Next Article in Journal
Decidualized Endometrial Stromal Cells Promote Mitochondrial Beta-Oxidation to Produce the Octanoic Acid Required for Implantation
Previous Article in Journal
Impact of Dropping on Postharvest Physiology of Tomato Fruits Harvested at Green and Red Ripeness Stages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unveiling the Impact of BMP9 in Liver Diseases: Insights into Pathogenesis and Therapeutic Potential

1
Center of Infectious Diseases, West China Hospital of Sichuan University, Chengdu 610041, China
2
Laboratory of Infectious and Liver Diseases, Institute of Infectious Diseases, West China Hospital of Sichuan University, Chengdu 610041, China
3
Department of Gastroenterology, Kanazawa University Hospital, Kanazawa 9208641, Japan
*
Authors to whom correspondence should be addressed.
Biomolecules 2024, 14(8), 1013; https://doi.org/10.3390/biom14081013
Submission received: 17 July 2024 / Revised: 8 August 2024 / Accepted: 13 August 2024 / Published: 15 August 2024
(This article belongs to the Section Molecular Biomarkers)

Abstract

:
Bone morphogenetic proteins (BMPs) are a group of growth factors belonging to the transforming growth factor β(TGF-β) family. While initially recognized for their role in bone formation, BMPs have emerged as significant players in liver diseases. Among BMPs with various physiological activities, this comprehensive review aims to delve into the involvement of BMP9 specifically in liver diseases and provide insights into the complex BMP signaling pathway. Through an enhanced understanding of BMP9, we anticipate the discovery of new therapeutic options and potential strategies for managing liver diseases.

Graphical Abstract

1. Introduction

Bone morphogenetic proteins (BMPs) constitute a group of signaling molecules that belong to the transforming growth factor β(TGF-β) superfamily [1]. Their significance extends beyond their initial identification as inducers of bone formation [2], as they are now recognized for their vital roles in diverse biological processes such as embryonic development, tissue regeneration, and organ homeostasis. BMPs exert regulatory control over cell proliferation, differentiation, apoptosis, and metabolism across various tissues and organs in the body [3,4,5]. These multifaceted effects position BMPs as crucial factors in both physiological and pathological conditions [6,7,8,9].
The TGF-β superfamily comprises three subfamilies: the TGF-β subfamily, the bone morphogenetic protein and growth differentiation factor (BMP/GDF) subfamily, and the activin and inhibin subfamily. The BMP/GDF subfamily, characterized by sequence homology and functional diversity, can be further divided into distinct groups: BMP2/4, BMP5/6/7/8A/8B, BMP9/BMP10, BMP12/13/14, BMP15/GDF9, BMP11/GDF8, BMP3/BMP3B, and GDF1/3 [9,10]. These proteins, known by alternative names, interact with various type I and II receptors, phosphorylate Co-Smads downstream, translocate into the nucleus, and regulate targeted genes (Figure 1) [11,12]. Among these members, BMP9 (also known as GDF2) stands out for its predominant expression in the liver. Initially recognized for its role in hepatogenesis [13], hematopoiesis [14], osteogenesis [15], and chondrogenesis [16,17,18]. BMP9 has emerged as a regulator of hepatic glucose and lipid metabolism [19], iron balance [20], cholinergic neuron differentiation [21,22], endothelial tip/stalk cell selection [23], angiogenesis [24,25,26], lymph angiogenesis [27], liver regeneration, and fibrosis [28,29]. BMP9 is an important mediator of vascular quiescence, and the loss-of-function mutations in the BMP9—ENG—ALK1 signaling pathway lead to hereditary hemorrhagic telangiectasia (HHT), an inherited disorder characterized by vascular dysplasia and potential bleeding tendencies [30,31]. Furthermore, BMP9 deficiency is associated with pulmonary arterial hypertension (PAH), a progressive obliterative vasculopathy affecting the distal pulmonary arterial circulation [30,32,33,34]. Additionally, BMP9 has been implicated in liver diseases, including nonalcoholic fatty liver disease (NAFLD), liver fibrosis, and hepatocellular carcinoma(HCC) [35]. However, many aspects regarding its precise involvement and underlying mechanisms in these conditions remain elusive [36]. This review aims to provide a comprehensive overview of BMP9’s roles and mechanisms in liver diseases. By examining its involvement, we can explore potential therapeutic options for the treatment of liver diseases.

2. Receptors and Intracellular Signaling of BMP9

The BMP9 protein, which is encoded by the GDF2 gene, is located at 10q11.22. BMP9 is initially synthesized as a precursor protein called pre-pro-BMP9, which consists of a total of 429 amino acids (aa). This includes a signal peptide (22 aa), a prodomain (297 aa), and a mature protein region (110 aa). The pre-pro-BMP9 then homodimerizes to form unprocessed pro-BMP9, which remains in an inactive state [37,38]. The processing of BMP typically occurs intracellularly within the Golgi apparatus. After being cleaved by a serine endoprotease, two active forms of BMP9 are synthesized. These include the short mature form (25 kDa) and the complexed form (100 kDa), both of which result from the Furin cleavage of pre-pro-BMP9 [37,38]. The stable BMP9 dimers can either form with an intermolecular disulfide bond (D-form) or without it (M-form). By changing the redox potential, these two forms can be converted into each other, leading to a significant alteration in BMP9 stability. The M-form is more susceptible to redox-dependent cleavage by proteases found in serum [39]. In both humans and mice, biologically active concentrations of BMP9 can be detected in the circulation. The circulating forms of BMP9 comprise the active complexed form of BMP9 (60%) and the inactive pro-BMP9 (40%), with both forms exceeding 100 kDa in size (Figure 2) [40,41]. It has been reported that BMP9 and BMP10 in plasma can form heterodimers, which contribute to the majority of the biological BMP activity [42]. Currently, commercial kits utilizing the Enzyme-Linked Immuno-Sorbent Assay (ELISA) are available for detecting various forms of BMP9 in serum. Some antibodies specifically target the mature form of BMP9, displacing the cleaved prodomain from the BMP9 complexed form. As a result, this type of ELISA quantifies mature BMP9 regardless of its association with the prodomain. However, it does not recognize the unprocessed BMP9 precursor. Alternatively, a combination of anti-prodomain BMP9 and anti-mature BMP9 antibodies as capture and detection antibodies, respectively, has proven effective in detecting the unprocessed forms of BMP9 [43].
BMP9 exerts its biological functions by binding to specific receptors. BMP receptors form heterotetrameric complexes, consisting of two BMP type I receptors and two type II receptors. The BMP type I receptors include ALK1, ALK2, ALK3 (BMPRIA), and ALK6 (BMPRIB), while the BMP type II receptors consist of BMPR-II, ActR-IIA, and ActR-IIB (Figure 1) [33,44]. BMP9 exhibits sub-nanomolar affinities (EC50 of around 50 pg/mL) when binding to the type I receptor ALK1 [45]. Both mature BMP9 and the BMP9 complexed form bind to ALK1 with the same level of affinity. Both the D-form and M-form of mature BMP9 are capable of binding to the prodomain and ALK1 [39]. However, BMP9’s affinity for ALK2 is significantly lower compared to ALK1. Additionally, the BMP9 complexed form has an even lower affinity for ALK2 than the mature BMP9. Regarding the type II receptors, some studies suggest that BMP9 binds with similar affinities to ALK1 and ActR-IIB, while showing lower affinities towards BMPRII and even lower affinities towards ActR-IIA [31,37,38,45]. Nevertheless, contradictory findings suggest that BMP9 has a much higher affinity for ALK1 than for the type II receptors, correlating with their expression levels in endothelial cells. Moreover, it has been observed that the binding of BMP9 to ALK2 is greatly facilitated when BMPR-II or ActR-IIA/B receptors are co-expressed, indicating that in such cases, binding to the type II receptors precedes binding to the type I receptor. Also, even in the absence of ALK1, cells are still capable of responding to BMP9 through ALK2 [46,47]. The expression of ALK2 and ActR-IIA was detected in whole human hepatic tissue, while BMPR-II expression was observed in primary mouse hepatocytes and HCC cell lines [48]. ALK1 was found to be expressed on the hepatoma cell line HLE [49]. It is intriguing to analyze the expression patterns of BMP9 receptors in various HCC cells, as this can provide valuable insights into the mechanisms by which BMP9 influences HCC [35].
BMP9 exhibits high affinity direct binding to the co-receptor endoglin (ENG), which is highly expressed in endothelial cells [50,51,52]. Endoglin enhances the signaling output, as evidenced by increased levels of pSmad1/5/8 [53,54,55]. The binding of BMP9 to the type II receptors leads to the displacement of the prodomain and cell surface endoglin, resulting in the formation of a signaling complex [31,52]. Furthermore, soluble endoglin can be detected in the bloodstream and has traditionally been believed to function as a ligand trap for circulating BMP9 [51]. In parallel, upon binding of TGF-β1, ALK5/TGFBR2 stimulation leads to the phosphorylation of Smad2/3. pSmad2/3 subsequently interacts with Smad4, facilitating its translocation to the nucleus to regulate multiple gene expressions [56]. It should be noted that Smad3 has the ability to directly bind to Smad-binding elements found within gene promoters, leading to an enhancement of transcription. On the other hand, neither Smad2 nor Smad4 possess DNA binding domains. Instead, they function as regulators of Smad3-mediated gene transcription [57,58,59,60]. pSmad1/5/8 could suppress Smad3-dependent gene transcription, and this may be the reason why BMP9—pSmad1/5/8 signaling counteracts TGF-β1—pSmad2/3 signaling [61,62,63,64]. The I-Smads (or Inhibitory Smads) Smad6 and Smad7 inhibit the phosphorylation of Smad1/5/8 and Smad2/3 (Figure 3) [65,66].
Upon binding of BMP9, the type II receptor initiates the phosphorylation of the type I receptor, leading to the subsequent phosphorylation of Smad1/5/8 within 5 min. Compared to the D-form of mature BMP9, the M-form shows a lesser degree of sustained induction in Smad1/5/8 phosphorylation [39]. The duration of this phosphorylation period depends on the concentration of BMP9, ranging from a few hours with low doses (0.1 ng/mL) to up to 24 h with higher doses (10 ng/mL) [53,67]. The BMP9/ALK1 complex undergoes rapid endocytosis and can either be recycled back to the membrane or degraded through proteolysis [68]. Additionally, it has been observed that in endothelial cells, BMP9 can induce the phosphorylation of Smad2, although at a much lower level compared to Smad1/5/8 phosphorylation [31,69,70]. Apart from the well-studied canonical Smad-dependent pathway, significant research has been conducted on the non-canonical non-Smad BMP pathways. The downstream effects of BMP9 involve various signaling molecules such as mitogen-activated protein kinases (MAPKs, including p38, extracellular-signal-regulated kinase (ERK), and c-Jun N-terminal kinases (JNK)), Phosphoinositide 3-kinases/AKT (PI3K/AKT), LIM Domain Kinase (LIMK), Rho, Rho-Associated Coiled-Coil-Containing Protein Kinase (ROCK), and TAK1 [71,72,73]. Additionally, there are crucial interactions between Notch, VEGF, Hippo, EGF, and BMP9 [74,75,76,77]. Some of the most prominent target genes of BMP9 include ID1, ID2, and ID3 [78,79], along with other notable targets such as Hepcidin and Snail (Figure 4) [35,80,81,82].
The regulation of BMP9 involves several processes, including transcription, microRNA, and ligand traps. A study on acetaminophen (APAP)-induced acute liver injury (ALI) demonstrated that BMP9 expression is regulated by CCAAT/enhancer binding protein α (C/EBPα), both in vitro and in vivo [83]. Additionally, miR-761 has been identified as a targeting inhibitor of BMP9 [84]. Furthermore, Crossveinless 2 (CV2), soluble ENG (sENG), and soluble chimeric ALK1 protein (ALK1-Fc) have been considered potential ligand traps for BMP9 [41,52,85]. However, our understanding of BMP9 regulation remains incomplete, and further studies are needed to elucidate the regulators and regulatory mechanisms involved.

3. BMP9 and Liver Cell Biology

In both humans and mice, BMP9 is predominantly expressed in the liver and originates from various cell types including hepatocytes, cholangiocytes, and hepatic stellate cells (HSCs), among which intrahepatic biliary epithelial cells express the highest levels of BMP9 [41]. Furthermore, BMP9 expression is observed to increase upon HSC activation [28,86]. In rats, however, BMP9 is primarily produced by HSCs, Kupffer cells (KCs), and endothelial cells [87]. Interestingly, studies have shown a higher expression of BMP9 in aged mice compared to their younger counterparts [83].
BMP9 plays a crucial regulatory role in liver cell functions. In healthy liver, it has the ability to inhibit hepatocyte proliferation and epithelial-to-mesenchymal transition (EMT) while preserving the expression of important metabolic enzymes such as cytochrome P450 [28]. Furthermore, BMP9 is involved in the maturation and differentiation of hepatic progenitor cells [88]. It facilitates the functioning of these cells, which are responsible for repopulating the liver and restoring its functionality after severe damage, through the HGF/c-Met signaling pathway [88,89]. The expression of BMP9 in Dissè’s spaces indicates its regenerative role in the liver [29]. In terms of oval cells, BMP9 directly inhibits their expansion through the ALK2—pSmad1/5/8 pathway, as ALK2 expression is significantly higher than that of ALK1 in oval cells [90]. The BMP9/BMP10—ALK1 signaling pathway controls the identity and self-renewal of Kupffer cells (KCs) through a Smad4-dependent pathway. The development of Kupffer cells crucially depends on their crosstalk with hepatic stellate cells via the evolutionarily conserved ALK1-BMP9/10 axis [91]. However, it should be noted that ALK1 is not essential for the maintenance of macrophages located in the lung, kidney, spleen, and brain [92]. The loss of BMP9 in the 129/Ola genetic background leads to spontaneous liver fibrosis due to the capillarization of LSEC and kidney lesions [93,94]. Furthermore, hepatic endothelial ALK1 signaling protects against the development of vascular malformations, preserving organ-specific endothelial differentiation and angiocrine signaling. On the other hand, the loss of ALK1 causes hepatic vascular malformations [95].

4. BMP9 and Viral Hepatitis

Currently, there have been reports indicating an inverse correlation between disease stage and circulating levels of BMP9 in patients with hepatitis C virus (HCV) [96]. However, there is a lack of available references directly addressing the effects of BMP9 on viral hepatitis. Nonetheless, several studies have investigated the interactions of other BMPs with hepatic viral infections. It has been observed that there exists a mutual antagonistic relationship between HCV infection and the BMP—Smad pathway. HCV infection attenuates the induction of hepcidin expression by BMP6, possibly through the TNF-mediated downregulation of the BMP co-receptor hemojuvelin. Conversely, BMP—Smad activity influences antiviral immunity. BMP6 regulates a gene profile reminiscent of type I interferon (IFN) signaling, including the upregulation of interferon regulatory factors (IRFs) and the downregulation of an inhibitor factor of IFN signaling called Ubiquitin specific peptidase 18 (USP18). BMP6 enhances the transcriptional and antiviral responses to IFN while also effectively inhibiting HCV replication independently of IFN, along with related activin proteins. Furthermore, both BMP6 and activin A have been found to suppress the growth of hepatitis B virus (HBV) in cell culture [97]. However, it is important to note that the current studies on the interactions between BMPs and viral hepatitis are limited. Furthermore, BMP9 shows relatively low homology with BMP6. It is not advisable to speculate on the mutual interaction between BMP9 and viral hepatitis based on the interaction between BMP6 and viral hepatitis. Therefore, further research is necessary to fully understand and explore the potential associations and implications of BMP9 in the context of viral hepatitis.

5. BMP9 and Acute Liver Injury

Acute liver injury induced by partial hepatectomy, as well as single injections of carbon tetrachloride (CCl4) or lipopolysaccharide (LPS) in mice, leads to a temporary decrease in hepatic BMP9 mRNA expression. Interestingly, the application of BMP9 following partial hepatectomy has been shown to significantly aggravate liver damage and disrupt the regenerative response [28]. However, the underlying mechanism remains unclear and requires further investigation. Notably, it has been discovered that LPS triggers the secretion of factors such as IL-6 from human liver sinusoidal endothelial cells (LSECs), which, in turn, upregulate BMP9 expression in human liver myofibroblasts [86]. Another intriguing finding is that aging-related C/EBPα plays a role in upregulating BMP9 expression, contributing to the exacerbation of APAP-induced ALI associated with aging. Firstly, BMP9 directly increases the susceptibility of hepatocytes to APAP-induced cell death. Secondly, it contributes to the polarization of M1 macrophages. These results highlight the significant impact of aging-associated C/EBPα on BMP9 expression, leading to the deterioration of APAP-ALI during the aging process [83]. Furthermore, in a model of cholestatic liver injury induced by 3,5-diethoxicarbonyl-1,4-dihydrocollidine (DDC), BMP9 has been identified as a negative regulator of oval cell expansion. The deletion of BMP9 in mice reduces liver damage and fibrosis but intensifies inflammation upon DDC feeding through the activation of the PI3K—AKT, ERK—MAPK, and c-Met signaling pathways. Additionally, due to higher expression of ALK2 compared to ALK1 in oval cells, BMP9 inhibits oval cell expansion via the ALK2—p-Smad1/5/8 pathway [90]. These findings collectively suggest that BMP9 is downregulated during acute or chronic liver injury caused by hepatotoxic substances or partial surgical excision. In such scenarios, elevated levels of BMP9 may further exacerbate liver damage.

6. BMP9 and NAFLD

6.1. BMP9 in Lipid Metabolism and Obesity

The beneficial effects of BMP9 on NAFLD are evident through its negative correlation with liver steatosis. A cross-sectional study observed that patients with both type 2 diabetes mellitus (T2DM) and NAFLD had significantly lower levels of serum BMP9 [98]. Furthermore, the study identified independent factors, such as fasting insulin (FINS), low-density lipoprotein cholesterol (LDL-C), high-density lipoprotein cholesterol (HDL-C), and body mass index (BMI), that influenced serum BMP9 levels. Among the fibroblast growth factor (FGF) family, FGF19 plays a crucial role in regulating glucose and lipid metabolisms [99]. Considering this, increased hepatic BMP9 levels would likely be beneficial, especially in pre-steatotic conditions, as BMP9 directly induces FGF19 in the gut [100]. Similarly to FGF19, FGF21 also plays a critical role in regulating glucose and lipid metabolisms. Animal studies have demonstrated that FGF21 can improve insulin sensitivity in obese animals, reduce body weight, lower the levels of low-density lipoprotein cholesterol, and reverse hepatic steatosis [101,102]. By enhancing FGF21 expression, MB109 (a recombinant derivative of human BMP9) effectively reduces obesity-induced liver pathologies, including inhibiting lipid accumulation in the liver and significantly decreasing serum levels of alanine aminotransferase (ALT), aspartate transaminase (AST), and total cholesterol. Additionally, MB109 exhibits dose-dependent improvements in the glucose and lipid metabolisms of obese mice. However, it is important to note that insufficient doses of MB109 may not adequately address the aforementioned pathological manifestations [103]. BMP9 has also been shown to suppress NAFLD, reduce obesity, improve glucose metabolism, alleviate hepatic steatosis, and decrease inflammation by reducing the promoter chromatin accessibility of Cers6, Fabp4, Fos, and Tlr1 [104]. Another study involving both animal models and cell lines demonstrated that BMP9 ablation promotes liver steatosis, with the mechanism being that BMP9 treatment attenuates triglyceride accumulation by enhancing peroxisome proliferator-activated receptor α (PPARα) promoter activity through the activation of phosphorylated Smad (pSmad) [105].
Adipose tissue plays a crucial role in maintaining energy balance and presents a potential target in the fight against obesity [106]. It is closely associated with energy homeostasis and offers an opportunity to address obesity, insulin resistance (IR), and T2DM. There are two distinct types of adipose tissue: white adipose tissue (WAT) and brown adipose tissue (BAT). WAT primarily serves as a lipid reservoir, while BAT dissipates energy through thermogenesis, generating heat. Furthermore, WAT functions as an endocrine organ by releasing adipokines. In cases of obesity, dysfunctional WAT has been linked to the secretion of pro-inflammatory adipokines, leading to IR, T2DM, and NAFLD/nonalcoholic steatohepatitis (NASH) [107]. On the other hand, BAT has been shown to regulate glucose homeostasis and improve insulin sensitivity. Therefore, inducing Adipose Tissue Browning is considered a strategy to combat obesity [106,108].
It has been discovered that BMP9 plays a role in activating thermogenesis in BAT and promoting the “browning” process in WAT [109]. Consequently, it has gained recognition as a potent inducer of brown adipogenesis with significant potential in combating obesity [110]. Moreover, when exposed to cold temperatures, BMP9 expression is triggered by the binding of cAMP response-element binding protein (CREB) and CREB-binding protein (CBP) to the BMP9 promoter. This activation leads to the stimulation of the Smad1 and MAPK pathways, which subsequently induce the expressions of uncoupling protein 1 (UCP1) and peroxisome proliferator-activated receptor-gamma coactivator 1 alpha (PGC1α), ultimately facilitating the “browning” of white adipose tissue. These findings underscore the substantial therapeutic value of BMP9 in addressing obesity-related NAFLD.
However, a cross-sectional study found a positive correlation between the serum levels of BMP9 and NAFLD/NASH. Specifically, patients with BMP9 levels greater than 1188 pg/mL had more severe NAFLD/NASH, suggesting that BMP9 could serve as a potential biomarker for these conditions [96]. In animal studies, BMP9 promoted nonalcoholic steatohepatitis by polarizing M1 macrophages through enhancing NF-κB signaling activation [111]. The conflicting results may be attributed to the small sample sizes and the use of different animal models in these clinical studies. Thus, future research should focus on precise and large-scale clinical studies with well-designed protocols, as well as comprehensive animal experiments, to better understand the regulatory role of BMP9 in NAFLD. Endoglin is a target of BMP9. An animal model demonstrated that mice fed with a high fat, fructose, and cholesterol (FFC) diet showed elevated serum and hepatic endoglin expression levels. Elevated levels of human serum endoglin lead to increased cholesterol deposition in the liver due to reduced conversion into bile acids. This also alters triglyceride (TAG) metabolism, causing their accumulation in the liver. This occurs through a combination of decreased TAG elimination via β-oxidation and reduced hepatic efflux of TAG. Elevated soluble endoglin (sEng) levels may contribute to the progression of NASH by compromising the crucial protective mechanism that prevents the excessive accumulation of TAG and cholesterol in the liver. These findings emphasize the significance of high serum endoglin levels in individuals susceptible to developing NASH [112].

6.2. BMP9 in Glucose Metabolism and T2DM

T2DM is a significant risk factor for NAFLD, and there is a strong association between these conditions [113,114]. The prevalence of NAFLD in patients with T2DM exceeds 60% [115]. In individuals with T2DM, impaired insulin regulation leads to increased liver fat accumulation, contributing to the development of NAFLD. NAFLD, in turn, worsens insulin resistance, creating a detrimental cycle between the two conditions [116]. Therefore, addressing these interconnected conditions and adopting a holistic approach to treating metabolic diseases could be highly beneficial [117].
Studies have reported a negative association between circulating BMP9 and metabolic syndrome as well as insulin resistance [118]. Circulating BMP9 levels were significantly lower in newly diagnosed T2DM patients compared to healthy subjects. BMP9 levels negatively correlated with HbA1c, fasting blood glucose, oral glucose tolerance tests (OGTT), the area under the curve for glucose (AUC glucose), and the homoeostasis model assessment of insulin resistance (HOMA-IR). Thus, the decreased levels of circulating BMP9 in type 2 diabetes patients serve as a marker of insulin resistance [119]. BMP9 has a protective effect against insulin resistance. Platycodins induce the activation of the BMP9/Smad4 pathway, which corrects blood glucose and lipid metabolism disorders, improves the liver index, and protects the liver function in cases of liver complications related to type 2 diabetes [120]. Conversely, the deletion of BMP9 elevated the phosphoenolpyruvate carboxykinase (PEPCK) protein levels and lowered the levels of InsR and Akt phosphorylation in mouse models fed a high-fat diet, resulting in exacerbated insulin resistance and glucose intolerance [121]. Additionally, BMP9 deletion in non-diabetic rats leads to insulin resistance and glucose intolerance [122]. BMP9 facilitates insulin action by reducing phosphoenolpyruvate carboxykinase expression, inhibiting hepatic glucose production, and increasing glycogen synthesis through the activation of the Smad5—Akt2—glycogen synthase kinase 3 (GSK3) pathway in L6 myotubes [123,124]. From the above research, we can infer that BMP9 possesses insulin-like properties. Therefore, BMP9 could be a potential candidate for enhancing hepatic insulin sensitivity. In conclusion, BMP9 plays a positive role in insulin resistance and T2DM, and it has the potential to become a new target for the treatment of NAFLD (Table 1).

7. BMP9 and Hepatic Fibrosis

Hepatic fibrosis refers to the excessive accumulation of collagen and other extracellular matrix components in the liver. It is a progressive condition that can result from chronic liver injury, including viral hepatitis, alcohol abuse, NAFLD, autoimmune liver diseases, or prolonged exposure to toxins. While the liver has an impressive regenerative capacity, continuous or severe injury can trigger an abnormal wound healing response [127,128,129]. This response involves the activation of HSCs, which are responsible for producing and depositing collagen and other connective tissue proteins [130,131]. Over time, this deposition leads to the formation of scar tissue, impairing liver function and disrupting its normal architecture. As hepatic fibrosis progresses, it can develop into a more advanced stage known as cirrhosis. Cirrhosis is characterized by extensive fibrotic scarring, nodules, and distortion of the liver structure [132]. It can lead to complications such as portal hypertension, impaired liver function, increased risk of liver cancer, and ultimately, liver failure [133]. Considerable research has been conducted to investigate the impact of BMP9, its receptors, and targets on liver fibrosis [81]. BMP9 has been found to have a protective effect on liver fibrosis by controlling the fenestration of liver sinusoidal endothelial cells and modulating the signaling of secreted cytokines, thus attenuating hepatic fibrosis [134]. In mouse models, the deletion of BMP9 resulted in liver fibrosis, specifically in the 129/Ola background [94]. However, it has also been observed that serum BMP9 levels are significantly higher in patients with fibrosis compared to healthy individuals (p < 0.01) [134]. Therefore, further extensive research is needed to clarify whether BMP9 promotes or prevents hepatic fibrosis, as well as to elucidate the specific mechanisms involved. Given that the BMP9 receptors ALK1 and Endoglin, as well as the BMP9 downstream targets Hepcidin and ID1, have significant roles in liver fibrosis, we will comprehensively discuss each of these components individually.

7.1. BMP9–ALK1 Axis in Liver Fibrosis

The activation of liver non-parenchymal cells plays a crucial role in the process of liver fibrosis. Among these cells, ALK1 is the primary receptor for BMP9 in the liver [35]. When ligands bind to ALK1, it triggers the phosphorylation of Smad1, which subsequently promotes the expression of ID1. This induction leads to the differentiation of HSCs into fibroblasts, contributing to the production of extracellular matrix (ECM) proteins [64,135]. In the mouse liver, HSCs serve as both the source and target of BMP9. A deficiency of BMP9 or the overexpression of the selective BMP9 antagonist ALK1-Fc has been found to reduce collagen accumulation and subsequent fibrosis in cases of chronic liver damage in mice [28]. Furthermore, an herbal compound called Cpd861 has shown potential in inhibiting the activation of LX-2 cells by suppressing the TGF-β1–ALK1–Smad1 pathway, thereby restraining the expression of alpha smooth muscle actin (α-SMA). This inhibition was also observed in rats with liver fibrosis induced by bile duct ligation (BDL), where Cpd861 induced the expression of the SnoN protein and inhibited the TGF-β1–Smad signaling pathway, resulting in the attenuation of liver fibrosis [136]. The ectopic expression of BMP9 has been found to exacerbate liver fibrosis in mouse models induced by CCl4 and BDL. BMP9 activates the ALK1–Smad1/5/8–ID1 pathway, promoting the activation of HSCs during the progression of liver fibrosis. Conversely, inhibiting BMP9 expression or neutralizing BMP9 protein using anti-BMP9 monoclonal antibodies could alleviate hepatic fibrosis [134]. Therefore, there is a need to further elucidate the role and molecular mechanisms of the BMP9–ALK1 axis in regulating hepatic fibrosis, which could have significant clinical and biological implications.

7.2. BMP9–Endoglin Axis in Liver Fibrosis

Endoglin is a co-receptor for both TGF-β and BMPs. While it cannot bind ligands on its own, it can interact with BMP9 in the presence of type I or type II signaling receptors. Endoglin is expressed in various cell types including endothelial cells, epithelial cells, fibroblasts, septal myofibroblasts, and HSCs. The upregulation of endoglin has been commonly observed in liver injuries, suggesting its potential as a serum biomarker for liver fibrosis. In a study of pediatric cholestatic liver disease, endoglin, along with interleukin-8 (IL-8) and matrix metalloproteinase-7 (MMP-7), showed a significant correlation with increased liver stiffness in children with biliary atresia [137]. Additionally, patients with alcoholic liver cirrhosis exhibited significantly decreased concentrations of endoglin, zinc, and selenium in their serum. However, patients with liver cirrhosis and carcinoma had elevated levels of endoglin in their serum [138,139]. Patients with HCV-related cirrhosis also demonstrated higher levels of serum endoglin compared to those with fibrosis but without cirrhosis [140]. Furthermore, in patients with advanced HCV-related fibrosis, intrahepatic expression of endoglin was significantly higher than in patients with early fibrosis and normal liver [138,140]. HCV infection can induce liver pathogenesis by increasing endoglin expression and activating the ALK1–Smad1/5 pathway [141]. A specific variant of endoglin, Thr5Met, has been associated with an increased risk of fibrosis development in HCV carriers [142]. Endoglin is involved in fibrogenic Smad signaling [138,143]. It regulates TGF-β and BMP signals in liver fibrosis by inhibiting the ALK5–Smad2/3 pathway and enhancing the ALK1–Smad1/5 pathway [144]. In a mouse model, different isoforms of endoglin were found to be upregulated in liver samples from patients with chronic and acute liver injury. Notably, a deficiency of endoglin in HSCs significantly exacerbated fibrosis, suggesting that endoglin may play a protective role against fibrotic injury through the modulation of TGF-β signaling [145]. In conclusion, endoglin plays a role in fibrogenic Smad signaling and regulates TGF-β and BMP signals in liver fibrosis by modulating the ALK5–Smad2/3 pathway and enhancing the ALK1–Smad1/5 pathway. Further studies are needed to fully understand the role of the BMP9–endoglin axis in liver fibrosis and its underlying mechanisms.

7.3. BMP9–Hepcidin Axis in Liver Fibrosis

Iron deposition in the liver is recognized as a critical factor in the development and progression of liver disease [146]. Hepcidin, a peptide hormone primarily synthesized and secreted by the liver, plays a crucial role in regulating iron metabolism in the body. It acts as a key regulator of systemic iron homeostasis by controlling iron absorption, recycling, and storage. BMP ligands can regulate the hepcidin levels through BMP receptors and the co-receptor hemojuvelin with the exception of BMP9, which does not interact functionally with hemojuvelin [147]. BMP9 has been shown to upregulate hepcidin expression through SMAD1/5 signaling. In a mouse model, hepatocytes lacking Smad1/5/8 exhibited severe tissue iron loading and liver fibrosis [148]. A meta-analysis of several studies revealed that hepcidin levels were significantly lower in subjects with HCV infection but higher in subjects with NAFLD. No significant differences in hepcidin levels were observed in subjects with HBV infection or controls. Furthermore, patients with cirrhosis of any etiology generally had lower hepcidin levels [149,150,151,152]. However, a cross-sectional study including treatment-naïve patients with chronic hepatitis B (CHB) indicated that reduced hepcidin levels were associated with increased fibrosis severity and histological activity index (HAI) in CHB patients. In autoimmune liver diseases, such as autoimmune hepatitis (AIH) and primary biliary cholangitis/primary sclerosing cholangitis (PBC/PSC), both serum hepcidin levels and the serum hepcidin/ferritin ratio were significantly lower compared to HBV, HCV, or NAFLD cases. These levels also correlated negatively with serum alkaline phosphatase (ALP) levels. PBC/PSC and AIH patients maintained low serum hepcidin throughout their two-year treatment period [153]. These findings suggest that hepcidin may play a protective role in liver fibrosis. Interestingly, the absence of hepcidin expression inhibits hepatic lipid accumulation caused by a high-fat diet, but it can also lead to the early development of fibrosis [154]. Conversely, the overexpression of hepcidin has been shown to alleviate steatohepatitis and fibrosis in a mouse model of diet-induced nonalcoholic steatohepatitis [155]. In summary, hepcidin levels are closely associated with liver fibrosis, but whether they are increased or decreased depends on the specific etiology. Further research is needed to fully understand the complex role of hepcidin in liver fibrosis and its underlying mechanisms.

7.4. BMP9-ID1 Axis in Liver Fibrosis

ID1 is a direct target of BMP9 [156] and plays a crucial role in hepatic fibrogenesis. It is interesting that in HSCs, ID1 could be induced by TGF-β1 through the ALK1–Smad1 pathway but not by the ALK5–Smad2/3 pathway and knockdown of ID1 in HSCs interfered with α-SMA fiber formation [157]. This indicates that ID1 may be a protective factor for liver fibrosis. ID1 has been identified as a potential target for therapeutic interventions of liver fibrosis. Ursodeoxycholic acid (UDCA), for instance, can alleviate liver fibrosis by promoting liver regeneration through the activation of the ID1-Wnt Family Member 2 (WNT2)/hepatocyte growth factor (HGF) pathway. The protective effects of UDCA were diminished when ID1 was knocked down in BDL mice. In a CCl4-induced liver fibrosis model, the deletion of Smad4 in hepatocytes resulted in suppressed ID1 expression and connective tissue growth factor (CTGF) secretion, subsequently activating the p38 and p65 signaling pathways in HSCs and thereby alleviating liver fibrosis [158]. Since ID1 is associated with HCC development, this suggests that an elevated expression of ID1 may play a significant role in the early stages of hepatocarcinogenesis in patients with liver cirrhosis [159]. Therefore, further studies are needed to investigate the role of the BMP9–ID1 axis in liver fibrosis. Besides ID1, TGF-β1 also promotes the activities of several signaling pathways including Smad2/3, TAK1, JNK, p38, MAPKs, RhoA–ROCK, JAK–STAT3, and PI3K–AKT, which contribute to the induction of hepatic fibrosis [160].

8. BMP9 and Portopulmonary Hypertension

Portopulmonary hypertension (PoPH) is a form of PAH that occurs in individuals with portal hypertension and liver cirrhosis and is characterized by increased blood pressure in the portal vein responsible for carrying blood from the digestive organs to the liver. In PoPH, the elevated pressure within the portal system induces changes in the pulmonary blood vessels, resulting in their constriction and reduced blood flow. Consequently, this leads to increased pressure in the pulmonary arteries, ultimately causing PAH [161,162]. Recent studies have highlighted the crucial roles of BMP9 in PoPH [163]. Reduced levels of circulating BMP10 and BMP9, along with elevated levels of endoglin, have been associated with disease severity, decompensation, and pulmonary vascular syndromes in patients with cirrhosis [164]. Circulating levels of BMP9 were found to be decreased in PoPH patients compared to healthy individuals or those with other etiologies of PAH, suggesting BMP9 as a potential biomarker for PoPH [165]. Additionally, studies have shown decreased levels of both BMP9 and BMP10 in plasma samples and liver specimens from patients with decompensated cirrhosis, including those with hepatopulmonary syndrome (HPS) or PoPH. Concurrently, the increased expression of endoglin mRNA was observed in cirrhotic livers, and elevated levels of circulating sEng in PoPH and HPS patients indicated enhanced endothelial sEng shedding in these conditions [161,164,166]. Animal models have demonstrated that a deficiency of BMP9 exacerbates PoPH, while the protective effect of BMP9 in PoPH has been attributed to its interaction with ALK1 and endoglin in pulmonary vascular endothelial cells [165]. These findings suggest that BMP9 is a sensitive and specific biomarker of PoPH, predicting transplant-free survival and the presence of PAH in liver disease.

9. BMP9 and Hepatopulmonary Syndrome

Hepatopulmonary syndrome (HPS) is a condition that affects the liver, lungs, and circulation. It is characterized by the presence of liver disease, abnormalities in arterial oxygenation, and the development of intrapulmonary vascular dilations called pulmonary arteriovenous malformations (PAVMs). Hereditary hemorrhagic telangiectasia is associated with approximately 80% of PAVMs [167]. Considering this, there has been speculation about the association between BMP9 and HPS. Some studies have reported that Krüppel-like factor 6 (KLF6) mediates pulmonary angiogenesis in an experimental rat model of hepatopulmonary syndrome and is exacerbated by BMP9 [168]. Also, research on a rat model showed that elevated circulating BMP9, secreted from the liver, promotes pulmonary angiogenesis in HPS rats via the ALK1–endoglin–Smad1/5/9 pathway [169]. And another study including HPS patients and an animal model showed that portal hypertension-induced loss of BMP9 signaling contributes to HPS development [170]. However, it is worth noting that some researchers believe that the evidence linking BMP9 to HPS is insufficient, and further investigation is required to establish a clear relationship between BMP9 and HPS [125,171]. Therefore, more research is needed to fully elucidate the role of BMP9 in HPS and provide stronger evidence for its involvement.

10. BMP9 and HCC

HCC is the most prevalent form of liver cancer and is associated with a poor prognosis. In recent decades, there has been a consistent increase in the incidence of HCC and the number of deaths caused by this disease. Although BMP9 has been implicated in cancers, its precise role in the development and progression of HCC remains unclear [172]. Previous studies have reported that BMP9 is elevated in approximately 40% of HCC tissues [172,173], while our own previous study showed an even higher percentage at nearly 70% [174]. However, data from the TCGA-LIHC database suggests that the expression level of BMP9 is lower in HCC compared to normal liver tissue. This discrepancy might be attributed to different etiologies and subtypes of liver cancer, which can have varying molecular characteristics and BMP9 expression patterns. Further research is needed to better understand the specific contexts in which BMP9 is involved in HCC and its implications for diagnosis, treatment, and prognosis.
The process of EMT involves a series of steps in which epithelial cells undergo temporary de-differentiation, resulting in a change in their plasticity and the adoption of a mesenchymal phenotype. EMT plays a crucial role in the progression of carcinoma, particularly in cancer invasion and metastasis. During EMT, cells lose their intercellular contacts, primarily due to the downregulation of E-cadherin, leading to increased motility and the ability to spread into nearby or distant tissues. Epithelial plasticity has gained significant attention in HCC. Several families of transcription factors, such as the SNAI family (including Snail (SNAI1) and Slug (SNAI2)), Twist family (Twist1, Twist2, E12, E47, and IDs), and ZEB family (ZEB1 and ZEB2), play important roles in driving the progression of EMT [175]. In HCC, it has been observed that BMP9 expression is associated with increased Smad1 phosphorylation, elevated levels of N-cadherin and Snail, and decreased levels of E-cadherin. Treatment with BMP9 at a concentration from 10 to 50 ng/mL induced cell migration, morphological conversion from epithelial to fibroblastic cells, and the upregulation of the mesenchymal marker Vimentin while reducing the expression of E-cadherin in the HCC cell lines HLE and HepG2 [49,176]. These findings suggest that BMP9 may promote EMT in HCC, contributing to the acquisition of a more invasive and migratory phenotype by HCC cells. In contrast to its promotion of EMT in HCC, BMP9 inhibits EMT in hepatocytes. The opposing effects of BMP9 could be attributed to the distinct cellular characteristics between hepatocytes (non-malignant) and HCC cells (malignant), as well as the differential activation of signaling pathways within each cell type [35].
MAPKs are a family of kinases that play a critical role in various cancer processes [177,178]. Among them, the p38–MAPK pathway is particularly important in regulating cell cycle progression by influencing the transcriptional machinery. The activation of the p38 MAPK pathway has been found to inhibit HCC cell proliferation and promote apoptosis [179,180]. Importantly, the p38–MAPK pathway represents a Smad-independent signaling mechanism that can be activated by BMPs [71,181]. It has been observed that p38 MAPK activation is essential for the protective effect of BMP9 against serum detoxification-induced apoptosis, and its involvement in promoting HepG2 cell growth has also been noted [173]. These findings highlight the significance of the p38–MAPK pathway in mediating the effects of BMP9, indicating its potential role in modulating cell proliferation and apoptosis in HCC cells.
BMP9 has the ability to induce cancer stem cell properties in a specific subtype of liver cancer stem cells (LCSCs) that are characterized by the presence of an epithelial cell adhesion molecule (EpCAM). The Wnt/β-catenin signaling pathway has been extensively associated with various human cancers. Abnormal activation of this pathway is closely linked to increased cancer prevalence, malignant progression, poor prognostics, and even elevated cancer-related mortality rates [182,183]. C-myc, an oncogene, is identified as a downstream target of the Wnt/β-catenin signaling pathway [184,185]. Patients with a high expression of BMP9 in liver cancer tissues tend to have a poorer prognosis. It has been observed that BMP9 can activate the Wnt/β-catenin signaling pathway through ID1, leading to the upregulation of C-myc expression. This activation ultimately induces cancer stem cell properties in cell lines containing EpCAM-positive LCSCs. To counteract these effects, the application of BMP receptor inhibitors, specifically K02288 and LDN-212854, has shown effectiveness in diminishing the cancer stem cell properties induced by BMP9 in EpCAM-positive HCC cells [174]. Another study has also revealed that BMP9 can inhibit m6A methylation modification and promote cell cycle progression in HCC cells. BMP9 is capable of inducing the expression of fat mass and obesity-associated protein (FTO) while reducing the expression of YTH N6-methyladenosine RNA binding protein F2 (YTHDF2). This leads to decreases in global RNA m6A methylation in HCC and m6A methylation within the 5‘-UTR regions of CyclinD1 mRNA. Consequently, the increased expression of CyclinD1 facilitates cell cycle progression in HCC cells [186].
HCC is a highly vascularized solid tumor, and angiogenesis plays a crucial role in its development [187,188]. The HIF-1α/VEGFA signaling pathway is widely implicated in the occurrence, progression, and prognosis of HCC [189,190]. It has been observed that BMP9-ID1 signaling is correlated with HIF-1α expression in HCC tissues. BMP9 can induce the activation of HIF-1α/VEGFA signaling, promoting the secretion of VEGFA, which leads to lumen formation in HUVECs (human umbilical vein endothelial cells). Furthermore, the application of BMP receptor inhibitors can diminish the activation of HIF-1α/VEGFA signaling induced by BMP9, as well as the secretion of VEGFA and lumen formation in HUVECs [191]. Another study conversely demonstrated that BMP9 could induce vascular normalization in hepatitis B virus-associated HCC. In patients with HBV-related HCC, reduced BMP9 expression caused vascular abnormalities that hindered intra-tumoral cytotoxic lymphocyte infiltration. The overexpression of BMP9 in HBV-infected HCC cells facilitated cytotoxic lymphocyte infiltration through vascular normalization. This effect was achieved by inhibiting the Rho–ROCK–myosin light chain (MLC) signaling cascade, ultimately improving the efficacy of immunotherapy. Additionally, ultrasound-targeted microbubble destruction (UTMD)-mediated BMP9 delivery showed therapeutic efficacy when combined with a programmed death-ligand 1 (PD-L1) antibody in HepG2.2.15 cell xenografts in immune-deficient mice [192].
These findings suggest that the influence of BMP9 signaling on HCC is complicated. Further research is needed to elucidate the mechanisms through which BMP9 signaling regulates HCC derived from different etiologies.

11. Summary and Future Expectations

BMP9 is a multifunctional cytokine that plays a crucial role in various cellular processes, including bone and cartilage formation, glucose and lipid metabolisms, iron balance, neuron differentiation, angiogenesis, and lymphangiogenesis. The liver serves as the primary organ for BMP9 expression and secretion, indicating its potential involvement in the progression of liver diseases [193]. In this review, we have summarized the functions and mechanisms of BMP9 in diverse liver diseases, including viral hepatitis, acute liver injury, NAFLD, hepatic fibrosis, PoPH, HPS, and HCC (Figure 5). It has been observed that BMP9 is downregulated in cases of HCV infection and acute liver injury. Moreover, BMP9 has shown beneficial effects on two risk factors for NAFLD and obesity. Its effects include regulating glucose and lipid metabolisms by inhibiting liver gluconeogenesis, promoting the transformation of WAT to BAT, enhancing insulin sensitivity, and inhibiting liver lipid deposition. Regarding liver fibrosis, BMP9 has the potential to induce its progression through the activation of the ALK1–Smad1/5 signaling pathway and associated targets such as ID1, Hepcidin and Snail. However, the co-receptor endoglin exerts a protective effect against liver fibrosis through the ALK5–Smad2/3 signaling pathway. In the case of PoPH, decreased BMP9 levels suggest a possible protective effect in this condition. The role of BMP9 in HPS remains unclear, and further research is required to understand its specific involvement. In HCC, BMP9 has been found to induce the EMT process and upregulate the expression of EMT markers such as Vimentin while reducing E-cadherin in HCC cell lines. Additionally, BMP9 promotes HCC cell proliferation through the activation of both the Smad1/5–ID1 signaling pathway and the p38–MAPK signaling pathway. Moreover, BMP9 enhances the cancer stem cell properties in Huh7 and MT cell lines characterized by EpCAM positive LCSCs. Furthermore, BMP9 induces angiogenesis by activating the HIF-1α/VEGFA signaling pathway in the Huh7 and MT cell lines. In HBV-related HCC, it has been observed that HBV infection reduces the expression of BMP9. However, BMP9 treatment can induce vascular normalization and facilitate the infiltration of cytotoxic lymphocytes through the Rho–ROCK–MLC signaling pathway. These findings suggest potential therapeutic efficacy when combining BMP9 with a PD-L1 antibody for treating HBV-related HCC.
In addition to BMP9, ALK1 is also a promising therapeutic target for the treatment of HCC. Currently, there are three drugs under development targeting ALK1. They are the monoclonal antibody Ascrinvacumab, the bispecific antibody ALK1/VEGF, and the monoclonal antibody TATX-21. Previously, there was also a fusion protein called Dalantercept, but it showed limited efficacy in HCC [194]. These three investigational drugs are at different stages of development. Ascrinvacumab has entered clinical trials, while the other two drugs are still in early stage research. These drugs have different therapeutic areas. Ascrinvacumab mainly targets solid tumors, including HCC [195,196], and TATX-21 could be used for the treatment of arterial sclerosis and heart disease. In summary, as a novel target for solid tumors and cardiovascular diseases, ALK1 holds promising potential for breakthroughs.
In conclusion, BMP9 plays a dual role depending on specific conditions and diseases. A thorough comprehension of BMP9 will significantly enhance our understanding of viral hepatitis, acute liver injury, obesity, T2DM, NAFLD, liver fibrosis or cirrhosis, PoPH, and HCC, potentially establishing it as a valuable biomarker for diagnosing and monitoring various liver diseases. However, more evidence is needed to support the therapeutic potential of the BMP9-ALK1 axis for HCC treatment.

Author Contributions

Conceptualization, H.C. and H.T.; validation, H.C., Y.-Y.L. and K.N.; data curation, H.C.; writing—original draft preparation, H.C.; writing—review and editing, K.N., Y.-Y.L. and H.T.; visualization, H.C.; supervision, H.T.; project administration, H.T.; funding acquisition, H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (NO.2022YFC2304800) and 1.3.5 project for disciplines of excellence, West China Hospital, Sichuan University (ZYGD23030).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

α-SMAalpha smooth muscle actin
APAPacetaminophen
ALIacute liver injury
ALPalkaline phosphatase
AIHautoimmune hepatitis
AUCthe area under the curve for glucose
BATbrown adipose tissue
BMIbody mass index
BMPsbone morphogenetic proteins
CBPCREB-binding protein
CCl4carbon tetrachloride
C/EBPαCCAAT/enhancer binding protein (C/EBP) alpha
CHBchronic hepatitis B
CREBcAMP response-element binding protein
CTGFconnective tissue growth factor
DDC3,5-diethoxicarbonyl-1,4-dihydrocollidine
5-diethoxicarbonyl-14-dihydrocollidine
ECMextracellular matrix
EGFepidermal growth factor
EMTepithelial-to-mesenchymal transition
ENGendoglin
EpCAMepithelial cell adhesion molecule
ERKextracellular signal-regulated kinase
FBGfasting blood glucose
FFAfree fatty acid
FFChigh fat, fructose, and cholesterol diet
FGFfibroblast growth factor
FINSfasting insulin
FTOfat mass and obesity-associated protein
GDFgrowth differentiation factor
GSK3glycogen synthase kinase 3
HAIhistological activity index
HCChepatocellular carcinoma
HCVhepatitis C virus
HDL-Chigh-density lipoprotein cholesterol
HGFhepatocyte growth factor
HHThereditary hemorrhagic telangiectasia
HIF-1αhypoxia inducible factor 1 subunit alpha
HOMA-IRhomoeostasis model assessment of insulin resistance
HPShepatopulmonary syndrome
HSChepatic stellate cell
HUVEChuman umbilical vein endothelial cell
ID1/2/3inhibitor of differentiation-1/2/3
IFNinterferon
IL-8interleukin-8
InsRinsulin receptor
IRinsulin resistance
IRFinterferon regulatory factors
JNKc-Jun N-terminal kinases
KCKupffer cells
KLF6Krüppel-like factor 6
LDL-Clow-density lipoprotein cholesterol
LIMKLIM Domain Kinase
LPSlipopolysaccharide
LSECliver sinusoidal endothelial cell
LCSCliver cancer stem cell
MAPKmitogen-activated protein kinases
MB109recombinant derivative of human BMP9
MLCmyosin light chain
MMP-7matrix metalloproteinase-7
NAFLDnonalcoholic fatty liver disease
NASHnonalcoholic steatohepatitis
OGTToral glucose tolerance tests
PAHpulmonary arterial hypertension
PAVMpulmonary arteriovenous malformations
PBC/PSCprimary biliary cholangitis/primary sclerosing cholangitis
PD-L1programmed death-ligand 1
PEPCKphosphoenolpyruvate carboxykinase
PGC1αproliferator-activated receptor-gamma coactivator 1 alpha
PI3KPhosphoinositide 3-kinases
PoPHportopulmonary hypertension
PPARαperoxisome proliferator-activated receptor α
ROCKRho-Associated Coiled-Coil-Containing Protein Kinase
TAGtriglyceride
TGF-βtransforming growth factor β
T2DMtype 2 diabetes mellitus
UCP1uncoupling protein 1
UDCAUrsodesoxycholic acid
USP18Ubiquitin specific peptidase 18
UTMDultrasound-targeted microbubble destruction
VEGFvascular endothelial growth factor
WATwhite adipose tissue
WHRWaist-hip ratio
WNT2Wnt Family Member 2
YTHDF2YTH N6-methyladenosine RNA binding protein F2
ZEBzinc finger E-box binding

References

  1. Chen, D.; Zhao, M.; Mundy, G.R. Bone morphogenetic proteins. Growth Factors 2004, 22, 233–241. [Google Scholar] [CrossRef]
  2. Elima, K. Osteoinductive proteins. Ann. Med. 1993, 25, 395–402. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, Y.; Ma, B.; Wang, X.; Zha, X.; Sheng, C.; Yang, P.; Qu, S. Potential Functions of the BMP Family in Bone, Obesity, and Glucose Metabolism. J. Diabetes Res. 2021, 2021, 6707464. [Google Scholar] [CrossRef] [PubMed]
  4. Sartori, R.; Sandri, M. BMPs and the muscle-bone connection. Bone 2015, 80, 37–42. [Google Scholar] [CrossRef]
  5. Liu, W.; Deng, Z.; Zeng, Z.; Fan, J.; Feng, Y.; Wang, X.; Cao, D.; Zhang, B.; Yang, L.; Liu, B.; et al. Highly expressed BMP9/GDF2 in postnatal mouse liver and lungs may account for its pleiotropic effects on stem cell differentiation, angiogenesis, tumor growth and metabolism. Genes Dis. 2020, 7, 235–244. [Google Scholar] [CrossRef]
  6. Chen, W.; Ten Dijke, P. Immunoregulation by members of the TGFbeta superfamily. Nat. Rev. Immunol. 2016, 16, 723–740. [Google Scholar] [CrossRef] [PubMed]
  7. Wu, J.; Jackson-Weaver, O.; Xu, J. The TGFbeta superfamily in cardiac dysfunction. Acta Biochim. Biophys. Sin. 2018, 50, 323–335. [Google Scholar] [CrossRef] [PubMed]
  8. Lui, P.P. Histopathological changes in tendinopathy--potential roles of BMPs? Rheumatology 2013, 52, 2116–2126. [Google Scholar] [CrossRef]
  9. Liu, M.; Goldman, G.; MacDougall, M.; Chen, S. BMP Signaling Pathway in Dentin Development and Diseases. Cells 2022, 11, 2216. [Google Scholar] [CrossRef]
  10. Katagiri, T.; Watabe, T. Bone Morphogenetic Proteins. Cold Spring Harb. Perspect. Biol. 2016, 8. [Google Scholar] [CrossRef]
  11. Lin, S.; Svoboda, K.K.; Feng, J.Q.; Jiang, X. The biological function of type I receptors of bone morphogenetic protein in bone. Bone Res. 2016, 4, 16005. [Google Scholar] [CrossRef] [PubMed]
  12. Sanchez-Duffhues, G.; Williams, E.; Goumans, M.J.; Heldin, C.H.; Ten Dijke, P. Bone morphogenetic protein receptors: Structure, function and targeting by selective small molecule kinase inhibitors. Bone 2020, 138, 115472. [Google Scholar] [CrossRef] [PubMed]
  13. Song, J.J.; Celeste, A.J.; Kong, F.M.; Jirtle, R.L.; Rosen, V.; Thies, R.S. Bone morphogenetic protein-9 binds to liver cells and stimulates proliferation. Endocrinology 1995, 136, 4293–4297. [Google Scholar] [CrossRef] [PubMed]
  14. Ploemacher, R.E.; Engels, L.J.; Mayer, A.E.; Thies, S.; Neben, S. Bone morphogenetic protein 9 is a potent synergistic factor for murine hemopoietic progenitor cell generation and colony formation in serum-free cultures. Leukemia 1999, 13, 428–437. [Google Scholar] [CrossRef] [PubMed]
  15. Lamplot, J.D.; Qin, J.; Nan, G.; Wang, J.; Liu, X.; Yin, L.; Tomal, J.; Li, R.; Shui, W.; Zhang, H.; et al. BMP9 signaling in stem cell differentiation and osteogenesis. Am. J. Stem Cells 2013, 2, 1–21. [Google Scholar]
  16. Ruschke, K.; Hiepen, C.; Becker, J.; Knaus, P. BMPs are mediators in tissue crosstalk of the regenerating musculoskeletal system. Cell Tissue Res. 2012, 347, 521–544. [Google Scholar] [CrossRef] [PubMed]
  17. Tsumaki, N.; Yoshikawa, H. The role of bone morphogenetic proteins in endochondral bone formation. Cytokine Growth Factor. Rev. 2005, 16, 279–285. [Google Scholar] [CrossRef] [PubMed]
  18. Thielen, N.G.M.; van der Kraan, P.M.; van Caam, A.P.M. TGFbeta/BMP Signaling Pathway in Cartilage Homeostasis. Cells 2019, 8, 969. [Google Scholar] [CrossRef] [PubMed]
  19. Yang, M.; Liang, Z.; Yang, M.; Jia, Y.; Yang, G.; He, Y.; Li, X.; Gu, H.F.; Zheng, H.; Zhu, Z.; et al. Role of bone morphogenetic protein-9 in the regulation of glucose and lipid metabolism. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2019, 33, 10077–10088. [Google Scholar] [CrossRef]
  20. Truksa, J.; Peng, H.; Lee, P.; Beutler, E. Different regulatory elements are required for response of hepcidin to interleukin-6 and bone morphogenetic proteins 4 and 9. Br. J. Haematol. 2007, 139, 138–147. [Google Scholar] [CrossRef]
  21. Schnitzler, A.C.; Mellott, T.J.; Lopez-Coviella, I.; Tallini, Y.N.; Kotlikoff, M.I.; Follettie, M.T.; Blusztajn, J.K. BMP9 (bone morphogenetic protein 9) induces NGF as an autocrine/paracrine cholinergic trophic factor in developing basal forebrain neurons. J. Neurosci. Off. J. Soc. Neurosci. 2010, 30, 8221–8228. [Google Scholar] [CrossRef]
  22. Lopez-Coviella, I.; Follettie, M.T.; Mellott, T.J.; Kovacheva, V.P.; Slack, B.E.; Diesl, V.; Berse, B.; Thies, R.S.; Blusztajn, J.K. Bone morphogenetic protein 9 induces the transcriptome of basal forebrain cholinergic neurons. Proc. Natl. Acad. Sci. USA 2005, 102, 6984–6989. [Google Scholar] [CrossRef] [PubMed]
  23. Ahmed, T.; Ramonett, A.; Kwak, E.A.; Kumar, S.; Flores, P.C.; Ortiz, H.R.; Langlais, P.R.; Hund, T.J.; Mythreye, K.; Lee, N.Y. Endothelial tip/stalk cell selection requires BMP9-induced beta(IV)-spectrin expression during sprouting angiogenesis. Mol. Biol. Cell 2023, 34, ar72. [Google Scholar] [CrossRef] [PubMed]
  24. Ayuso-Inigo, B.; Mendez-Garcia, L.; Pericacho, M.; Munoz-Felix, J.M. The Dual Effect of the BMP9-ALK1 Pathway in Blood Vessels: An Opportunity for Cancer Therapy Improvement? Cancers 2021, 13, 5412. [Google Scholar] [CrossRef]
  25. Xiao, H.; Wang, X.; Wang, C.; Dai, G.; Zhu, Z.; Gao, S.; He, B.; Liao, J.; Huang, W. BMP9 exhibits dual and coupled roles in inducing osteogenic and angiogenic differentiation of mesenchymal stem cells. Biosci. Rep. 2020, 40, BSR20201262. [Google Scholar] [CrossRef]
  26. Ricard, N.; Ciais, D.; Levet, S.; Subileau, M.; Mallet, C.; Zimmers, T.A.; Lee, S.J.; Bidart, M.; Feige, J.J.; Bailly, S. BMP9 and BMP10 are critical for postnatal retinal vascular remodeling. Blood 2012, 119, 6162–6171. [Google Scholar] [CrossRef] [PubMed]
  27. Levet, S.; Ciais, D.; Merdzhanova, G.; Mallet, C.; Zimmers, T.A.; Lee, S.J.; Navarro, F.P.; Texier, I.; Feige, J.J.; Bailly, S.; et al. Bone morphogenetic protein 9 (BMP9) controls lymphatic vessel maturation and valve formation. Blood 2013, 122, 598–607. [Google Scholar] [CrossRef] [PubMed]
  28. Breitkopf-Heinlein, K.; Meyer, C.; Konig, C.; Gaitantzi, H.; Addante, A.; Thomas, M.; Wiercinska, E.; Cai, C.; Li, Q.; Wan, F.; et al. BMP-9 interferes with liver regeneration and promotes liver fibrosis. Gut 2017, 66, 939–954. [Google Scholar] [CrossRef] [PubMed]
  29. Sosa, I.; Cvijanovic, O.; Celic, T.; Cuculic, D.; Crncevic-Orlic, Z.; Vukelic, L.; Zoricic Cvek, S.; Dudaric, L.; Bosnar, A.; Bobinac, D. Hepatoregenerative role of bone morphogenetic protein-9. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2011, 17, HY33–HY35. [Google Scholar] [CrossRef]
  30. Tillet, E.; Bailly, S. Emerging roles of BMP9 and BMP10 in hereditary hemorrhagic telangiectasia. Front. Genet. 2014, 5, 456. [Google Scholar] [CrossRef]
  31. Desroches-Castan, A.; Tillet, E.; Bouvard, C.; Bailly, S. BMP9 and BMP10: Two close vascular quiescence partners that stand out. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2022, 251, 178–197. [Google Scholar] [CrossRef]
  32. Medina-Jover, F.; Riera-Mestre, A.; Vinals, F. Rethinking growth factors: The case of BMP9 during vessel maturation. Vasc. Biol. 2022, 4, R1–R14. [Google Scholar] [CrossRef]
  33. Li, W.; Salmon, R.M.; Jiang, H.; Morrell, N.W. Regulation of the ALK1 ligands, BMP9 and BMP10. Biochem. Soc. Trans. 2016, 44, 1135–1141. [Google Scholar] [CrossRef]
  34. Robert, F.; Berrebeh, N.; Guignabert, C.; Humbert, M.; Bailly, S.; Tu, L.; Savale, L. Dysfunction of endothelial BMP-9 signaling in pulmonary vascular disease. Rev. Mal. Respir. 2023, 40, 234–238. [Google Scholar] [CrossRef]
  35. Herrera, B.; Dooley, S.; Breitkopf-Heinlein, K. Potential roles of bone morphogenetic protein (BMP)-9 in human liver diseases. Int. J. Mol. Sci. 2014, 15, 5199. [Google Scholar] [CrossRef] [PubMed]
  36. Herrera, B.; Sanchez, A.; Fabregat, I. BMPS and liver: More questions than answers. Curr. Pharm. Des. 2012, 18, 4114–4125. [Google Scholar] [CrossRef]
  37. Mi, L.Z.; Brown, C.T.; Gao, Y.; Tian, Y.; Le, V.Q.; Walz, T.; Springer, T.A. Structure of bone morphogenetic protein 9 procomplex. Proc. Natl. Acad. Sci. USA 2015, 112, 3710–3715. [Google Scholar] [CrossRef]
  38. Brown, M.A.; Zhao, Q.; Baker, K.A.; Naik, C.; Chen, C.; Pukac, L.; Singh, M.; Tsareva, T.; Parice, Y.; Mahoney, A.; et al. Crystal structure of BMP-9 and functional interactions with pro-region and receptors. J. Biol. Chem. 2005, 280, 25111–25118. [Google Scholar] [CrossRef] [PubMed]
  39. Wei, Z.; Salmon, R.M.; Upton, P.D.; Morrell, N.W.; Li, W. Regulation of bone morphogenetic protein 9 (BMP9) by redox-dependent proteolysis. J. Biol. Chem. 2014, 289, 31150–31159. [Google Scholar] [CrossRef]
  40. Wang, Y.; Ma, C.; Sun, T.; Ren, L. Potential roles of bone morphogenetic protein-9 in glucose and lipid homeostasis. J. Physiol. Biochem. 2020, 76, 503–512. [Google Scholar] [CrossRef] [PubMed]
  41. Bidart, M.; Ricard, N.; Levet, S.; Samson, M.; Mallet, C.; David, L.; Subileau, M.; Tillet, E.; Feige, J.J.; Bailly, S. BMP9 is produced by hepatocytes and circulates mainly in an active mature form complexed to its prodomain. Cell. Mol. Life Sci. CMLS 2012, 69, 313–324. [Google Scholar] [CrossRef] [PubMed]
  42. Tillet, E.; Ouarne, M.; Desroches-Castan, A.; Mallet, C.; Subileau, M.; Didier, R.; Lioutsko, A.; Belthier, G.; Feige, J.J.; Bailly, S. A heterodimer formed by bone morphogenetic protein 9 (BMP9) and BMP10 provides most BMP biological activity in plasma. J. Biol. Chem. 2018, 293, 10963–10974. [Google Scholar] [CrossRef]
  43. Kienast, Y.; Jucknischke, U.; Scheiblich, S.; Thier, M.; de Wouters, M.; Haas, A.; Lehmann, C.; Brand, V.; Bernicke, D.; Honold, K.; et al. Rapid Activation of Bone Morphogenic Protein 9 by Receptor-mediated Displacement of Pro-domains. J. Biol. Chem. 2016, 291, 3395–3410. [Google Scholar] [CrossRef]
  44. Yadin, D.; Knaus, P.; Mueller, T.D. Structural insights into BMP receptors: Specificity, activation and inhibition. Cytokine Growth Factor. Rev. 2016, 27, 13–34. [Google Scholar] [CrossRef]
  45. Salmon, R.M.; Guo, J.; Wood, J.H.; Tong, Z.; Beech, J.S.; Lawera, A.; Yu, M.; Grainger, D.J.; Reckless, J.; Morrell, N.W.; et al. Molecular basis of ALK1-mediated signalling by BMP9/BMP10 and their prodomain-bound forms. Nat. Commun. 2020, 11, 1621. [Google Scholar] [CrossRef]
  46. Scharpfenecker, M.; van Dinther, M.; Liu, Z.; van Bezooijen, R.L.; Zhao, Q.; Pukac, L.; Lowik, C.W.; ten Dijke, P. BMP-9 signals via ALK1 and inhibits bFGF-induced endothelial cell proliferation and VEGF-stimulated angiogenesis. J. Cell Sci. 2007, 120, 964–972. [Google Scholar] [CrossRef] [PubMed]
  47. Herrera, B.; van Dinther, M.; Ten Dijke, P.; Inman, G.J. Autocrine bone morphogenetic protein-9 signals through activin receptor-like kinase-2/Smad1/Smad4 to promote ovarian cancer cell proliferation. Cancer Res. 2009, 69, 9254–9262. [Google Scholar] [CrossRef]
  48. Xia, Y.; Babitt, J.L.; Sidis, Y.; Chung, R.T.; Lin, H.Y. Hemojuvelin regulates hepcidin expression via a selective subset of BMP ligands and receptors independently of neogenin. Blood 2008, 111, 5195–5204. [Google Scholar] [CrossRef]
  49. Li, Q.; Gu, X.; Weng, H.; Ghafoory, S.; Liu, Y.; Feng, T.; Dzieran, J.; Li, L.; Ilkavets, I.; Kruithof-de Julio, M.; et al. Bone morphogenetic protein-9 induces epithelial to mesenchymal transition in hepatocellular carcinoma cells. Cancer Sci. 2013, 104, 398–408. [Google Scholar] [CrossRef] [PubMed]
  50. Saito, T.; Bokhove, M.; Croci, R.; Zamora-Caballero, S.; Han, L.; Letarte, M.; de Sanctis, D.; Jovine, L. Structural Basis of the Human Endoglin-BMP9 Interaction: Insights into BMP Signaling and HHT1. Cell Rep. 2017, 19, 1917–1928. [Google Scholar] [CrossRef]
  51. Castonguay, R.; Werner, E.D.; Matthews, R.G.; Presman, E.; Mulivor, A.W.; Solban, N.; Sako, D.; Pearsall, R.S.; Underwood, K.W.; Seehra, J.; et al. Soluble endoglin specifically binds bone morphogenetic proteins 9 and 10 via its orphan domain, inhibits blood vessel formation, and suppresses tumor growth. J. Biol. Chem. 2011, 286, 30034–30046. [Google Scholar] [CrossRef]
  52. Lawera, A.; Tong, Z.; Thorikay, M.; Redgrave, R.E.; Cai, J.; van Dinther, M.; Morrell, N.W.; Afink, G.B.; Charnock-Jones, D.S.; Arthur, H.M.; et al. Role of soluble endoglin in BMP9 signaling. Proc. Natl. Acad. Sci. USA 2019, 116, 17800–17808. [Google Scholar] [CrossRef] [PubMed]
  53. David, L.; Mallet, C.; Mazerbourg, S.; Feige, J.J.; Bailly, S. Identification of BMP9 and BMP10 as functional activators of the orphan activin receptor-like kinase 1 (ALK1) in endothelial cells. Blood 2007, 109, 1953–1961. [Google Scholar] [CrossRef] [PubMed]
  54. Mallet, C.; Lamribet, K.; Giraud, S.; Dupuis-Girod, S.; Feige, J.J.; Bailly, S.; Tillet, E. Functional analysis of endoglin mutations from hereditary hemorrhagic telangiectasia type 1 patients reveals different mechanisms for endoglin loss of function. Hum. Mol. Genet. 2015, 24, 1142–1154. [Google Scholar] [CrossRef] [PubMed]
  55. Nolan-Stevaux, O.; Zhong, W.; Culp, S.; Shaffer, K.; Hoover, J.; Wickramasinghe, D.; Ruefli-Brasse, A. Endoglin requirement for BMP9 signaling in endothelial cells reveals new mechanism of action for selective anti-endoglin antibodies. PLoS ONE 2012, 7, e50920. [Google Scholar] [CrossRef] [PubMed]
  56. Deng, Z.; Fan, T.; Xiao, C.; Tian, H.; Zheng, Y.; Li, C.; He, J. TGF-beta signaling in health, disease, and therapeutics. Signal Transduct. Target. Ther. 2024, 9, 61. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, S.J.; Yuan, W.; Mori, Y.; Levenson, A.; Trojanowska, M.; Varga, J. Stimulation of type I collagen transcription in human skin fibroblasts by TGF-beta: Involvement of Smad 3. J. Invest. Dermatol. 1999, 112, 49–57. [Google Scholar] [CrossRef] [PubMed]
  58. Dennler, S.; Itoh, S.; Vivien, D.; ten Dijke, P.; Huet, S.; Gauthier, J.M. Direct binding of Smad3 and Smad4 to critical TGF beta-inducible elements in the promoter of human plasminogen activator inhibitor-type 1 gene. EMBO J. 1998, 17, 3091–3100. [Google Scholar] [CrossRef] [PubMed]
  59. Yuan, W.; Varga, J. Transforming growth factor-beta repression of matrix metalloproteinase-1 in dermal fibroblasts involves Smad3. J. Biol. Chem. 2001, 276, 38502–38510. [Google Scholar] [CrossRef]
  60. Piek, E.; Ju, W.J.; Heyer, J.; Escalante-Alcalde, D.; Stewart, C.L.; Weinstein, M.; Deng, C.; Kucherlapati, R.; Bottinger, E.P.; Roberts, A.B. Functional characterization of transforming growth factor beta signaling in Smad2- and Smad3-deficient fibroblasts. J. Biol. Chem. 2001, 276, 19945–19953. [Google Scholar] [CrossRef]
  61. Zeisberg, M.; Hanai, J.; Sugimoto, H.; Mammoto, T.; Charytan, D.; Strutz, F.; Kalluri, R. BMP-7 counteracts TGF-beta1-induced epithelial-to-mesenchymal transition and reverses chronic renal injury. Nat. Med. 2003, 9, 964–968. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, S.; Hirschberg, R. Bone morphogenetic protein-7 signals opposing transforming growth factor beta in mesangial cells. J. Biol. Chem. 2004, 279, 23200–23206. [Google Scholar] [CrossRef]
  63. Meng, X.M.; Nikolic-Paterson, D.J.; Lan, H.Y. TGF-beta: The master regulator of fibrosis. Nat. Rev. Nephrol. 2016, 12, 325–338. [Google Scholar] [CrossRef]
  64. Munoz-Felix, J.M.; Gonzalez-Nunez, M.; Lopez-Novoa, J.M. ALK1-Smad1/5 signaling pathway in fibrosis development: Friend or foe? Cytokine Growth Factor. Rev. 2013, 24, 523–537. [Google Scholar] [CrossRef] [PubMed]
  65. Luyckx, I.; Verstraeten, A.; Goumans, M.J.; Loeys, B. SMAD6-deficiency in human genetic disorders. NPJ Genom. Med. 2022, 7, 68. [Google Scholar] [CrossRef]
  66. de Ceuninck van Capelle, C.; Spit, M.; Ten Dijke, P. Current perspectives on inhibitory SMAD7 in health and disease. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 691–715. [Google Scholar] [CrossRef]
  67. Theilmann, A.L.; Hawke, L.G.; Hilton, L.R.; Whitford, M.K.M.; Cole, D.V.; Mackeil, J.L.; Dunham-Snary, K.J.; Mewburn, J.; James, P.D.; Maurice, D.H.; et al. Endothelial BMPR2 Loss Drives a Proliferative Response to BMP (Bone Morphogenetic Protein) 9 via Prolonged Canonical Signaling. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 2605–2618. [Google Scholar] [CrossRef]
  68. Tao, B.; Kraehling, J.R.; Ghaffari, S.; Ramirez, C.M.; Lee, S.; Fowler, J.W.; Lee, W.L.; Fernandez-Hernando, C.; Eichmann, A.; Sessa, W.C. BMP-9 and LDL crosstalk regulates ALK-1 endocytosis and LDL transcytosis in endothelial cells. J. Biol. Chem. 2020, 295, 18179–18188. [Google Scholar] [CrossRef]
  69. Upton, P.D.; Davies, R.J.; Trembath, R.C.; Morrell, N.W. Bone morphogenetic protein (BMP) and activin type II receptors balance BMP9 signals mediated by activin receptor-like kinase-1 in human pulmonary artery endothelial cells. J. Biol. Chem. 2009, 284, 15794–15804. [Google Scholar] [CrossRef] [PubMed]
  70. van Caam, A.; Blaney Davidson, E.; Garcia de Vinuesa, A.; van Geffen, E.; van den Berg, W.; Goumans, M.J.; ten Dijke, P.; van der Kraan, P. The high affinity ALK1-ligand BMP9 induces a hypertrophy-like state in chondrocytes that is antagonized by TGFbeta1. Osteoarthr. Cartil. 2015, 23, 985–995. [Google Scholar] [CrossRef] [PubMed]
  71. Wu, M.; Chen, G.; Li, Y.P. TGF-beta and BMP signaling in osteoblast, skeletal development, and bone formation, homeostasis and disease. Bone Res. 2016, 4, 16009. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, G.; Deng, C.; Li, Y.P. TGF-beta and BMP signaling in osteoblast differentiation and bone formation. Int. J. Biol. Sci. 2012, 8, 272–288. [Google Scholar] [CrossRef] [PubMed]
  73. Zhong, J.; Zou, H. BMP signaling in axon regeneration. Curr. Opin. Neurobiol. 2014, 27, 127–134. [Google Scholar] [CrossRef] [PubMed]
  74. Jin, Y.; Kaluza, D.; Jakobsson, L. VEGF, Notch and TGFbeta/BMPs in regulation of sprouting angiogenesis and vascular patterning. Biochem. Soc. Trans. 2014, 42, 1576–1583. [Google Scholar] [CrossRef] [PubMed]
  75. Zhao, X. BMP9 promotes tibial fracture healing of rats through regulating notch signaling pathway. Minerva Medica 2021, 112, 828–829. [Google Scholar] [CrossRef] [PubMed]
  76. Young, K.; Tweedie, E.; Conley, B.; Ames, J.; FitzSimons, M.; Brooks, P.; Liaw, L.; Vary, C.P. BMP9 Crosstalk with the Hippo Pathway Regulates Endothelial Cell Matricellular and Chemokine Responses. PLoS ONE 2015, 10, e0122892. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, X.; Qin, J.; Luo, Q.; Bi, Y.; Zhu, G.; Jiang, W.; Kim, S.H.; Li, M.; Su, Y.; Nan, G.; et al. Cross-talk between EGF and BMP9 signalling pathways regulates the osteogenic differentiation of mesenchymal stem cells. J. Cell. Mol. Med. 2013, 17, 1160–1172. [Google Scholar] [CrossRef]
  78. Peng, Y.; Kang, Q.; Luo, Q.; Jiang, W.; Si, W.; Liu, B.A.; Luu, H.H.; Park, J.K.; Li, X.; Luo, J.; et al. Inhibitor of DNA binding/differentiation helix-loop-helix proteins mediate bone morphogenetic protein-induced osteoblast differentiation of mesenchymal stem cells. J. Biol. Chem. 2004, 279, 32941–32949. [Google Scholar] [CrossRef] [PubMed]
  79. Yang, J.; Li, X.; Li, Y.; Southwood, M.; Ye, L.; Long, L.; Al-Lamki, R.S.; Morrell, N.W. Id proteins are critical downstream effectors of BMP signaling in human pulmonary arterial smooth muscle cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2013, 305, L312–L321. [Google Scholar] [CrossRef]
  80. Truksa, J.; Peng, H.; Lee, P.; Beutler, E. Bone morphogenetic proteins 2, 4, and 9 stimulate murine hepcidin 1 expression independently of Hfe, transferrin receptor 2 (Tfr2), and IL-6. Proc. Natl. Acad. Sci. USA 2006, 103, 10289–10293. [Google Scholar] [CrossRef]
  81. Bi, J.; Ge, S. Potential roles of BMP9 in liver fibrosis. Int. J. Mol. Sci. 2014, 15, 20656. [Google Scholar] [CrossRef]
  82. Ma, J.; van der Zon, G.; Goncalves, M.; van Dinther, M.; Thorikay, M.; Sanchez-Duffhues, G.; Ten Dijke, P. TGF-beta-Induced Endothelial to Mesenchymal Transition Is Determined by a Balance Between SNAIL and ID Factors. Front. Cell Dev. Biol. 2021, 9, 616610. [Google Scholar] [CrossRef]
  83. Liu, R.; Xu, W.; Zhu, H.; Dong, Z.; Dong, H.; Yin, S. Aging aggravates acetaminophen-induced acute liver injury and inflammation through inordinate C/EBPalpha-BMP9 crosstalk. Cell Biosci. 2023, 13, 61. [Google Scholar] [CrossRef] [PubMed]
  84. Mo, L.; Jiang, H.B.; Tian, G.R.; Lu, G.J. The proliferation and migration of atherosclerosis-related HVSMCs were inhibited by downregulation of lncRNA XIST via regulation of the miR-761/BMP9 axis. Kaohsiung J. Med. Sci. 2022, 38, 18–29. [Google Scholar] [CrossRef]
  85. Yao, Y.; Jumabay, M.; Ly, A.; Radparvar, M.; Wang, A.H.; Abdmaulen, R.; Bostrom, K.I. Crossveinless 2 regulates bone morphogenetic protein 9 in human and mouse vascular endothelium. Blood 2012, 119, 5037–5047. [Google Scholar] [CrossRef]
  86. Gaitantzi, H.; Karch, J.; Germann, L.; Cai, C.; Rausch, V.; Birgin, E.; Rahbari, N.; Seitz, T.; Hellerbrand, C.; Konig, C.; et al. BMP-9 Modulates the Hepatic Responses to LPS. Cells 2020, 9, 617. [Google Scholar] [CrossRef]
  87. Miller, A.F.; Harvey, S.A.; Thies, R.S.; Olson, M.S. Bone morphogenetic protein-9. An autocrine/paracrine cytokine in the liver. J. Biol. Chem. 2000, 275, 17937–17945. [Google Scholar] [CrossRef]
  88. Tao, L.; Fang, S.Y.; Zhao, L.; He, T.C.; He, Y.; Bi, Y. Indocyanine Green Uptake and Periodic Acid-Schiff Staining Method for Function Detection of Liver Cells are Affected by Different Cell Confluence. Cytotechnology 2021, 73, 159–167. [Google Scholar] [CrossRef]
  89. Addante, A.; Gonzalez-Corralejo, C.; Roncero, C.; Lazcanoiturburu, N.; Garcia-Saez, J.; Herrera, B.; Sanchez, A. BMP9 Promotes an Epithelial Phenotype and a Hepatocyte-like Gene Expression Profile in Adult Hepatic Progenitor Cells. Cells 2022, 11, 365. [Google Scholar] [CrossRef]
  90. Addante, A.; Roncero, C.; Almale, L.; Lazcanoiturburu, N.; Garcia-Alvaro, M.; Fernandez, M.; Sanz, J.; Hammad, S.; Nwosu, Z.C.; Lee, S.J.; et al. Bone morphogenetic protein 9 as a key regulator of liver progenitor cells in DDC-induced cholestatic liver injury. Liver Int. Off. J. Int. Assoc. Study Liver 2018, 38, 1664–1675. [Google Scholar] [CrossRef]
  91. Guilliams, M.; Bonnardel, J.; Haest, B.; Vanderborght, B.; Wagner, C.; Remmerie, A.; Bujko, A.; Martens, L.; Thone, T.; Browaeys, R.; et al. Spatial proteogenomics reveals distinct and evolutionarily conserved hepatic macrophage niches. Cell 2022, 185, 379–396.e338. [Google Scholar] [CrossRef] [PubMed]
  92. Zhao, D.; Yang, F.; Wang, Y.; Li, S.; Li, Y.; Hou, F.; Yang, W.; Liu, D.; Tao, Y.; Li, Q.; et al. ALK1 signaling is required for the homeostasis of Kupffer cells and prevention of bacterial infection. J. Clin. Investig. 2022, 132, e150489. [Google Scholar] [CrossRef]
  93. Desroches-Castan, A.; Tillet, E.; Ricard, N.; Ouarne, M.; Mallet, C.; Belmudes, L.; Coute, Y.; Boillot, O.; Scoazec, J.Y.; Bailly, S.; et al. Bone Morphogenetic Protein 9 Is a Paracrine Factor Controlling Liver Sinusoidal Endothelial Cell Fenestration and Protecting Against Hepatic Fibrosis. Hepatology 2019, 70, 1392–1408. [Google Scholar] [CrossRef]
  94. Desroches-Castan, A.; Tillet, E.; Ricard, N.; Ouarne, M.; Mallet, C.; Feige, J.J.; Bailly, S. Differential Consequences of Bmp9 Deletion on Sinusoidal Endothelial Cell Differentiation and Liver Fibrosis in 129/Ola and C57BL/6 Mice. Cells 2019, 8, 1079. [Google Scholar] [CrossRef] [PubMed]
  95. Schmid, C.D.; Olsavszky, V.; Reinhart, M.; Weyer, V.; Trogisch, F.A.; Sticht, C.; Winkler, M.; Kurschner, S.W.; Hoffmann, J.; Ola, R.; et al. ALK1 controls hepatic vessel formation, angiodiversity, and angiocrine functions in hereditary hemorrhagic telangiectasia of the liver. Hepatology 2023, 77, 1211–1227. [Google Scholar] [CrossRef]
  96. Bocci, G.; Orlandi, P.; Manca, M.L.; Rossi, C.; Salvati, A.; Brunetto, M.R.; Solini, A. Predictive Power of Tissue and Circulating Biomarkers for the Severity of Biopsy-Validated Chronic Liver Diseases. J. Clin. Med. 2022, 11, 5985. [Google Scholar] [CrossRef]
  97. Eddowes, L.A.; Al-Hourani, K.; Ramamurthy, N.; Frankish, J.; Baddock, H.T.; Sandor, C.; Ryan, J.D.; Fusco, D.N.; Arezes, J.; Giannoulatou, E.; et al. Antiviral activity of bone morphogenetic proteins and activins. Nat. Microbiol. 2019, 4, 339–351. [Google Scholar] [CrossRef]
  98. Hao, J.; Wang, Y.; Huo, L.; Sun, T.; Zhen, Y.; Gao, Z.; Chen, S.; Ren, L. Circulating Bone Morphogenetic Protein-9 is Decreased in Patients with Type 2 Diabetes and Non-Alcoholic Fatty Liver Disease. Int. J. Gen. Med. 2022, 15, 8539–8546. [Google Scholar] [CrossRef]
  99. Fisher, F.M.; Maratos-Flier, E. Understanding the Physiology of FGF21. Annu. Rev. Physiol. 2016, 78, 223–241. [Google Scholar] [CrossRef] [PubMed]
  100. Drexler, S.; Cai, C.; Hartmann, A.L.; Moch, D.; Gaitantzi, H.; Ney, T.; Kraemer, M.; Chu, Y.; Zheng, Y.; Rahbari, M.; et al. Intestinal BMP-9 locally upregulates FGF19 and is down-regulated in obese patients with diabetes. Mol. Cell. Endocrinol. 2023, 570, 111934. [Google Scholar] [CrossRef]
  101. Yano, K.; Yamaguchi, K.; Seko, Y.; Okishio, S.; Ishiba, H.; Tochiki, N.; Takahashi, A.; Kataoka, S.; Okuda, K.; Liu, Y.; et al. Hepatocyte-specific fibroblast growth factor 21 overexpression ameliorates high-fat diet-induced obesity and liver steatosis in mice. Lab. Investig. 2022, 102, 281–289. [Google Scholar] [CrossRef] [PubMed]
  102. Berglund, E.D.; Li, C.Y.; Bina, H.A.; Lynes, S.E.; Michael, M.D.; Shanafelt, A.B.; Kharitonenkov, A.; Wasserman, D.H. Fibroblast growth factor 21 controls glycemia via regulation of hepatic glucose flux and insulin sensitivity. Endocrinology 2009, 150, 4084–4093. [Google Scholar] [CrossRef]
  103. Kim, S.; Choe, S.; Lee, D.K. BMP-9 enhances fibroblast growth factor 21 expression and suppresses obesity. Biochim. Biophys. Acta 2016, 1862, 1237–1246. [Google Scholar] [CrossRef] [PubMed]
  104. Sun, Q.J.; Cai, L.Y.; Jian, J.; Cui, Y.L.; Huang, C.K.; Liu, S.Q.; Lu, J.L.; Wang, W.; Zeng, X.; Zhong, L. The Role of Bone Morphogenetic Protein 9 in Nonalcoholic Fatty Liver Disease in Mice. Front. Pharmacol. 2020, 11, 605967. [Google Scholar] [CrossRef]
  105. Yang, Z.; Li, P.; Shang, Q.; Wang, Y.; He, J.; Ge, S.; Jia, R.; Fan, X. CRISPR-mediated BMP9 ablation promotes liver steatosis via the down-regulation of PPARalpha expression. Sci. Adv. 2020, 6, 48. [Google Scholar] [CrossRef]
  106. Cheng, L.; Wang, J.; Dai, H.; Duan, Y.; An, Y.; Shi, L.; Lv, Y.; Li, H.; Wang, C.; Ma, Q.; et al. Brown and beige adipose tissue: A novel therapeutic strategy for obesity and type 2 diabetes mellitus. Adipocyte 2021, 10, 48–65. [Google Scholar] [CrossRef]
  107. Lee, E.; Korf, H.; Vidal-Puig, A. An adipocentric perspective on the development and progression of non-alcoholic fatty liver disease. J. Hepatol. 2023, 78, 1048–1062. [Google Scholar] [CrossRef] [PubMed]
  108. Kurylowicz, A.; Puzianowska-Kuznicka, M. Induction of Adipose Tissue Browning as a Strategy to Combat Obesity. Int. J. Mol. Sci. 2020, 21, 6241. [Google Scholar] [CrossRef] [PubMed]
  109. Poher, A.L.; Altirriba, J.; Veyrat-Durebex, C.; Rohner-Jeanrenaud, F. Brown adipose tissue activity as a target for the treatment of obesity/insulin resistance. Front. Physiol. 2015, 6, 4. [Google Scholar] [CrossRef]
  110. Kuo, M.M.; Kim, S.; Tseng, C.Y.; Jeon, Y.H.; Choe, S.; Lee, D.K. BMP-9 as a potent brown adipogenic inducer with anti-obesity capacity. Biomaterials 2014, 35, 3172–3179. [Google Scholar] [CrossRef]
  111. Jiang, Q.; Li, Q.; Liu, B.; Li, G.; Riedemann, G.; Gaitantzi, H.; Breitkopf-Heinlein, K.; Zeng, A.; Ding, H.; Xu, K. BMP9 promotes methionine- and choline-deficient diet-induced nonalcoholic steatohepatitis in non-obese mice by enhancing NF-kappaB dependent macrophage polarization. Int. Immunopharmacol. 2021, 96, 107591. [Google Scholar] [CrossRef]
  112. Igreja Sa, I.C.; Tripska, K.; Hroch, M.; Hyspler, R.; Ticha, A.; Lastuvkova, H.; Schreiberova, J.; Dolezelova, E.; Eissazadeh, S.; Vitverova, B.; et al. Soluble Endoglin as a Potential Biomarker of Nonalcoholic Steatohepatitis (NASH) Development, Participating in Aggravation of NASH-Related Changes in Mouse Liver. Int. J. Mol. Sci. 2020, 21, 9021. [Google Scholar] [CrossRef]
  113. Caussy, C.; Aubin, A.; Loomba, R. The Relationship Between Type 2 Diabetes, NAFLD, and Cardiovascular Risk. Curr. Diabetes Rep. 2021, 21, 15. [Google Scholar] [CrossRef]
  114. Targher, G.; Corey, K.E.; Byrne, C.D.; Roden, M. The complex link between NAFLD and type 2 diabetes mellitus—Mechanisms and treatments. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 599–612. [Google Scholar] [CrossRef]
  115. Younossi, Z.; Tacke, F.; Arrese, M.; Chander Sharma, B.; Mostafa, I.; Bugianesi, E.; Wai-Sun Wong, V.; Yilmaz, Y.; George, J.; Fan, J.; et al. Global Perspectives on Nonalcoholic Fatty Liver Disease and Nonalcoholic Steatohepatitis. Hepatology 2019, 69, 2672–2682. [Google Scholar] [CrossRef]
  116. Tanase, D.M.; Gosav, E.M.; Costea, C.F.; Ciocoiu, M.; Lacatusu, C.M.; Maranduca, M.A.; Ouatu, A.; Floria, M. The Intricate Relationship between Type 2 Diabetes Mellitus (T2DM), Insulin Resistance (IR), and Nonalcoholic Fatty Liver Disease (NAFLD). J. Diabetes Res. 2020, 2020, 3920196. [Google Scholar] [CrossRef] [PubMed]
  117. Ferguson, D.; Finck, B.N. Emerging therapeutic approaches for the treatment of NAFLD and type 2 diabetes mellitus. Nat. Rev. Endocrinol. 2021, 17, 484–495. [Google Scholar] [CrossRef]
  118. Xu, X.; Li, X.; Yang, G.; Li, L.; Hu, W.; Zhang, L.; Liu, H.; Zheng, H.; Tan, M.; Zhu, D. Circulating bone morphogenetic protein-9 in relation to metabolic syndrome and insulin resistance. Sci. Rep. 2017, 7, 17529. [Google Scholar] [CrossRef]
  119. Luo, Y.; Li, L.; Xu, X.; Wu, T.; Yang, M.; Zhang, C.; Mou, H.; Zhou, T.; Jia, Y.; Cai, C.; et al. Decreased circulating BMP-9 levels in patients with Type 2 diabetes is a signature of insulin resistance. Clin. Sci. 2017, 131, 239–246. [Google Scholar] [CrossRef]
  120. Luan, H.; Yang, L.; Liu, L.; Liu, S.; Zhao, X.; Sui, H.; Wang, J.; Wang, S. Effects of platycodins on liver complications of type 2 diabetes. Mol. Med. Rep. 2014, 10, 1597–1603. [Google Scholar] [CrossRef]
  121. Jia, Y.; Niu, D.; Li, Q.; Huang, H.; Li, X.; Li, K.; Li, L.; Zhang, C.; Zheng, H.; Zhu, Z.; et al. Effective gene delivery of shBMP-9 using polyethyleneimine-based core-shell nanoparticles in an animal model of insulin resistance. Nanoscale 2019, 11, 2008–2016. [Google Scholar] [CrossRef]
  122. Caperuto, L.C.; Anhe, G.F.; Cambiaghi, T.D.; Akamine, E.H.; do Carmo Buonfiglio, D.; Cipolla-Neto, J.; Curi, R.; Bordin, S. Modulation of bone morphogenetic protein-9 expression and processing by insulin, glucose, and glucocorticoids: Possible candidate for hepatic insulin-sensitizing substance. Endocrinology 2008, 149, 6326–6335. [Google Scholar] [CrossRef]
  123. Chen, C.; Grzegorzewski, K.J.; Barash, S.; Zhao, Q.; Schneider, H.; Wang, Q.; Singh, M.; Pukac, L.; Bell, A.C.; Duan, R.; et al. An integrated functional genomics screening program reveals a role for BMP-9 in glucose homeostasis. Nat. Biotechnol. 2003, 21, 294–301. [Google Scholar] [CrossRef]
  124. Anhe, F.F.; Lellis-Santos, C.; Leite, A.R.; Hirabara, S.M.; Boschero, A.C.; Curi, R.; Anhe, G.F.; Bordin, S. Smad5 regulates Akt2 expression and insulin-induced glucose uptake in L6 myotubes. Mol. Cell. Endocrinol. 2010, 319, 30–38. [Google Scholar] [CrossRef]
  125. Capasso, T.L.; Trucco, S.M.; Hindes, M.; Schwartze, T.; Bloch, J.L.; Kreutzer, J.; Cook, S.C.; Hinck, C.S.; Treggiari, D.; Feingold, B.; et al. In Search of "Hepatic Factor": Lack of Evidence for ALK1 Ligands BMP9 and BMP10. Am. J. Respir. Crit. Care Med. 2021, 203, 249–251. [Google Scholar] [CrossRef]
  126. Hong, O.K.; Kim, E.S.; Son, J.W.; Kim, S.R.; Yoo, S.J.; Kwon, H.S.; Lee, S.S. Alcohol-induced increase in BMP levels promotes fatty liver disease in male prediabetic stage Otsuka Long-Evans Tokushima Fatty rats. J. Cell. Biochem. 2023, 124, 459–472. [Google Scholar] [CrossRef]
  127. Friedman, S.L.; Pinzani, M. Hepatic fibrosis 2022: Unmet needs and a blueprint for the future. Hepatology 2022, 75, 473–488. [Google Scholar] [CrossRef] [PubMed]
  128. Kisseleva, T.; Brenner, D. Molecular and cellular mechanisms of liver fibrosis and its regression. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 151–166. [Google Scholar] [CrossRef] [PubMed]
  129. Parola, M.; Pinzani, M. Liver fibrosis: Pathophysiology, pathogenetic targets and clinical issues. Mol. Asp. Med. 2019, 65, 37–55. [Google Scholar] [CrossRef]
  130. Zhang, M.; Serna-Salas, S.; Damba, T.; Borghesan, M.; Demaria, M.; Moshage, H. Hepatic stellate cell senescence in liver fibrosis: Characteristics, mechanisms and perspectives. Mech. Ageing Dev. 2021, 199, 111572. [Google Scholar] [CrossRef]
  131. Tsuchida, T.; Friedman, S.L. Mechanisms of hepatic stellate cell activation. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 397–411. [Google Scholar] [CrossRef]
  132. Gines, P.; Krag, A.; Abraldes, J.G.; Sola, E.; Fabrellas, N.; Kamath, P.S. Liver cirrhosis. Lancet 2021, 398, 1359–1376. [Google Scholar] [CrossRef]
  133. Devarbhavi, H.; Asrani, S.K.; Arab, J.P.; Nartey, Y.A.; Pose, E.; Kamath, P.S. Global burden of liver disease: 2023 update. J. Hepatol. 2023, 79, 516–537. [Google Scholar] [CrossRef]
  134. Li, P.; Li, Y.; Zhu, L.; Yang, Z.; He, J.; Wang, L.; Shang, Q.; Pan, H.; Wang, H.; Ma, X.; et al. Targeting secreted cytokine BMP9 gates the attenuation of hepatic fibrosis. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 709–720. [Google Scholar] [CrossRef]
  135. Shen, H.; Fan, J.; Burczynski, F.; Minuk, G.Y.; Cattini, P.; Gong, Y. Increased Smad1 expression and transcriptional activity enhances trans-differentiation of hepatic stellate cells. J. Cell. Physiol. 2007, 212, 764–770. [Google Scholar] [CrossRef]
  136. Chi, C.; Liu, X.Y.; Hou, F.; Yu, X.Z.; Li, C.Y.; Cui, L.J.; Liu, R.X.; Yin, C.H. Herbal compound 861 prevents hepatic fibrosis by inhibiting the TGF-beta1/Smad/SnoN pathway in bile duct-ligated rats. BMC Complement. Altern. Med. 2018, 18, 52. [Google Scholar] [CrossRef]
  137. Leung, D.H.; Devaraj, S.; Goodrich, N.P.; Chen, X.; Rajapakshe, D.; Ye, W.; Andreev, V.; Minard, C.G.; Guffey, D.; Molleston, J.P.; et al. Serum biomarkers correlated with liver stiffness assessed in a multicenter study of pediatric cholestatic liver disease. Hepatology 2023, 77, 530–545. [Google Scholar] [CrossRef]
  138. Clemente, M.; Nunez, O.; Lorente, R.; Rincon, D.; Matilla, A.; Salcedo, M.; Catalina, M.V.; Ripoll, C.; Iacono, O.L.; Banares, R.; et al. Increased intrahepatic and circulating levels of endoglin, a TGF-beta1 co-receptor, in patients with chronic hepatitis C virus infection: Relationship to histological and serum markers of hepatic fibrosis. J. Viral Hepat. 2006, 13, 625–632. [Google Scholar] [CrossRef]
  139. Yagmur, E.; Rizk, M.; Stanzel, S.; Hellerbrand, C.; Lammert, F.; Trautwein, C.; Wasmuth, H.E.; Gressner, A.M. Elevation of endoglin (CD105) concentrations in serum of patients with liver cirrhosis and carcinoma. Eur. J. Gastroenterol. Hepatol. 2007, 19, 755–761. [Google Scholar] [CrossRef] [PubMed]
  140. Rath, T.; Hage, L.; Kugler, M.; Menendez Menendez, K.; Zachoval, R.; Naehrlich, L.; Schulz, R.; Roderfeld, M.; Roeb, E. Serum proteome profiling identifies novel and powerful markers of cystic fibrosis liver disease. PLoS ONE 2013, 8, e58955. [Google Scholar] [CrossRef]
  141. Kwon, Y.C.; Sasaki, R.; Meyer, K.; Ray, R. Hepatitis C Virus Core Protein Modulates Endoglin (CD105) Signaling Pathway for Liver Pathogenesis. J. Virol. 2017, 91, 10–128. [Google Scholar] [CrossRef] [PubMed]
  142. About, F.; Bibert, S.; Jouanguy, E.; Nalpas, B.; Lorenzo, L.; Rattina, V.; Zarhrate, M.; Hanein, S.; Munteanu, M.; Mullhaupt, B.; et al. Identification of an Endoglin Variant Associated With HCV-Related Liver Fibrosis Progression by Next-Generation Sequencing. Front. Genet. 2019, 10, 1024. [Google Scholar] [CrossRef] [PubMed]
  143. Meurer, S.K.; Tihaa, L.; Borkham-Kamphorst, E.; Weiskirchen, R. Expression and functional analysis of endoglin in isolated liver cells and its involvement in fibrogenic Smad signalling. Cell. Signal. 2011, 23, 683–699. [Google Scholar] [CrossRef] [PubMed]
  144. Finnson, K.W.; Philip, A. Endoglin in liver fibrosis. J. Cell Commun. Signal. 2012, 6, 1–4. [Google Scholar] [CrossRef] [PubMed]
  145. Alsamman, M.; Sterzer, V.; Meurer, S.K.; Sahin, H.; Schaeper, U.; Kuscuoglu, D.; Strnad, P.; Weiskirchen, R.; Trautwein, C.; Scholten, D. Endoglin in human liver disease and murine models of liver fibrosis-A protective factor against liver fibrosis. Liver Int. Off. J. Int. Assoc. Study Liver 2018, 38, 858–867. [Google Scholar] [CrossRef] [PubMed]
  146. Kouroumalis, E.; Tsomidis, I.; Voumvouraki, A. Iron as a therapeutic target in chronic liver disease. World J. Gastroenterol. 2023, 29, 616–655. [Google Scholar] [CrossRef] [PubMed]
  147. Babitt, J.L.; Huang, F.W.; Wrighting, D.M.; Xia, Y.; Sidis, Y.; Samad, T.A.; Campagna, J.A.; Chung, R.T.; Schneyer, A.L.; Woolf, C.J.; et al. Bone morphogenetic protein signaling by hemojuvelin regulates hepcidin expression. Nat. Genet. 2006, 38, 531–539. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, C.Y.; Xiao, X.; Bayer, A.; Xu, Y.; Dev, S.; Canali, S.; Nair, A.V.; Masia, R.; Babitt, J.L. Ablation of Hepatocyte Smad1, Smad5, and Smad8 Causes Severe Tissue Iron Loading and Liver Fibrosis in Mice. Hepatology 2019, 70, 1986–2002. [Google Scholar] [CrossRef] [PubMed]
  149. Sharma, R.; Zhao, W.; Zafar, Y.; Murali, A.R.; Brown, K.E. Serum hepcidin levels in chronic liver disease: A systematic review and meta-analysis. Clin. Chem. Lab. Med. 2023, 62, 373–384. [Google Scholar] [CrossRef]
  150. Fujita, N.; Sugimoto, R.; Takeo, M.; Urawa, N.; Mifuji, R.; Tanaka, H.; Kobayashi, Y.; Iwasa, M.; Watanabe, S.; Adachi, Y.; et al. Hepcidin expression in the liver: Relatively low level in patients with chronic hepatitis C. Mol. Med. 2007, 13, 97–104. [Google Scholar] [CrossRef]
  151. Tsochatzis, E.; Papatheodoridis, G.V.; Koliaraki, V.; Hadziyannis, E.; Kafiri, G.; Manesis, E.K.; Mamalaki, A.; Archimandritis, A.J. Serum hepcidin levels are related to the severity of liver histological lesions in chronic hepatitis C. J. Viral Hepat. 2010, 17, 800–806. [Google Scholar] [CrossRef] [PubMed]
  152. Mantovani, A.; Csermely, A.; Castagna, A.; Antinori, E.; Danese, E.; Zusi, C.; Sani, E.; Ravaioli, F.; Colecchia, A.; Maffeis, C.; et al. Associations between higher plasma ferritin and hepcidin levels with liver stiffness in patients with type 2 diabetes: An exploratory study. Liver Int. Off. J. Int. Assoc. Study Liver 2023, 43, 2434–2444. [Google Scholar] [CrossRef] [PubMed]
  153. Lyberopoulou, A.; Chachami, G.; Gatselis, N.K.; Kyratzopoulou, E.; Saitis, A.; Gabeta, S.; Eliades, P.; Paraskeva, E.; Zachou, K.; Koukoulis, G.K.; et al. Low Serum Hepcidin in Patients with Autoimmune Liver Diseases. PLoS ONE 2015, 10, e0135486. [Google Scholar] [CrossRef]
  154. Lu, S.; Bennett, R.G.; Kharbanda, K.K.; Harrison-Findik, D.D. Lack of hepcidin expression attenuates steatosis and causes fibrosis in the liver. World J. Hepatol. 2016, 8, 211–225. [Google Scholar] [CrossRef] [PubMed]
  155. Chen, H.; Zhao, W.; Yan, X.; Huang, T.; Yang, A. Overexpression of Hepcidin Alleviates Steatohepatitis and Fibrosis in a Diet-induced Nonalcoholic Steatohepatitis. J. Clin. Transl. Hepatol. 2022, 10, 577–588. [Google Scholar] [CrossRef] [PubMed]
  156. Miyazono, K.; Miyazawa, K. Id: A target of BMP signaling. Sci. STKE Signal Transduct. Knowl. Environ. 2002, 2002, pe40. [Google Scholar] [CrossRef]
  157. Wiercinska, E.; Wickert, L.; Denecke, B.; Said, H.M.; Hamzavi, J.; Gressner, A.M.; Thorikay, M.; ten Dijke, P.; Mertens, P.R.; Breitkopf, K.; et al. Id1 is a critical mediator in TGF-beta-induced transdifferentiation of rat hepatic stellate cells. Hepatology 2006, 43, 1032–1041. [Google Scholar] [CrossRef] [PubMed]
  158. Wei, M.; Yan, X.; Xin, X.; Chen, H.; Hou, L.; Zhang, J. Hepatocyte-Specific Smad4 Deficiency Alleviates Liver Fibrosis via the p38/p65 Pathway. Int. J. Mol. Sci. 2022, 23, 1696. [Google Scholar] [CrossRef]
  159. Matsuda, Y.; Yamagiwa, S.; Takamura, M.; Honda, Y.; Ishimoto, Y.; Ichida, T.; Aoyagi, Y. Overexpressed Id-1 is associated with a high risk of hepatocellular carcinoma development in patients with cirrhosis without transcriptional repression of p16. Cancer 2005, 104, 1037–1044. [Google Scholar] [CrossRef]
  160. Dong, X.C.; Chowdhury, K.; Huang, M.; Kim, H.G. Signal Transduction and Molecular Regulation in Fatty Liver Disease. Antioxid. Redox Signal 2021, 35, 689–717. [Google Scholar] [CrossRef]
  161. Certain, M.C.; Robert, F.; Baron, A.; Sitbon, O.; Humbert, M.; Guignabert, C.; Tu, L.; Savale, L. Hepatopulmonary syndrome: Prevalence, pathophysiology and clinical implications. Rev. Mal. Respir. 2022, 39, 84–89. [Google Scholar] [CrossRef] [PubMed]
  162. Raevens, S.; Boret, M.; Fallon, M.B. Hepatopulmonary syndrome. JHEP Rep. Innov. Hepatol. 2022, 4, 100527. [Google Scholar] [CrossRef] [PubMed]
  163. Toshner, M. BMP9 Morphs into a Potential Player in Portopulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2019, 199, 819–821. [Google Scholar] [CrossRef] [PubMed]
  164. Owen, N.E.; Alexander, G.J.; Sen, S.; Bunclark, K.; Polwarth, G.; Pepke-Zaba, J.; Davenport, A.P.; Morrell, N.W.; Upton, P.D. Reduced circulating BMP10 and BMP9 and elevated endoglin are associated with disease severity, decompensation and pulmonary vascular syndromes in patients with cirrhosis. EBioMedicine 2020, 56, 102794. [Google Scholar] [CrossRef] [PubMed]
  165. Nikolic, I.; Yung, L.M.; Yang, P.; Malhotra, R.; Paskin-Flerlage, S.D.; Dinter, T.; Bocobo, G.A.; Tumelty, K.E.; Faugno, A.J.; Troncone, L.; et al. Bone Morphogenetic Protein 9 Is a Mechanistic Biomarker of Portopulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2019, 199, 891–902. [Google Scholar] [CrossRef] [PubMed]
  166. Rochon, E.R.; Krowka, M.J.; Bartolome, S.; Heresi, G.A.; Bull, T.; Roberts, K.; Hemnes, A.; Forde, K.A.; Krok, K.L.; Patel, M.; et al. BMP9/10 in Pulmonary Vascular Complications of Liver Disease. Am. J. Respir. Crit. Care Med. 2020, 201, 1575–1578. [Google Scholar] [CrossRef] [PubMed]
  167. Roman, B.L.; Hinck, A.P. ALK1 signaling in development and disease: New paradigms. Cell. Mol. Life Sci. CMLS 2017, 74, 4539–4560. [Google Scholar] [CrossRef] [PubMed]
  168. Yang, Y.; Yu, H.; Yang, C.; Zhang, Y.; Ai, X.; Wang, X.; Lu, K.; Yi, B. Kruppel-like factor 6 mediates pulmonary angiogenesis in rat experimental hepatopulmonary syndrome and is aggravated by bone morphogenetic protein 9. Biol. Open 2019, 8, bio040121. [Google Scholar] [CrossRef]
  169. Yang, C.; Sun, M.; Yang, Y.; Han, Y.; Wu, X.; Wu, X.; Cao, H.; Chen, L.; Lei, Y.; Hu, X.; et al. Elevated circulating BMP9 aggravates pulmonary angiogenesis in hepatopulmonary syndrome rats through ALK1-Endoglin-Smad1/5/9 signalling. Eur. J. Clin. Invest. 2024, 54, e14212. [Google Scholar] [CrossRef]
  170. Robert, F.; Certain, M.C.; Baron, A.; Thuillet, R.; Duhaut, L.; Ottaviani, M.; Kamel Chelgham, M.; Normand, C.; Berrebeh, N.; Ricard, N.; et al. Disrupted BMP-9 Signaling Impairs Pulmonary Vascular Integrity in Hepatopulmonary Syndrome. Am. J. Respir. Crit. Care Med. 2024. [Google Scholar] [CrossRef]
  171. Sangam, S.; Yu, P.B. Hepatopulmonary Syndrome, Another Face of Dysregulated BMP9 Signaling. Am. J. Respir. Crit. Care Med. 2024. [Google Scholar] [CrossRef]
  172. Alkhathami, A.G.; Abdullah, M.R.; Ahmed, M.; Hassan Ahmed, H.; Alwash, S.W.; Muhammed Mahdi, Z.; Alsaikhan, F.; Dera, A.A. Bone morphogenetic protein (BMP)9 in cancer development: Mechanistic, diagnostic, and therapeutic approaches? J. Drug Target. 2023, 31, 714–724. [Google Scholar] [CrossRef]
  173. Garcia-Alvaro, M.; Addante, A.; Roncero, C.; Fernandez, M.; Fabregat, I.; Sanchez, A.; Herrera, B. BMP9-Induced Survival Effect in Liver Tumor Cells Requires p38MAPK Activation. Int. J. Mol. Sci. 2015, 16, 20431. [Google Scholar] [CrossRef] [PubMed]
  174. Chen, H.; Nio, K.; Yamashita, T.; Okada, H.; Li, R.; Suda, T.; Li, Y.; Doan, P.T.B.; Seki, A.; Nakagawa, H.; et al. BMP9-ID1 signaling promotes EpCAM-positive cancer stem cell properties in hepatocellular carcinoma. Mol. Oncol. 2021, 15, 2203–2218. [Google Scholar] [CrossRef]
  175. Giannelli, G.; Koudelkova, P.; Dituri, F.; Mikulits, W. Role of epithelial to mesenchymal transition in hepatocellular carcinoma. J. Hepatol. 2016, 65, 798–808. [Google Scholar] [CrossRef] [PubMed]
  176. Kim, M.; Choi, O.; Pyo, S.; Choi, S.U.; Park, C.H. Identification of novel ALK2 inhibitors and their effect on cancer cells. Biochem. Biophys. Res. Commun. 2017, 492, 121–127. [Google Scholar] [CrossRef] [PubMed]
  177. Dhillon, A.S.; Hagan, S.; Rath, O.; Kolch, W. MAP kinase signalling pathways in cancer. Oncogene 2007, 26, 3279–3290. [Google Scholar] [CrossRef]
  178. Garcia-Hernandez, L.; Garcia-Ortega, M.B.; Ruiz-Alcala, G.; Carrillo, E.; Marchal, J.A.; Garcia, M.A. The p38 MAPK Components and Modulators as Biomarkers and Molecular Targets in Cancer. Int. J. Mol. Sci. 2021, 23, 370. [Google Scholar] [CrossRef]
  179. He, H.; Qiao, K.; Wang, C.; Yang, W.; Xu, Z.; Zhang, Z.; Jia, Y.; Zhang, C.; Peng, L. Hydrazinocurcumin Induces Apoptosis of Hepatocellular Carcinoma Cells Through the p38 MAPK Pathway. Clin. Transl. Sci. 2021, 14, 2075–2084. [Google Scholar] [CrossRef]
  180. Han, Z.; Zhao, X.; Zhang, E.; Ma, J.; Zhang, H.; Li, J.; Xie, W.; Li, X. Resistomycin Induced Apoptosis and Cycle Arrest in Human Hepatocellular Carcinoma Cells by Activating p38 MAPK Pathway In Vitro and In Vivo. Pharmaceuticals 2021, 14, 958. [Google Scholar] [CrossRef]
  181. Gomez-Puerto, M.C.; Iyengar, P.V.; Garcia de Vinuesa, A.; Ten Dijke, P.; Sanchez-Duffhues, G. Bone morphogenetic protein receptor signal transduction in human disease. J. Pathol. 2019, 247, 9–20. [Google Scholar] [CrossRef] [PubMed]
  182. Yu, F.; Yu, C.; Li, F.; Zuo, Y.; Wang, Y.; Yao, L.; Wu, C.; Wang, C.; Ye, L. Wnt/beta-catenin signaling in cancers and targeted therapies. Signal Transduct. Target. Ther. 2021, 6, 307. [Google Scholar] [CrossRef]
  183. Zhang, Y.; Wang, X. Targeting the Wnt/beta-catenin signaling pathway in cancer. J. Hematol. Oncol. 2020, 13, 165. [Google Scholar] [CrossRef]
  184. Katoh, M. Multi-layered prevention and treatment of chronic inflammation, organ fibrosis and cancer associated with canonical WNT/beta-catenin signaling activation (Review). Int. J. Mol. Med. 2018, 42, 713–725. [Google Scholar] [CrossRef]
  185. Deldar Abad Paskeh, M.; Mirzaei, S.; Ashrafizadeh, M.; Zarrabi, A.; Sethi, G. Wnt/beta-Catenin Signaling as a Driver of Hepatocellular Carcinoma Progression: An Emphasis on Molecular Pathways. J. Hepatocell. Carcinoma 2021, 8, 1415–1444. [Google Scholar] [CrossRef]
  186. Chen, H.; Zhang, M.; Li, J.; Liu, M.; Cao, D.; Li, Y.Y.; Yamashita, T.; Nio, K.; Tang, H. BMP9-ID1 Pathway Attenuates N(6)-Methyladenosine Levels of CyclinD1 to Promote Cell Proliferation in Hepatocellular Carcinoma. Int. J. Mol. Sci. 2024, 25, 981. [Google Scholar] [CrossRef]
  187. Morse, M.A.; Sun, W.; Kim, R.; He, A.R.; Abada, P.B.; Mynderse, M.; Finn, R.S. The Role of Angiogenesis in Hepatocellular Carcinoma. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2019, 25, 912–920. [Google Scholar] [CrossRef] [PubMed]
  188. Yao, C.; Wu, S.; Kong, J.; Sun, Y.; Bai, Y.; Zhu, R.; Li, Z.; Sun, W.; Zheng, L. Angiogenesis in hepatocellular carcinoma: Mechanisms and anti-angiogenic therapies. Cancer Biol. Med. 2023, 20, 25–43. [Google Scholar] [CrossRef]
  189. Chu, Q.; Gu, X.; Zheng, Q.; Zhu, H. Regulatory mechanism of HIF-1alpha and its role in liver diseases: A narrative review. Ann. Transl. Med. 2022, 10, 109. [Google Scholar] [CrossRef] [PubMed]
  190. Bao, M.H.; Wong, C.C. Hypoxia, Metabolic Reprogramming, and Drug Resistance in Liver Cancer. Cells 2021, 10, 1715. [Google Scholar] [CrossRef]
  191. Chen, H.; Nio, K.; Tang, H.; Yamashita, T.; Okada, H.; Li, Y.; Doan, P.T.B.; Li, R.; Lv, J.; Sakai, Y.; et al. BMP9-ID1 Signaling Activates HIF-1alpha and VEGFA Expression to Promote Tumor Angiogenesis in Hepatocellular Carcinoma. Int. J. Mol. Sci. 2022, 23, 1475. [Google Scholar] [CrossRef]
  192. Han, Y.; Pan, Q.; Guo, Z.; Du, Y.; Zhang, Y.; Liu, Y.; Zhao, J.; Xu, J.; Yang, J.; Ouyang, D.; et al. BMP9-induced vascular normalisation improves the efficacy of immunotherapy against hepatitis B virus-associated hepatocellular carcinoma. Clin. Transl. Med. 2023, 13, e1247. [Google Scholar] [CrossRef] [PubMed]
  193. Jiang, Q.Q.; Liu, B.B.; Xu, K.S. New insights into BMP9 signaling in liver diseases. Mol. Cell. Biochem. 2021, 476, 3591–3600. [Google Scholar] [CrossRef]
  194. Abou-Alfa, G.K.; Miksad, R.A.; Tejani, M.A.; Williamson, S.; Gutierrez, M.E.; Olowokure, O.O.; Sharma, M.R.; El Dika, I.; Sherman, M.L.; Pandya, S.S. A Phase Ib, Open-Label Study of Dalantercept, an Activin Receptor-Like Kinase 1 Ligand Trap, plus Sorafenib in Advanced Hepatocellular Carcinoma. Oncologist 2019, 24, 161-e70. [Google Scholar] [CrossRef] [PubMed]
  195. Goff, L.W.; Cohen, R.B.; Berlin, J.D.; de Braud, F.G.; Lyshchik, A.; Noberasco, C.; Bertolini, F.; Carpentieri, M.; Stampino, C.G.; Abbattista, A.; et al. A Phase I Study of the Anti-Activin Receptor-Like Kinase 1 (ALK-1) Monoclonal Antibody PF-03446962 in Patients with Advanced Solid Tumors. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2016, 22, 2146–2154. [Google Scholar] [CrossRef]
  196. Simonelli, M.; Zucali, P.; Santoro, A.; Thomas, M.B.; de Braud, F.G.; Borghaei, H.; Berlin, J.; Denlinger, C.S.; Noberasco, C.; Rimassa, L.; et al. Phase I study of PF-03446962, a fully human monoclonal antibody against activin receptor-like kinase-1, in patients with hepatocellular carcinoma. Ann. Oncol. 2016, 27, 1782–1787. [Google Scholar] [CrossRef]
Figure 1. BMP proteins and receptors. The summary of BMP family members and their alternative names, type I and type II receptors, and Co-Smads. OP, osteogenic protein; Vgr, vegetal related; RAI, atrial isomerism right; KFS3, Klippel-Feil syndrome 3, autosomal dominant; ALK, anaplastic lymphoma kinase; BMPRIA, bone morphogenetic protein receptor IA; BMPRIB, bone morphogenetic protein receptor IB; ACVR1, Activin A receptor, type I.
Figure 1. BMP proteins and receptors. The summary of BMP family members and their alternative names, type I and type II receptors, and Co-Smads. OP, osteogenic protein; Vgr, vegetal related; RAI, atrial isomerism right; KFS3, Klippel-Feil syndrome 3, autosomal dominant; ALK, anaplastic lymphoma kinase; BMPRIA, bone morphogenetic protein receptor IA; BMPRIB, bone morphogenetic protein receptor IB; ACVR1, Activin A receptor, type I.
Biomolecules 14 01013 g001
Figure 2. Synthesis procedure for BMP9. BMP9 is synthesized as a 429 amino acid (aa) precursor protein known as pre–pro-BMP9. It consists of a 22 aa signal peptide, a 297 aa prodomain, and a 110 aa mature protein. After synthesis, the pre-pro-BMP9 undergoes homodimerization and subsequent cleavage, resulting in two active forms: the short mature BMP9 (25 kDa) and the complexed form (100 kDa). The short mature BMP9 (25 kDa) includes two forms of dimers, with (D-form) or without (M-form) an intermolecular disulfide bond. Within circulating BMP9, the inactive unprocessed pro-BMP9 accounts for 40%, while the active complexed form accounts for 60%.
Figure 2. Synthesis procedure for BMP9. BMP9 is synthesized as a 429 amino acid (aa) precursor protein known as pre–pro-BMP9. It consists of a 22 aa signal peptide, a 297 aa prodomain, and a 110 aa mature protein. After synthesis, the pre-pro-BMP9 undergoes homodimerization and subsequent cleavage, resulting in two active forms: the short mature BMP9 (25 kDa) and the complexed form (100 kDa). The short mature BMP9 (25 kDa) includes two forms of dimers, with (D-form) or without (M-form) an intermolecular disulfide bond. Within circulating BMP9, the inactive unprocessed pro-BMP9 accounts for 40%, while the active complexed form accounts for 60%.
Biomolecules 14 01013 g002
Figure 3. Diagram of TGF-β and BMP9 signaling pathways. BMP9 binds to heterotetrameric receptor complexes (ALK1/ALK2 and BMPR-II/ActR-IIA/ActR-IIB), resulting in the phosphorylation of downstream Smad1/5/8 (pSmad1/5/8). pSmad1/5/8 subsequently interacts with Smad4, translocates into the nucleus, and binds to the promoters of target genes, thereby promoting gene expression. The co-receptor endoglin plays a crucial role in activating the BMP9—pSmad1/5/8 pathway. Similarly, TGF-β binds to heterotetrameric receptor complexes (ALK5 and TGFBR2), leading to the phosphorylation of downstream Smad2/3 (pSmad2/3). pSmad2/3 then interacts with Smad4, translocates into the nucleus, and binds to the promoters of target genes, thereby promoting gene expression. pSmad1/5/8 are negative regulators of Smad3-dependent gene transcription. Smad6 and Smad7 inhibit the phosphorylation of Smad1/5/8 and Smad2/3.
Figure 3. Diagram of TGF-β and BMP9 signaling pathways. BMP9 binds to heterotetrameric receptor complexes (ALK1/ALK2 and BMPR-II/ActR-IIA/ActR-IIB), resulting in the phosphorylation of downstream Smad1/5/8 (pSmad1/5/8). pSmad1/5/8 subsequently interacts with Smad4, translocates into the nucleus, and binds to the promoters of target genes, thereby promoting gene expression. The co-receptor endoglin plays a crucial role in activating the BMP9—pSmad1/5/8 pathway. Similarly, TGF-β binds to heterotetrameric receptor complexes (ALK5 and TGFBR2), leading to the phosphorylation of downstream Smad2/3 (pSmad2/3). pSmad2/3 then interacts with Smad4, translocates into the nucleus, and binds to the promoters of target genes, thereby promoting gene expression. pSmad1/5/8 are negative regulators of Smad3-dependent gene transcription. Smad6 and Smad7 inhibit the phosphorylation of Smad1/5/8 and Smad2/3.
Biomolecules 14 01013 g003
Figure 4. Smad-dependent pathway and non-Smad pathways.
Figure 4. Smad-dependent pathway and non-Smad pathways.
Biomolecules 14 01013 g004
Figure 5. Schematic diagram of the effects of BMP9 on liver diseases.
Figure 5. Schematic diagram of the effects of BMP9 on liver diseases.
Biomolecules 14 01013 g005
Table 1. The roles of BMP9 in NALFD.
Table 1. The roles of BMP9 in NALFD.
Molecular MechanismEffects of the PathwayEffects of BMP9Study TypeReference
BMP9 promotes M1 macrophage polarizationBMP9 enhances NF-κB signalingBMP9 promotes nonalcoholic steatohepatitisAnimal modelJiang (2021) [111,125]
BMP9 reduces obesity, improves glucose metabolism, alleviates hepatic steatosis, and decreases inflammationBMP9 decreases the promoter chromatin accessibility of CERS6, FABP4, FOS, and TLR1BMP9 suppresses NALFDAnimal modelSun (2021) [104]
FINS, LDL-C, HDL-C, and BMI were independent factors impacting serum BMP9 levelsSerum levels of BMP9 are significantly lower in patients with T2DM and NAFLDBMP9 is an independent risk factor for patients with T2DM and NAFLDCross-sectional studyHao (2022) [98]
Alcohol induces the activation of BMP signalingAlcohol induces the expression of BMP2, BMP4, BMP7, BMP9, Smad1, and Smad4BMP signaling exerts anti- and/or pro-fibrotic effectsAnimal modelHong (2023) [126]
BMP9 attenuates triglyceride accumulation by enhancing PPARα promoter activity via the activation of pSmadsBMP9 attenuates triglyceride accumulation by enhancing PPARα promoter activity via the activation of pSmadsLoss of BMP9 promotes liver steatosisCell line and animal modelYang (2020) [105]
BMP9 decreases ALT as well as cholesterol and enhances brown adipogenesisBMP9 promotes UCP1 expression through the activation of FGF21BMP9 suppresses NALFD and obesityAnimal modelKim (2016) [103]
There is a positive correlation between BMP9 levels and NASH/NAFLD Patients with BMP9 > 1188 pg/mL show worse disease in NASH/NAFLDBMP9 can be biomarker for NASH/NAFLDCross-sectional studyBocci (2022) [96]
Hepatic BMP9 expression negatively correlates with steatosis in obese patients with diabetesBMP9 directly induces FGF19 in gutUnder pre-steatotic conditions, it is likely that increased levels of BMP9 would have beneficial effectsCross-sectional studyDrexler (2023) [100]
BMP9 is negatively associated with WHR, FBG, 2h-OGTT, HbA1c, TG levels, and HOMA-IR; on the other hand, BMP9 is positively associated with FFA and HDL levelsCirculating BMP9 is negatively associated with metabolic syndrome and insulin resistanceBMP9 is independently associated with T2DM, HOMA-IR, and FFACross-sectional studyXu (2017) [118]
BMP9 level is negatively correlated with HbA1c, FBG, OGTT, AUC glucose, and HOMA-IRCirculating BMP9 levels are significantly lower in newly diagnosed T2DM patients compared to healthy peopleThe decreased levels of circulating BMP9 could serve as a marker of insulin resistance in T2DM patients Cross-sectional studyLuo (2017) [119]
Platycodin induces BMP9 expression and reduces Smad4 expressionBMP9 rectifies blood glucose and lipid metabolism disordersAnimal modelLuan (2014) [120]
The deletion of BMP9 elevates the PEPCK protein levels and reduces the levels of InsR and Akt phosphorylation The deletion of BMP9 exacerbates insulin resistance and glucose intolerance in mouse models fed a high-fat dietAnimal modelJia (2019) [121]
CERS6, Ceramide synthase 6; FABP4, Fatty Acid Binding Protein 4; TLR1, Toll Like Receptor 1; NAFLD, nonalcoholic fatty liver disease; FINS, fasting insulin; LDL-C, low-density lipoprotein cholesterol; HDL-C, high-density lipoprotein cholesterol; BMI, body mass index; T2DM, type 2 diabetes mellitus; UCP1, uncoupling protein 1; FGF21, fibroblast growth factor 21; WHR, waist–hip ratio; FBG, fasting blood glucose; 2h-OGTT, 2 h blood glucose after glucose overload; TG, triglyceride; HOMA-IR, homoeostasis model assessment of insulin resistance; FFA, free fatty acid; AUC glucose, the area under the curve for glucose; PEPCK, phosphoenolpyruvate carboxykinase; InsR, insulin receptor.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, H.; Li, Y.-Y.; Nio, K.; Tang, H. Unveiling the Impact of BMP9 in Liver Diseases: Insights into Pathogenesis and Therapeutic Potential. Biomolecules 2024, 14, 1013. https://doi.org/10.3390/biom14081013

AMA Style

Chen H, Li Y-Y, Nio K, Tang H. Unveiling the Impact of BMP9 in Liver Diseases: Insights into Pathogenesis and Therapeutic Potential. Biomolecules. 2024; 14(8):1013. https://doi.org/10.3390/biom14081013

Chicago/Turabian Style

Chen, Han, Ying-Yi Li, Kouki Nio, and Hong Tang. 2024. "Unveiling the Impact of BMP9 in Liver Diseases: Insights into Pathogenesis and Therapeutic Potential" Biomolecules 14, no. 8: 1013. https://doi.org/10.3390/biom14081013

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop