wu CL, Tsai HC, Chen ZW, Wu CM, Li TM, Fong YC, et al. Ras activation mediates WISP-1-induced increases in cell motility and matrix metalloproteinase expression in human osteosarcoma. Cell Signal. 2013;25(12):2812–22. https://doi.org/10.1016/j.cellsig.2013.09.005.

    Article  CAS  PubMed  Google Scholar 

  33. Hou CH, Chiang YC, Fong YC, Tang CH. WISP-1 increases MMP-2 expression and cell motility in human chondrosarcoma cells. Biochem Pharmacol. 2011;81(11):1286–95. https://doi.org/10.1016/j.bcp.2011.03.016.

    Article  CAS  PubMed  Google Scholar 

  34. Bradham DM, Igarashi A, Potter RL, Grotendorst GR. Connective tissue growth factor: a cysteine-rich mitogen secreted by human vascular endothelial cells is related to the SRC-induced immediate early gene product CEF-10. J Cell Biol. 1991;114(6):1285–94.

    Article  CAS  PubMed  Google Scholar 

  35. Phanish MK, Winn SK, Dockrell MEC. Connective tissue growth factor-(CTGF, CCN2)—a marker, mediator and therapeutic target for renal fibrosis. Nephron Exp Nephrol. 2010;114(3):83–92.

    Article  Google Scholar 

  36. Yang W, Wagener J, Wolf N, Schmidt M, Kimmig R, Winterhager E, et al. Impact of CCN3 (NOV) glycosylation on migration/invasion properties and cell growth of the choriocarcinoma cell line Jeg3. Hum Reprod. 2011;26(10):2850–60.

    Article  CAS  PubMed  Google Scholar 

  37. Soon LL, Yie TA, Shvarts A, Levine AJ, Su F, Tchou-Wong KM. Overexpression of WISP-1 down-regulated motility and invasion of lung cancer cells through inhibition of Rac activation. J Biol Chem. 2003;278(13):11465–70. https://doi.org/10.1074/jbc.M210945200.

    Article  CAS  PubMed  Google Scholar 

  38. Neupane S, Berardinelli SJ, Cameron DC, Grady RC, Komatsu DE, Percival CJ, et al. O-fucosylation of thrombospondin type 1 repeats is essential for ECM remodeling and signaling during bone development. Matrix Biol. 2022;107:77–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Pendurthi UR, Tran TT, Post M, Rao LVM. Proteolysis of CCN1 by plasmin: functional implications. Cancer Res. 2005;65(21):9705–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Choi J, Lin A, Shrier E, Lau LF, Grant MB, Chaqour B. Degradome products of the matricellular protein CCN1 as modulators of pathological angiogenesis in the retina. J Biol Chem. 2013;288(32):23075–89. https://doi.org/10.1074/jbc.M113.475418.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Hashimoto G, Inoki I, Fujii Y, Aoki T, Ikeda E, Okada Y. Matrix metalloproteinases cleave connective tissue growth factor and reactivate angiogenic activity of vascular endothelial growth factor 165. J Biol Chem. 2002;277(39):36288–95. https://doi.org/10.1074/jbc.M201674200.

    Article  CAS  PubMed  Google Scholar 

  42. Guillon-munos A, Oikonomopoulou K, Smith CR, Canepa S, Reverdiau P, Heuze N, et al. Kallikrein-related peptidase 12 hydrolyzes matricellular proteins of the CCN family and modifies interactions of CCN1 and CCN5 with growth factors. J Biol Chem. 2011;286(29):25505–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Mochizuki S, Tanaka R, Shimoda M, Onuma J, Fujii Y, Jinno H, et al. Connective tissue growth factor is a substrate of ADAM28. Biochem Biophys Res Commun. 2010;402(4):651–7.

    Article  CAS  PubMed  Google Scholar 

  44. Kaasbøll OJ, Gadicherla AK, Wang JH, Monsen VT, Hagelin EMV, Dong MQ, et al. Connective tissue growth factor (CCN2) is a matricellular preproprotein controlled by proteolytic activation. J Biol Chem. 2018;293(46):17953–70.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Butler GS, Connor AR, Sounni NE, Eckhard U, Morrison CJ, Noël A, et al. Degradomic and yeast 2-hybrid inactive catalytic domain substrate trapping identifies new membrane-type 1 matrix metalloproteinase (MMP14) substrates: CCN3 (Nov) and CCN5 (WISP2). Matrix Biol. 2017;59:23–38. https://doi.org/10.1016/j.matbio.2016.07.006.

    Article  CAS  PubMed  Google Scholar 

  46. Perbal B. Alternative splicing of CCN mRNAs… It has been upon us. J Cell Commun Signal. 2009;3(2):153–7.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Tanaka S, Sugimachi K, Saeki H, Kinoshita J, Ohga T, Shimada M, et al. A novel variant of WISP1 lacking a Von Willebrand type C module overexpressed in scirrhous gastric carcinoma. Oncogene. 2001;20(39):5525–32.

    Article  CAS  PubMed  Google Scholar 

  48. Tanaka S, Sugimachi K, Kameyama T, Maehara SI, Shirabe K, Shimada M, et al. Human WISP1v, a member of the CCN family, is associated with invasive cholangiocarcinoma. Hepatology. 2003;37(5):1122–9.

    Article  CAS  PubMed  Google Scholar 

  49. Inkson CA, Ono M, Kuznetsov SA, Fisher LW, Gehron P, Young MF. TGF-β1 and WISP-1/CCN-4 can regulate each other’s activity to cooperatively control osteoblast function. J Cell Biochem. 2009;104(5):1865–78.

    Article  Google Scholar 

  50. Cervello M, Giannitrapani L, Labbozzetta M, Notarbartolo M, D’Alessandro N, Lampiasi N, et al. Expression of WISPs and of their novel alternative variants in human hepatocellular carcinoma cells. Ann N Y Acad Sci. 2004;1028:432–9.

    Article  CAS  PubMed  Google Scholar 

  51. Yanagita T, Kubota S, Kawaki H, Kawata K, Kondo S, Takano-Yamamoto T, et al. Expression and physiological role of CCN4/Wnt-induced secreted protein 1 mRNA splicing variants in chondrocytes. FEBS J. 2007;274(7):1655–65.

    Article  CAS  PubMed  Google Scholar 

  52. Blom AB, Brockbank SM, Van Lent PL, Van Beuningen HM, Geurts J, Takahashi N, et al. Involvement of the Wnt signaling pathway in experimental and human osteoarthritis: prominent role of Wnt-induced signaling protein 1. Arthr Rheum. 2009;60(2):501–12.

    Article  CAS  Google Scholar 

  53. Chen YZ, Sun DQ, Zheng Y, Zheng GK, Chen RQ, Lin M, et al. WISP1 silencing confers protection against epithelial–mesenchymal transition of renal tubular epithelial cells in rats via inactivation of the wnt/β-catenin signaling pathway in uremia. J Cell Physiol. 2019;234(6):9673–86.

    Article  CAS  PubMed  Google Scholar 

  54. Madrigano J. Interleukin-18/WNT1-inducible signaling pathway protein-1 signaling mediates human saphenous vein smooth muscle cell proliferation. Occup Env Med. 2008;23(1):1–7.

    Google Scholar 

  55. Winkler J, Abisoye-Ogunniyan A, Metcalf KJ, Werb Z. Concepts of extracellular matrix remodelling in tumour progression and metastasis. Nat Commun. 2020. https://doi.org/10.1038/s41467-020-18794-x.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Liu Y, Song Y, Ye M, Hu X, Wang ZP, Zhu X. The emerging role of WISP proteins in tumorigenesis and cancer therapy. J Transl Med. 2019. https://doi.org/10.1186/s12967-019-1769-7.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Xi Y, LaCanna R, Ma HY, N’Diaye EN, Gierke S, Caplazi P, et al. A WISP1 antibody inhibits MRTF signaling to prevent the progression of established liver fibrosis. Cell Metab. 2022;34(9):1377-1393.e8. https://doi.org/10.1016/j.cmet.2022.07.009.

    Article  CAS  PubMed  Google Scholar 

  58. Kadam AH, Schnitzer JE. Characterization of acute lung injury in the bleomycin rat model. Physiol Rep. 2023;11(5):1–13.

    Article  Google Scholar 

  59. Kim MS, Kim SH, Jeon D, Kim HY, Lee K. Changes in expression of cytokines in polyhexamethylene guanidine-induced lung fibrosis in mice: comparison of bleomycin-induced lung fibrosis. Toxicology. 2017;2018(393):185–92.

    Google Scholar 

  60. Königshoff M, Kramer M, Balsara N, Wilhelm J, Amarie OV, Jahn A, Rose F, Fink L, Seeger W, Schaefer L, Günther A, Eickelberg O. WNT1-inducible signaling protein-1 mediates pulmonary fibrosis in mice and is upregulated in humans with idiopathic pulmonary fibrosis. J Clin Invest. 2009;119(4):772–87.

    PubMed  PubMed Central  Google Scholar 

  61. Wang A-R, Yan X-Q, Zhang C, Du C-Q, Long W-J, Zhan D, et al. Characterization of Wnt1-inducible signaling pathway protein-1 in obese children and adolescents. Curr Med Sci. 2018;38(5):868–74.

    Article  CAS  PubMed  Google Scholar 

  62. Maiese K. Prospects and perspectives for WISP1 (CCN4) in diabetes mellitus. Curr Neurovasc Res. 2020;17(3):327–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Klee S, Lehmann M, Wagner DE, Baarsma HA, Konigshoff M. WISP1 mediates IL-6-dependent proliferation in primary human lung fibroblasts. Sci Rep. 2016;6:1–11. https://doi.org/10.1038/srep20547.

    Article  CAS  Google Scholar 

  64. Venkatachalam K, Venkatesan B, Valente AJ, Melby PC, Nandish S, Reusch JEB, et al. WISP1, a pro-mitogenic, pro-survival factor, mediates tumor necrosis factor-α (TNF-α)-stimulated cardiac fibroblast proliferation but inhibits TNF-α-induced cardiomyocyte death. J Biol Chem. 2009;284(21):14414–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Colston JT, De Rosa SD, Koehler M, Gonzales K, Mestril R, Freeman GL, et al. Wnt-induced secreted protein-1 is a prohypertrophic and profibrotic growth factor. Am J Physiol Circ Physiol. 2007;293:78229.

    Article  Google Scholar 

  66. Wang S, Chong ZZ, Shang YC, Maiese K. WISP1 (CCN4) autoregulates its expression and nuclear trafficking of β-catenin during oxidant stress with limited effects upon neuronal autophagy. Curr Neurovasc Res. 2012;9(2):91–101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Shang YC, Chong ZZ, Wang S, Maiese K. Wnt1 inducible signaling pathway protein 1 (WISP1) targets PRAS40 to govern β-amyloid apoptotic injury of microglia. Curr Neurovasc Res. 2012;9(4):239–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Maiese K. WISP1: clinical insights for a proliferative and restorative member of the CCN family. Curr Neurovasc Res. 2014;11:378.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Brown JS, Amend SR, Austin RH, Gatenby RA, Hammarlund EU, Pienta KJ. Updating the definition of cancer. Mol Cancer Res. 2023;21(11):1142–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Shuel SL. Targeted cancer therapies. Can Fam Phys. 2022;68(7):515–8.

    Article  Google Scholar 

  71. Jia Q, Xu B, Zhang Y, Ali A, Liao X. CCN family proteins in cancer: insight into their structures and coordination role in tumor microenvironment. Front Genet. 2021;12:1–13.

    Article  CAS  Google Scholar 

  72. Yaribeygi H, Atkin SL, Sahebkar A. Wingless-type inducible signaling pathway protein-1 (WISP1) adipokine and glucose homeostasis. J Cell Physiol. 2019;234(10):16966–70.

    Article  CAS  PubMed  Google Scholar 

  73. Chen PP, Li WJ, Wang Y, Zhao S, Li DY, Feng LY, et al. Expression of Cyr61, CTGF, and WISP-1 correlates with clinical features of lung cancer. PLoS ONE. 2007;2(6):1–9.

    Article  Google Scholar 

  74. Li Y, Wang F, Liu T, Lv N, Yuan X, Li P. WISP1 induces ovarian cancer via the IGF1/αvβ3/Wnt axis. J Ovarian Res. 2022. https://doi.org/10.1186/s13048-022-01016-x.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Wu J, Long Z, Cai H, Du C, Liu X, Yu S, et al. High expression of WISP1 in colon cancer is associated with apoptosis, invasion and poor prognosis. Oncotarget. 2016;7(31):49834–47.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Zhu Y, Li W, Yang Y, Li Y, Zhao Y. WISP1 indicates poor prognosis and regulates cell proliferation and apoptosis in gastric cancer via targeting AKT/mTOR signaling pathway. Am J Transl Res. 2020;12(11):7297–311.

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Chiang KC, Yeh CN, Chung LC, Feng TH, Sun CC, Chen MF, et al. WNT-1 inducible signaling pathway protein-1 enhances growth and tumorigenesis in human breast cancer. Sci Rep. 2015;5:1–12.

    Article  Google Scholar 

  78. Zhang H, Luo H, Hu Z, Peng J, Jiang Z, Song T, et al. Targeting WISP1 to sensitize esophageal squamous cell carcinoma to irradiation. Oncotarget. 2015. https://doi.org/10.18632/oncotarget.3358.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Jung EK, Kim S, Yoon TM, Lee KH, Kim HK, Lee DH, et al. WNT1-inducible signaling pathway protein-1 contributes to tumor progression and treatment failure in oral squamous cell carcinoma. Oncol Lett. 2017;14:1719–24.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Wang QY, Feng YJ, Ji R. High expression of WISP1 promotes metastasis and predicts poor prognosis in hepatocellular carcinoma. Eur Rev Med Pharmacol Sci. 2020;24(20):10445–51.

    PubMed  Google Scholar 

  81. Fernandez A, Deng W, McLaughlin SL, Pirkey AC, Rellick SL, Razazan A, et al. Cell communication network factor 4 promotes tumor-induced immunosuppression in melanoma. EMBO Rep. 2022;23(4):1–16.

    Article  Google Scholar 

  82. Davies SR, Watkins G, Mansel RE, Jiang WG. Differential expression and prognostic implications of the CCN family members WISP-1, WISP-2, and WISP-3 in human breast cancer. Ann Surg Oncol. 2007;14(6):1909–18.

    Article  PubMed  Google Scholar 

  83. Taghavi A, Akbari ME, Bahremani MH. Gene expression profiling of the 8q22-24 position in human breast cancer: TSPYL5, MTDH, ATAD2 and CCNE2 genes are implicated in oncogenesis, while WISP1 and EXT1 genes may predict a risk of metastasis. Oncol Lett. 2016;12:3845–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Zhang H, Li W, Huang P, Lin L, Ye H, Lin D, et al. Expression of CCN family members correlates with the clinical features of hepatocellular carcinoma. Oncol Rep. 2015;33(3):1481–92.

    Article  CAS  PubMed  Google Scholar 

  85. Shao H, Cai L, Grichnik JM, Livingstone AS, Velazquez OC, Liu ZJ. Activation of Notch1 signaling in stromal fibroblasts inhibits melanoma growth by upregulating WISP-1. Oncogene. 2011;30(42):4316–26.

    Article  CAS  PubMed  Google Scholar 

  86. Anderson NM, Simon MC. The tumor microenvironment. Curr Biol. 2020;30(16):R921-5. https://doi.org/10.1016/j.cub.2020.06.081.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Tao W, Chu C, Zhou W, Huang Z, Zhai K, Fang X, et al. Dual Role of WISP1 in maintaining glioma stem cells and tumor-supportive macrophages in glioblastoma. Nat Commun. 2020. https://doi.org/10.1038/s41467-020-16827-z.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Zimmer AM, Pan YK, Chandrapalan T, Kwong RWM, Perry SF. Loss-of-function approaches in comparative physiology: is there a future for knockdown experiments in the era of genome editing ? J Exp Biol. 2019;222(7): jeb175737.

    Article  PubMed  Google Scholar 

  89. Prelich G. Gene overexpression: uses, mechanisms, and interpretation. Genetics. 2012;190:841–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Beltrao P, Bork P, Krogan NJ, Van NV. Evolution and functional cross-talk of protein post-translational modifications. Mol Syst Biol. 2013;9(714):1–13.

    Google Scholar 

  91. Tai H-C, Chang A-C, Yu H-J, Huang C-Y, Lai Y-W, Sun H-L, et al. Osteoblast-derived WISP-1 increases VCAM-1 expression and enhances prostate cancer metastasis by down-regulating miR-126. Oncotarget. 2014;5(17):7589.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Chang A-C, Lien M-Y, Tsai M-H, Hua C-H. WISP-1 promotes epithelial-mesenchymal transition in oral squamous cell carcinoma cells via the miR-153–3p/Snail axis. Cancers. 2019;11(1903):1–16.

    Google Scholar 

  93. Chuang JY, Chang AC, Chiang IP, Tsai MH, Tang CH. Apoptosis signal-regulating kinase 1 is involved in WISP-1-promoted cell motility in human oral squamous cell carcinoma cells. PLoS ONE. 2013;8(10): e78022.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Chuang JY, Chen PC, Tsao CW, Chang AC, Lein MY, Lin CC, et al. WISP-1, a novel angiogenic regulator of the CCN family, promotes oral squamous cell carcinoma angiogenesis through VEGF-A expression. Oncotarget. 2015;6(6):4239–52.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Lin CC, Chen PC, Lein MY, Tsao CW, Huang CC, Wang SW, et al. WISP-1 promotes VEGF-C-dependent lymphangiogenesis by inhibiting miR-300 in human oral squamous cell carcinoma cells. Oncotarget. 2016;7(9):9993–10005.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Lee H-L, Chiou H-L, Wang S-S, Hung S-C, Chou M-C, Yang S-F, et al. WISP1 genetic variants as predictors of tumor development with urothelial cell carcinoma. Urol Oncol Semin Orig Investig. 2018;36(4):160.e15-160.e21. https://doi.org/10.1016/j.urolonc.2017.11.023.

    Article  CAS  Google Scholar 

  97. Chen CT, Lee HL, Chiou HL, Chou CH, Wang PH, Yang SF, et al. Impacts of wnt1-inducible signaling pathway protein 1 polymorphism on hepatocellular carcinoma development. PLoS ONE. 2018;13(6):1–15.

    Article  Google Scholar 

  98. He J, Wang Z, Wang Y, Zou T, Li XP, Cao L, et al. The effects of WISP1 polymorphisms on the prognosis of lung cancer patients with platinum-based chemotherapy. Pharmgenomics Pers Med. 2021;14:1193–203.

    PubMed  PubMed Central  Google Scholar 

  99. Chen J, Yin J, Li X, Wang Y, Zheng Y, Qian C, et al. WISP1 polymorphisms contribute to platinum-based chemotherapy toxicity in lung cancer patients. Int J Mol Sci. 2014;15:21011–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Van Den Bosch MHJ, Ramos YFM, Den Hollander W, Bomer N, Nelissen RGHH, Bovée JVMG, et al. Increased WISP1 expression in human osteoarthritic articular cartilage is epigenetically regulated and decreases cartilage matrix production. Rheumatology. 2019;58(6):1065–74.

    Article  PubMed  Google Scholar 

  101. Xie D, Nakachi K, Wang H, Elashoff R, Koeffler HP. Elevated levels of connective tissue growth factor, WISP-1, and CYR61 in primary breast cancers associated with more advanced features 1. Cancer Res. 2001;61:8917–23.

    CAS  PubMed  Google Scholar 

  102. Fang F, Xu W, Zhang J, Gu J, Yang G. Ultrasound microbubble-mediated RNA interference targeting WNT1 inducible signaling pathway protein 1 (WISP1) suppresses the proliferation and metastasis of breast cancer cells. Bioengineered. 2022;13(4):11050–60. https://doi.org/10.1080/21655979.2022.2068738.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Ii DJK. Induction of Wnt-inducible signaling protein-1 correlates with invasive breast cancer oncogenesis and reduced type 1 cell-mediated cytotoxic immunity : a retrospective study. PLoS Comput Biol. 2014;10(1): e1003409.

    Article  Google Scholar 

  104. Mirza Z, Ansari S, Iqbal S, Ahmad N, Alganmi N, Banjar H, et al. Identification of novel diagnostic and prognostic gene signature biomarkers for breast cancer using artificial intelligence and machine learning assisted transcriptomics analysis. Cancers. 2023. https://doi.org/10.3390/cancers15123237.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Jia H, Janjanam J, Wu SC, Wang R, Pano G, Celestine M, et al. The tumor cell - secreted matricellular protein WISP 1 drives pro-metastatic collagen linearization. FASEB J. 2019. https://doi.org/10.1096/fj.202000713RR.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Liu S, Wu D, Sun X, Fan Y, Zha R, Jalali A, et al. Mechanical stimulations can inhibit local and remote tumor progression by downregulating WISP1. FASEB J. 2020;34(9):12847–59.

    Article  CAS  PubMed  Google Scholar 

  107. Li Y, Zhu Z, Hou X, Sun Y. LncRNA AFAP1-AS1 promotes the progression of colorectal cancer through miR-195–5p and WISP1. J Oncol. 2021;2021:1–10.

    Google Scholar 

  108. Nagai Y, Watanabe M, Ishikawa S, Karashima R, Kurashige J, Iwagami S, et al. Clinical significance of Wnt-induced secreted protein-1 ( WISP-1/CCN4) in esophageal squamous cell carcinoma. Anticancer Res. 2011;998:991–7.

    Google Scholar 

  109. Yang C, Li F, Zhou W, Huang J. Knockdown of long non-coding RNA CCAT2 suppresses growth and metastasis of esophageal squamous cell carcinoma by inhibiting the β-catenin/WISP1 signaling pathway. J Int Med Res. 2021;49(5):030006052110199.

    Article  Google Scholar 

  110. Jing D, Zhang Q, Yu H, Zhao Y, Shen L. Identification of WISP1 as a novel oncogene in glioblastoma. Int J Oncol. 2017;51:1261–70.

    Article  CAS  PubMed  Google Scholar 

  111. Wang L, Sun J, Gao P, Su K, Wu H, Junli L, et al. Wnt1-inducible signaling protein 1 regulates laryngeal squamous cell carcinoma glycolysis and chemoresistance via the YAP1/TEAD1/GLUT1 pathway. J Cell Physiol. 2019. https://doi.org/10.1002/jcp.28253.

    Article  PubMed  PubMed Central  Google Scholar 

  112. Wang L, Sun J, Cao H. MicroRNA-384 regulates cell proliferation and apoptosis through directly targeting WISP1 in laryngeal cancer. J Cell Biochem. 2018;2019:3018–26.

    Google Scholar 

  113. Yan J, Lei J, Chen L, Deng H, Dong D, Jin T, et al. Human leukocyte antigen F locus adjacent transcript 10 overexpression disturbs WISP1 protein and mRNA expression to promote hepatocellular carcinoma progression. Hepatology. 2018;68(6):2268–84.

    Article  CAS  PubMed  Google Scholar 

  114. Margalit O, Eisenbach L, Amariglio N, Kaminski N, Harmelin A, Pfeffer R, et al. Overexpression of a set of genes, including WISP-1, common to pulmonary metastases of both mouse D122 Lewis lung carcinoma and B16-F10.9 melanoma cell lines. Br J Cancer. 2003;89(2):314–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Clausen MJAM, Melchers LJ, Mastik MF, Slagter-menkema L. Identification and validation of WISP1 as an epigenetic regulator of metastasis in oral squamous cell carcinoma. Genes Chromosom Cancer. 2016;59:45–59.

    Article  Google Scholar 

  116. Tsai H-C, Tzeng H-E, Huang C-Y, Huang Y-L, Tsai C-H, Wang S-W, et al. WISP-1 positively regulates angiogenesis by controlling VEGF-A expression in human osteosarcoma. Cell Death Dis. 2017;8(4):e2750-11. https://doi.org/10.1038/cddis.2016.421.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Graumann J, Finkernagel F, Reinartz S, Stief T, Brödje D, Renz H, et al. Multi-platform affinity proteomics identify proteins linked to metastasis and immune suppression in ovarian cancer plasma. Front Oncol. 2019;9:1–12.

    Article  Google Scholar 

  118. Ono M, Inkson CA, Sonn R, Kilts TM, de Castro LF, Maeda A, et al. WISP1/CCN4: a potential target for inhibiting prostate cancer growth and spread to bone. PLoS ONE. 2013;8(8): e71709.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Mao A, Tang J, Tang D, Wang F, Liao S, Yuan H, et al. MicroRNA-29b-3p enhances radiosensitivity through modulating WISP1-mediated mitochondrial apoptosis in prostate cancer cells. J Cancer. 2020;11:6356.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Gaudreau PO, Clairefond S, Class CA, Boulay PL, Chrobak P, Allard B, et al. WISP1 is associated to advanced disease, EMT and an inflamed tumor microenvironment in multiple solid tumors. Oncoimmunology. 2019;8(5):1–18. https://doi.org/10.1080/2162402X.2019.1581545.

    Article  Google Scholar 

  121. Chang AC, Chen PC, Lin YF, Su CM, Liu JF, Lin TH, et al. Osteoblast-secreted WISP-1 promotes adherence of prostate cancer cells to bone via the VCAM-1/integrin α4β1 system. Cancer Lett. 2018;426:47–56. https://doi.org/10.1016/j.canlet.2018.03.050.

    Article  CAS  PubMed  Google Scholar 

  122. Chang KS, Chen ST, Sung HC, Hsu SY, Lin WY, Hou CP, et al. WNT1 inducible signaling pathway protein 1 is a stroma-specific secreting protein inducing a fibroblast contraction and carcinoma cell growth in the human prostate. Int J Mol Sci. 2022;23(19):1–18.

    Article  Google Scholar 

  123. Chang AC, Lin LW, Chen YC, Chen PC, Liu SC, Tai HC, et al. The ADAM9/WISP-1 axis cooperates with osteoblasts to stimulate primary prostate tumor growth and metastasis. Int J Biol Sci. 2023;19(3):760–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Zheng Y, Chen CJ, Lin ZY, Li JX, Liu J, Lin FJ, et al. Circ_KATNAL1 regulates prostate cancer cell growth and invasiveness through the mir-145-3p/WISP1 pathway. Biochem Cell Biol. 2020;98(3):396–404.

    Article  CAS  PubMed  Google Scholar 

  125. Deng W, Fernandez A, McLaughlin SL, Klinke DJ. WNT1-inducible signaling pathway protein 1 (WISP1/CCN4) stimulates melanoma invasion and metastasis by promoting the epithelial mesenchymal transition. J Biol Chem. 2019;294(14):5261–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Deng W, Fernandez A, McLaughlin SL, Klinke DJ. Cell communication network factor 4 (CCN4/WISP1) shifts melanoma cells from a fragile proliferative state to a resilient metastatic state. Cell Mol Bioeng. 2020;13(1):45–60.

    Article  CAS  PubMed  Google Scholar 

  127. Pirkey AC, Deng W, Norman D, Razazan A, Klinke DJ. Head-to-head comparison of CCN4, DNMT3A, PTPN11, and SPARC as suppressors of anti-tumor immunity. Cell Mol Bioeng. 2023;16(5–6):431–42. https://doi.org/10.1007/s12195-023-00787-7.

    Article  CAS  PubMed  Google Scholar 

  128. Plikus MV, Wang X, Sinha S, Forte E, Thompson SM. Fibroblasts: origins, definitions, and functions in health and disease. Cell. 2021;184(15):3852–72. https://doi.org/10.1016/j.cell.2021.06.024.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Kendall RT, Feghali-bostwick CA, Rauch BH. Fibroblasts in fibrosis: novel roles and mediators. Front Pharmacol. 2014;5:1–13.

    Article  CAS  Google Scholar 

  130. Wynn TA. Cellular and molecular mechanisms of fibrosis. J Pathol. 2008;214:199–210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Abuserewa ST, Duff R, Becker G. Treatment of idiopathic pulmonary fibrosis. Cureus. 2021;13(5):1–10.

    Google Scholar 

  132. Liu J, Lv S, Ma W, Yang D, Zhang X. Effect of WISP1 on paraquat-induced EMT. Toxicol Vitro. 2023;93:105693. https://doi.org/10.1016/j.tiv.2023.105693.

    Article  CAS  Google Scholar 

  133. Sun Z, Yang Z, Wang M, Huang C, Ren Y, Zhang W, et al. Paraquat induces pulmonary fibrosis through Wnt/β-catenin signaling pathway and myofibroblast differentiation. Toxicol Lett. 2020;333(June):170–83. https://doi.org/10.1016/j.toxlet.2020.08.004.

    Article  CAS  PubMed  Google Scholar 

  134. Berschneider B, Ellwanger DC, Baarsma HA, Thiel C, Shimbori C, White ES, et al. MiR-92a regulates TGF-β1-induced WISP1 expression in pulmonary fibrosis. Int J Biochem Cell Biol. 2014;53:432–41. https://doi.org/10.1016/j.biocel.2014.06.011.

    Article  CAS  PubMed  Google Scholar 

  135. Pommier A, Varilh J, Bleuse S, Delétang K, Bonini J, Bergougnoux A, et al. miRNA repertoires of cystic fibrosis ex vivo models highlight miR-181a and miR-101 that regulate WISP1 expression. J Pathol. 2021;253(2):186–97.

    Article  CAS  PubMed  Google Scholar 

  136. Injury E, Fibrosis IP. Epithelial injury and dysfunction in the pathogenesis of idiopathic pulmonary fibrosis. Am J Med Sci. 2020;357(5):374–8.

    Google Scholar 

  137. Singh K, Teyani RL, Moniri NH. Agonists and hydrogen peroxide mediate hyperoxidation of β 2-adrenergic receptor in airway epithelial cells : implications for tachyphylaxis to β 2-agonists in constrictive airway disorders. Biomed Pharmacother. 2023;168:115763. https://doi.org/10.1016/j.biopha.2023.115763.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Heijink IH, Kuchibhotla VNS, Roffel MP, Maes T, Knight DA, Sayers I, et al. Epithelial cell dysfunction, a major driver of asthma development. Allergy. 2020;75:1902–17.

    Article  PubMed  Google Scholar 

  139. Heise RL, Stober V, Cheluvaraju C, Hollingsworth JW, Garantziotis S. Mechanical stretch induces epithelial-mesenchymal transition in alveolar epithelia via hyaluronan activation of innate immunity. J Biol Chem. 2011;286(20):17435–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Stephens S, Palmer J, Konstantinova I, Pearce A, Jarai G, Day E. A functional analysis of Wnt inducible signalling pathway protein−1 (WISP-1/CCN4). J Cell Commun Signal. 2015;9(1):63–72.

    Article  PubMed  PubMed Central  Google Scholar 

  141. Zhang C, Zhu Y, Zhang Y, Gao L, Zhang N, Feng H. Therapeutic potential of umbilical cord mesenchymal stem cells for inhibiting myofibroblastic differentiation of irradiated human lung fibroblasts. Tohoku J Exp Med. 2015;236(3):209–17.

    Article  CAS  PubMed  Google Scholar 

  142. Li X, Liu X, Deng R, Gao S, Yu H, Huang K, et al. Nintedanib inhibits Wnt3a-induced myofibroblast activation by suppressing the Src/b-catenin pathway. Front Pharmacol. 2020;11:1–11.

    Google Scholar 

  143. Lehmann M, Buhl L, Alsafadi HN, Klee S, Hermann S, Mutze K, et al. Differential effects of Nintedanib and Pirfenidone on lung alveolar epithelial cell function in ex vivo murine and human lung tissue cultures of pulmonary fibrosis. Respir Res. 2018;19(1):1–12.

    Article  Google Scholar 

  144. Huang X, Cai H, Ammar R, Zhang Y, Wang Y, Ravi K, et al. Molecular characterization of a precision-cut rat liver slice model for the evaluation of antifibrotic compounds. Am J Physiol Gastrointest Liver Physiol. 2019;316(1):G15-24.

    Article  CAS  PubMed  Google Scholar 

  145. Pivovarova-Ramich O, Loske J, Hornemann S, Markova M, Seebeck N, Rosenthal A, et al. Hepatic wnt1 inducible signaling pathway protein 1 (WISP-1/CCN4) associates with markers of liver fibrosis in severe obesity. Cells. 2021;10(5):1048.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Chen Z, Ding X, Jin S, Pitt B, Zhang L, Billiar T, et al. WISP1-αvβ3 integrin signaling positively regulates TLR-triggered inflammation response in sepsis induced lung injury. Sci Rep. 2016;6:1–11. https://doi.org/10.1038/srep28841.

    Article  CAS  Google Scholar 

  147. Rong X, Liu J, Yao X, Jiang T, Wang Y, Xie F. Human bone marrow mesenchymal stem cells-derived exosomes alleviate liver fibrosis through the Wnt/β-catenin pathway. Stem Cell Res Ther. 2019;10(1):1–11.

    Article  Google Scholar 

  148. Jian YC, Wang JJ, Dong S, Hu JW, Hu LJ, Yang GM, et al. Wnt-induced secreted protein 1/CCN4 in liver fibrosis both in vitro and in vivo. Clin Lab. 2014;60(1):29–35.

    CAS  PubMed  Google Scholar 

  149. Li X, Chen Y, Ye W, Tao X, Zhu J, Wu S, et al. Blockade of CCN4 attenuates CCl4-induced liver fibrosis. Arch Med Sci. 2015;11(3):647–53.

    Article  PubMed  PubMed Central  Google Scholar 

  150. Ono M, Masaki A, Maeda A, Kilts TM. CCN4 / WISP1 controls cutaneous wound healing by modulating proliferation, migration and ECM expression in dermal fibroblasts via α 5 β 1 and TNF α. Matrix Biol. 2018;68–69:533–46. https://doi.org/10.1016/j.matbio.2018.01.004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Tank J, Lindner D, Wang X, Stroux A, Gilke L, Gast M, et al. Single-target RNA interference for the blockade of multiple interacting proinflammatory and profibrotic pathways in cardiac fibroblasts. J Mol Cell Cardiol. 2014;66:141–56.

    Article  CAS  PubMed  Google Scholar 

  152. Green LC, Slone S, Anthony SR, Guarnieri AR, Parkins S, Shearer SM, et al. HuR-dependent expression of Wisp1 is necessary for TGFβ-induced cardiac myofibroblast activity. J Mol Cell Cardiol. 2022;2023(174):38–46.

    Google Scholar 

  153. Zhong X, Tu YJ, Li Y, Zhang P, Wang W, Chen SS, et al. Serum levels of WNT1-inducible signaling pathway protein-1 (WISP-1): a noninvasive biomarker of renal fibrosis in subjects with chronic kidney disease. Am J Transl Res. 2017;9(6):2920–32.

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Zhou Y, Chai P, Wang J, Li L, Chen MH. Wingless/int-1induced secreted protein-1: a new biomarker for renal fibrosis. J Biol Regul Homeost Agents. 2021;35(1):33480221.

    Google Scholar 

  155. Hennemeier I, Humpf HU, Gekle M, Schwerdt G. The food contaminant and nephrotoxin ochratoxin A enhances Wnt1 inducible signaling protein 1 and tumor necrosis factor-α expression in human primary proximal tubule cells. Mol Nutr Food Res. 2012;56(9):1375–84.

    Article  CAS  PubMed  Google Scholar 

  156. Yang X, Wang H, Tu Y, Li Y, Zou Y, Li G, et al. WNT1-inducible signaling protein-1 mediates TGF-β1-induced renal fibrosis in tubular epithelial cells and unilateral ureteral obstruction mouse models via autophagy. J Cell Physiol. 2020;235(3):2009–22.

    Article  CAS  PubMed  Google Scholar 

  157. Wang B, Ding C, Ding X, Tesch G, Zheng J, Tian PY, et al. WNT1-inducible signaling pathway protein 1 regulates kidney inflammation through the NF-κB pathway. Clin Sci. 2022;136(1):29–44.

    Article  CAS  Google Scholar 

  158. Ai K, Yi L, Wang Y, Li Y. CircRNA_33702 promotes renal fibrosis by targeting the miR-29b-3p/WNT1-inducible signaling pathway protein 1 pathway. J Pharmacol Exp Ther. 2023;384(1):225–6.

    Article  Google Scholar 

  159. Liang J-N, Zou X, Fang X-H, Xu J-D, Xiao Z, Zhu J-N, et al. The Smad3-miR-29b/miR-29c axis mediates the protective effect of macrophage migration inhibitory factor against cardiac fibrosis. Biochim Biophys Acta Mol Basis Dis. 2019;1865(9):2441–50. https://doi.org/10.1016/j.bbadis.2019.06.004.

    Article  CAS  PubMed  Google Scholar 

  160. Gong X, Wang X, Zhou F. Liver microRNA-29b-3p positively correlates with relative enhancement values of magnetic resonance imaging and represses liver fibrosis. J Biochem. 2020;168(6):603–9.

    Article  CAS  PubMed  Google Scholar 

  161. Xue Y, Fan X, Yang R, Jiao Y, Li Y. miR-29b-3p inhibits post-infarct cardiac fibrosis by targeting FOS. Biosci Rep. 2020;40(9):1–11.

    Article  Google Scholar 

  162. Fan X, Gao Y, Zhang X, Lughmani HY, Kennedy DJ, Haller ST, et al. A strategic expression method of miR-29b and its anti-fibrotic effect based on RNA-sequencing analysis. PLoS ONE. 2020;15:1–16. https://doi.org/10.1371/journal.pone.0244065.

    Article  CAS  Google Scholar 

  163. Chen H, Fan Y, Jing H, Tang S, Huang Z, Liao M, et al. LncRNA Gm12840 mediates WISP1 to regulate ischemia-reperfusion-induced renal fibrosis by sponging miR-677-5p. Epigenomics. 2020;12(24):2205–18.

    Article  CAS  PubMed  Google Scholar 

  164. Abd El-Kader T, Kubota S, Janune D. Anti-fibrotic effect of CCN3 accompanied by altered gene expression profile of the CCN family. J Cell Commun Signal. 2013;7:11–8.

    Article  PubMed  Google Scholar 

  165. Leask A. Yin and Yang revisited: CCN3 as an anti-fibrotic therapeutic? J Cell Commun Signal. 2015;9(1):97–8.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Leask A. CCN3: a novel anti-fibrotic treatment in end-stage renal disease? J Cell Commun Signal. 2012;6(2):115–6.

    Article  PubMed  PubMed Central  Google Scholar 

  167. Lin X, Li H. Obesity: epidemiology, pathophysiology, and therapeutics. Front Endocrinol. 2021;12(September):1–9.

    Google Scholar 

  168. Coelho M, Oliveira T, Fernandes R. Biochemistry of adipose tissue: an endocrine organ. Arch Med Sci. 2013;9(2):191–200.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Ouchi N, Parker JL, Lugus JJ, Walsh K. Adipokines in inflammation and metabolic disease. Nat Rev Immunol. 2011;11:85–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Murahovschi V, Pivovarova O, Ilkavets I, Dmitrieva RM, Döcke S, Keyhani-Nejad F, et al. WISP1 Is a novel adipokine linked to inflammation in obesity. Diabetes. 2015;64(3):856–66.

    Article  CAS  PubMed  Google Scholar 

  171. Martinerie C, Garcia M, Do TTH, Antoine B, Moldes M, Dorothee G, et al. NOV/CCN3: A new adipocytokine involved in obesity-associated insulin resistance. Diabetes. 2016;65(9):2502–15.

    Article  CAS  PubMed  Google Scholar 

  172. Tacke C, Aleksandrova K, Rehfeldt M, Murahovschi V, Markova M, Kemper M, et al. Assessment of circulating Wnt1 inducible signalling pathway protein 1 (WISP-1)/CCN4 as a novel biomarker of obesity. J Cell Commun Signal. 2018;12(3):539–48.

    Article  PubMed  Google Scholar 

  173. Barchetta I, Cimini FA, Capoccia D, De Gioannis R, Porzia A, Mainiero F, et al. WISP1 is a marker of systemic and adipose tissue inflammation in dysmetabolic subjects with or without type 2 diabetes. J Endocr Soc. 2017;1:660–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Hörbelt T, Tacke C, Markova M, de Wiza DH, Van de Velde F, Bekaert M, et al. The novel adipokine WISP1 associates with insulin resistance and impairs insulin action in human myotubes and mouse hepatocytes. Diabetologia. 2018;61(9):2054–65.

    Article  PubMed  Google Scholar 

  175. Jung TW, Kang C, Goh J, Chae SI, Kim HC, Lee TJ, et al. WISP1 promotes non-alcoholic fatty liver disease and skeletal muscle insulin resistance via TLR4/JNK signaling. J Cell Physiol. 2018;233(8):6077–87.

    Article  CAS  PubMed  Google Scholar 

  176. Fernandez-ruiz R, García-alamán A, Esteban Y, Mir-coll J, Serra-navarro B, Fontcuberta-pisunyer M, et al. Wisp1 is a circulating factor that stimulates proliferation of adult mouse and human beta cells. Nat Commun. 2020. https://doi.org/10.1038/s41467-020-19657-1.

    Article  PubMed  PubMed Central  Google Scholar 

  177. Seong JK, Park J. Identification of differentially expressed mRNA during pancreas regeneration of rat by mRNA differential display. Biochem Biophys Res Commun. 2002;299:806–12.

    Article  PubMed  Google Scholar 

  178. Kapasa M, Sera I, Gavalas A, Kossida S. Genomics identification of phylogenetically conserved enhancer elements implicated in pancreas development in the WISP1 and CTGF orthologs. Genomics. 2008;92:301–8.

    Article  CAS  PubMed  Google Scholar 

  179. Chen Y, Li P, Lin M, Jiang Y, Tan G, Huang L, et al. Silencing of METTL3 prevents the proliferation, migration, epithelial-mesenchymal transition, and renal fibrosis of high glucose-induced HK2 cells by mediating WISP1 in m6A-dependent manner. Aging. 2024;16(2):1237–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Liu L, Hu J, Yang L, Wang N, Liu Y, Wei X, et al. Association of WISP1/CCN4 with risk of overweight and gestational diabetes mellitus in Chinese pregnant women. Dis Markers. 2020;2020:1–10.

    CAS  Google Scholar 

  181. Sahin Ersoy G, Altun Ensari T, Subas S, Giray B, Simsek EE, Cevik O. WISP1 is a novel adipokine linked to metabolic parameters in gestational diabetes mellitus. J Matern Neonatal Med. 2017;30(8):942–6.

    Article  CAS  Google Scholar 

  182. French DM, Kaul RJ, Souza ALD, Crowley CW, Bao M, Frantz GD, et al. WISP-1 is an osteoblastic regulator expressed during skeletal development and fracture repair. Am J Pathol. 2004;165(3):855–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Khanna D, Khanna S, Khanna P, Kahar P, Patel BM. Obesity: a chronic low-grade inflammation and its markers. Cureus. 2022;14(2): e2271.

    Google Scholar 

  184. Lumeng CN, Bodzin JL, Saltiel AR. Obesity induces a phenotypic switch in adipose tissue macrophage polarization. J Clin Investig. 2007;117(1):175–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Lacy P, Murray RZ, Moraes-vieira PM. The macrophage switch in obesity. Development. 2016;6(Jan):1–11.

    Google Scholar 

  186. Chang JS, Kim TH, Kong ID. Exercise intervention lowers aberrant serum WISP-1 levels with insulin resistance in breast cancer survivors : a randomized controlled trial. Sci Rep. 2020;10(2):1–9. https://doi.org/10.1038/s41598-020-67794-w.

    Article  CAS  Google Scholar 

  187. Yu Z, Wang T, Zhang L, Yang X, Li Q, Ding X. WISP1 and TLR4 on macrophages contribute to ventilator-induced lung injury. Inflammation. 2020;43(2):425–32.

    Article  CAS  PubMed  Google Scholar 

  188. Liu L, Xu S, Li P, Li L. A novel adipokine WISP1 attenuates lipopolysaccharide-induced cell injury in 3T3-L1 adipocytes by regulating the PI3K/Akt pathway. Obes Res Clin Pract. 2022;16(2):122–9. https://doi.org/10.1016/j.orcp.2022.03.001.

    Article  PubMed  Google Scholar 

  189. Ferrand N, Béreziat V, Moldes M, Zaoui M, Larsen AK, Sabbah M. WISP1/CCN4 inhibits adipocyte differentiation through repression of PPAR γ activity. Sci Rep. 2017;7:1–12.

    Article  CAS  Google Scholar 

  190. Senthelal S, Li J, Ardeshirzadeh S, Thomas MA. Arthritis. StatPearls: Treasure Island; 2023.

    Google Scholar 

  191. Eakin GS, Amodeo KL, Kahlon RS. Arthritis and its public health burden. Del J Public Heal. 2017;3(1):36–44.

    Article  Google Scholar 

  192. Komatsu M, Nakamura Y, Maruyama M, Abe K, Watanapokasin R, Kato H. Expression profiles of human CCN genes in patients with osteoarthritis or rheumatoid arthritis. J Orthop Sci. 2015;20(4):708–16.

    Article  CAS  PubMed  Google Scholar 

  193. Macsai CE, Georgiou KR, Foster BK, Zannettino ACW, Xian CJ. Microarray expression analysis of genes and pathways involved in growth plate cartilage injury responses and bony repair. Bone. 2012;50(5):1081–91. https://doi.org/10.1016/j.bone.2012.02.013.

    Article  CAS  PubMed  Google Scholar 

  194. Geyer M, Grässel S, Straub RH, Schett G, Dinser R, Grifka J, et al. Differential transcriptome analysis of intraarticular lesional vs intact cartilage reveals new candidate genes in osteoarthritis pathophysiology. Osteoarthr Cartil. 2009;17(3):328–35.

    Article  CAS  Google Scholar 

  195. Chou CH, Wu CC, Song IW, Chuang HP, Lu LS, Chang JH, et al. Genome-wide expression profiles of subchondral bone in osteoarthritis. Arthr Res Ther. 2013;15(6):R190.

    Article  Google Scholar 

  196. Chen S, Li B. MiR-128-3p post-transcriptionally inhibits WISP1 to suppress apoptosis and inflammation in human articular chondrocytes via the PI3K/AKT/NF-κB signaling pathway. Cell Transpl. 2020;29(8):1–13.

    Google Scholar 

  197. van den Bosch MH, Blom AB, Kram V, Maeda A, Sikka S, Gabet Y, et al. WISP1/CCN4 aggravates cartilage degeneration in experimental osteoarthritis. Osteoarthr Cartil. 2017;25(11):1900–11. https://doi.org/10.1016/j.joca.2017.07.012.

    Article  Google Scholar 

  198. Cheng C, Tian J, Zhang F, Deng Z, Tu M, Li L, et al. WISP1 protects against chondrocyte senescence and apoptosis by regulating αvβ3 and PI3K/Akt pathway in osteoarthritis. DNA Cell Biol. 2021;40(4):629–37.

    Article  CAS  PubMed  Google Scholar 

  199. Zhang M, Meng QC, Yang XF, Mu WD. TGF-β1/WISP1/Integrin-α interaction mediates human chondrocytes dedifferentiation. Eur Rev Med Pharmacol Sci. 2020;24(17):8675–84.

    CAS  PubMed  Google Scholar 

  200. Van den Bosch MH, Blom AB, van Lent PL, van Beuningen HM, Blaney Davidson EN, van der Kraan PM, et al. Canonical Wnt signaling skews TGF-β signaling in chondrocytes towards signaling via ALK1 and smad 1/5/8. Cell Signal. 2014;26(5):951–8. https://doi.org/10.1016/j.cellsig.2014.01.021.

    Article  CAS  PubMed  Google Scholar 

  201. Hou CH, Tang CH, Hsu CJ, Hou SM, Liu JF. CCN4 induces IL-6 production through αvβ5 receptor, PI3K, Akt, and NF-κB singling pathway in human synovial fibroblasts. Arthr Res Ther. 2013;15(1):R19.

    Article  CAS  Google Scholar 

  202. Charlier E, Deroyer C, Ciregia F, Malaise O, Neuville S, Plener Z, et al. Chondrocyte dedifferentiation and osteoarthritis (OA). Biochem Pharmacol. 2019;165(Feb):49–65. https://doi.org/10.1016/j.bcp.2019.02.036.

    Article  CAS  PubMed  Google Scholar 

  203. Timmermans RGM, Blom AB, Bloks NGC, Nelissen RGHH, van der Linden EHMJ, van der Kraan PM, et al. CCN4/WISP1 promotes migration of human primary osteoarthritic chondrocytes. Cartilage. 2023;14(1):67–75.

    Article  CAS  PubMed  Google Scholar 

  204. Ono M, Inkson CA, Kilts TM, Young MF. WISP-1/CCN4 regulates osteogenesis by enhancing BMP-2 activity. J Bone Miner Res. 2011;26(1):193–208.

    Article  CAS  PubMed  Google Scholar 

  205. Cernea M, Tang W, Guan H, Yang K. Wisp1 mediates Bmp-3 stimulated mesenchymal stem cell proliferation. J Mol Endocrinol. 2016;56(1):39–46.

    Article  CAS  PubMed  Google Scholar 

  206. Liu JF, Hou SM, Tsai CH, Huang CY, Hsu CJ, Tang CH. CCN4 induces vascular cell adhesion molecule-1 expression in human synovial fibroblasts and promotes monocyte adhesion. Biochim Biophys Acta Mol Cell Res. 2013;1833(5):966–75. https://doi.org/10.1016/j.bbamcr.2012.12.023.

    Article  CAS  Google Scholar 

  207. Urano T, Narusawa K, Shiraki M, Usui T, Sasaki N, Hosoi T, et al. Association of a single nucleotide polymorphism in the WISP1 gene with spinal osteoarthritis in postmenopausal Japanese women. J Bone Miner Metab. 2007;25(4):253–8.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors wish to acknowledge Dr. Kannan Thirunavukkarasu and Dr. Marta A. Witek for providing critical feedback on the manuscript. The authors would also like to thank the Department of Biotherapeutic Enabling Biology for all the help and support.

Funding

The work was supported and funded by Eli Lilly and Company.

Author information

Authors and Affiliations

Authors

Contributions

KS conceptualized, compiled, reviewed literature, edited, designed figures, tables and wrote the manuscript. SSO supervised, reviewed, edited, and provided critical feedback on the manuscript. All the authors read and approved the final manuscript.

Corresponding author

Correspondence to Sunday S. Oladipupo.

Ethics declarations

Ethics approval and consent to participate

Not Applicable.

Consent for publication

Not applicable.

Competing interests

The authors have no competing interests related to this work.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Singh, K., Oladipupo, S.S. An overview of CCN4 (WISP1) role in human diseases. J Transl Med 22, 601 (2024). https://doi.org/10.1186/s12967-024-05364-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12967-024-05364-8

Keywords