Next Article in Journal
Calcineurin-Dependent Homeostatic Response of C. elegans Muscle Cells upon Prolonged Activation of Acetylcholine Receptors
Previous Article in Journal
Morphology of Neutrophils during Their Activation and NETosis: Atomic Force Microscopy Study
Previous Article in Special Issue
Genetic and Pharmacological YAP Activation Induces Proliferation and Improves Survival in Human Induced Pluripotent Stem Cell-Derived Cardiomyocytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Human iPSCs as Model Systems for BMP-Related Rare Diseases

by
Gonzalo Sánchez-Duffhues
1,2,*,† and
Christian Hiepen
3,*,†
1
Nanomaterials and Nanotechnology Research Center (CINN-CSIC), ISPA-HUCA, Avda. de Roma, s/n, 33011 Oviedo, Spain
2
Department of Cell and Chemical Biology, Leiden University Medical Center, Einthovenweg 20, 2333 ZC Leiden, The Netherlands
3
Department of Engineering and Natural Sciences, Westphalian University of Applied Sciences, August-Schmidt-Ring 10, 45665 Recklinghausen, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2023, 12(17), 2200; https://doi.org/10.3390/cells12172200
Submission received: 18 July 2023 / Revised: 17 August 2023 / Accepted: 23 August 2023 / Published: 2 September 2023
(This article belongs to the Special Issue iPS Cells (iPSCs) for Modelling and Treatment of Human Diseases 2022)

Abstract

:
Disturbances in bone morphogenetic protein (BMP) signalling contribute to onset and development of a number of rare genetic diseases, including Fibrodysplasia ossificans progressiva (FOP), Pulmonary arterial hypertension (PAH), and Hereditary haemorrhagic telangiectasia (HHT). After decades of animal research to build a solid foundation in understanding the underlying molecular mechanisms, the progressive implementation of iPSC-based patient-derived models will improve drug development by addressing drug efficacy, specificity, and toxicity in a complex humanized environment. We will review the current state of literature on iPSC-derived model systems in this field, with special emphasis on the access to patient source material and the complications that may come with it. Given the essential role of BMPs during embryonic development and stem cell differentiation, gain- or loss-of-function mutations in the BMP signalling pathway may compromise iPSC generation, maintenance, and differentiation procedures. This review highlights the need for careful optimization of the protocols used. Finally, we will discuss recent developments towards complex in vitro culture models aiming to resemble specific tissue microenvironments with multi-faceted cellular inputs, such as cell mechanics and ECM together with organoids, organ-on-chip, and microfluidic technologies.
Keywords:
PAH; HHT; FOP; FAP; TGF-beta

Graphical Abstract

1. Introduction:

1.1. Use of iPSCs for Efficient BMP Rare Disease Research and Orphan Drug Discovery

The study of rare diseases associated with mutations in the bone morphogenetic protein (BMP) signalling pathway has been advanced by the advent of human induced pluripotent stem cell (hiPSC) technology through Sir John B. Gurdon, Shinya Yamanaka, and co-workers in 2006 [1], reviewed in [2]. The definition of “rare” or “orphan” disease depends on the jurisdiction of every country. A condition that affects 1 in 2500 individuals may be deemed “rare” based on a global average [3]. Genetic factors account for more than 80% of rare diseases [4]. Given the low numbers of individuals affected, which limits the commercial impact of a marketable drug, the pharmaceutical industry may have less interest in the study of rare diseases. To compensate for this, the pharmaceutical authorities often implement measures to attract the industry into the field of orphan diseases, for example, by accelerating the bureaucratic procedures, offering longer patent licenses, and reducing the number of individuals required for clinical trials. However, the status quo is more complex. Learning from rare diseases, pathogenic mechanisms may be identified that can be applied to more prevalent conditions. The BMP-related rare genetic disease (herein referred to as BMP rare disease) Fibrodysplasia ossificans progressiva (FOP) is one such example. In FOP, genetic mutations in the ACVR1 gene (ALK2) (described later in more detail) induce extraskeletal bone formation [5,6]. Mechanistically, the mutant ALK2 causes a stimuli-dependent over-activation of the BMP pathway, for which small molecule inhibitors have been developed. Studies with such compounds have revealed that systemic BMP inhibition may lead to iron overload. Therefore, some of these molecules, which may not be ultimately adequate for FOP, are repositioned into more common disorders like anaemia (clinical trial NCT05090891). Two related rare vascular diseases that are currently being investigated by iPSC technology and which will be covered here are Pulmonary arterial hypertension (PAH) and Hereditary haemorrhagic telangiectasia (HHT), also known as Morbus Osler–Weber–Rendu disease. A deeper understanding of PAH and HHT promises benefits for other, more prevalent (cardio)vascular diseases with related biological features.
Throughout the drug development pipeline, about 90% of the screened compounds usually drop out before the pre-clinical phase is initiated [7]. Even after entering pre-clinical testing, promising candidates may be discarded because results from in vitro experiments cannot be reproduced in animals. As we will discuss later in the cases of FOP and PAH, disease animal models do not necessarily resemble the entire human patient pathophysiology [8]. Therefore, drug mechanisms discovered in animals may translate poorly to patients during clinical trials, dampening the success rate of a therapeutic compound and its ability to reach the market. In order to facilitate the transition from preclinical research to clinical trials, it is crucial to develop and implement appropriate humanized cellular systems for evaluating drug efficacy, toxicity, and pharmacokinetics. Human derived iPSCs are superior to human primary cells or cell lines with unlimited laboratory life-span if iPSCs are generated, cultured, and differentiated under the right conditions. Applying an organoid culture of human-iPSC-derived cell types with complex environmental cues can mimic major aspects of the healthy and diseased human body [9]. These additional cues may include, for example, biomechanical stimulation, hypoxic and/or inflammatory microenvironments, and the right co-culture combination of cell types involved in the biological process. Ideally, these platforms will allow researchers to interrogate the complexity of BMP signalling, and identify novel molecules with potential as druggable targets and/or biomarkers in rare diseases.

1.2. Basics in BMP Signalling

Following their discovery by Marshall Urist as ectopic bone inducing agents in rodents and their subsequent isolation and characterization [10,11,12], the BMPs were included as members of the transforming growth factor (TGF)-β superfamily, due to their high sequence and structural homology with existing TGF-β members. The TGF-β family of cytokines consists of soluble growth and differentiation factors, which exert pleiotropic functions through the activation of specific heterotetrameric receptor complexes on the cell membrane (Figure 1). The BMP receptors exhibit serine/threonine kinase activity, transducing extracellular signals into intracellular signalling networks to regulate, e.g., gene expression. During the past decades, comprehensive studies have emphasized the complexity of the TGF-β signalling pathway at different levels. At the extracellular level, over 30 different ligands, membrane and soluble receptors, natural antagonists, and extracellular matrix (ECM) proteins interact and define the activation of different subsets of TGF-β receptors (TGFβRs). Intracellularly, the different branches of the pathway tend to counteract one another and crosstalk to a myriad of other signalling pathways. As a result, complex intertwined transcriptional networks are regulated, in cooperation with tissue- and stimulus-specific transcriptional (co)-factors.
Canonical BMP signalling is initiated upon the interaction between soluble dimeric BMP ligands and so-called BMP type I receptors (BMPRI) (Figure 1). Four BMPRI, also termed activin-like-kinase (ALK) receptors, have been described: ALK1, 2, 3, and 6. The different BMP ligands display distinct affinities for the ALK receptors (i.e., BMP9/10 are high affinity ALK1 binders, BMP-2/-4 preferably bind ALK3/6, and BMP-5/-6/-7 preferably bind ALK2, reviewed in [13,14]), but the interaction is determined by the expression levels and availability of ligands and receptors. As such, soluble and membrane-bound factors like the co-receptors endoglin (ENG) and Cripto (TDGF1) can enhance BMP-9/-10-induced interaction with ALK1 [15,16]. These co-receptors can be shed off the membrane and circulate in the extracellular space [15,17,18]. The repulsive guidance molecules (RGMs) can either hamper or enhance ligand–receptor interaction [19]. BAMBI (BMP and activin membrane-bound inhibitor) is a transmembrane protein able to bind extracellular ligands, but it lacks the intracellular BMP type I receptor kinase domain [20]. Neogenin can bind and inhibit signalling induced by BMP-2, -4, -6, and -7 [21]. The protein casein kinase II (CK2) appears to be associated with ALK3 in C2C12 cells and is released upon stimulation with recombinant BMP2. Genetic and pharmacological inhibition of membrane casein kinase II (CK2) augmented BMP2-induced transcriptional activity, suggesting a functional role for CK2 on BMP signalling [22]. Similarly, an siRNA-based screening using a luciferase reporter line identified the membrane proteins fragile histidine triad (FHIT) [23], lymphocyte cell-specific protein tyrosine kinase (LCK), Src-like Kinase (FYN) [24], and protein tyrosine phosphatase non-receptor type 1 (PTNP1) [25] as contributors to BMP transcriptional responses. Cysteine-rich motor neuron 1 protein (CRIM1) is a glycosylated transmembrane protein that interferes with BMP ligand maturation and extracellular expression [26]. Furthermore, BMP signalling is also regulated through interaction of BMPs with the ECM, thereby restricting the local BMP bioavailability. BMPs interact, e.g., with fibronectin [27,28], collagens [29,30], fibrillin [31,32], and heparins [33]. Heparin binding of some BMPs (including BMP-2) was shown to have dual activity. In some contexts, heparin binding represses [34,35], while, in other contexts, it increases the BMP signalling output [36,37]. Finally, there are soluble BMP antagonists like noggin, chordin, the gremlins, and the follistatins that can sequester and shield the ligands, in order to avoid their interaction with the extracellular domain of the receptors [38].
The association between the ligand dimer and a type I receptor dimer favours the recruitment of the type II receptors, thereby forming a heterotetrameric receptor complex (Figure 1). There are three types of BMP type II kinase receptors, named ActRIIA, ActRIIB, and BMPR2 (Figure 1). Unlike the type I receptors, the BMP type II receptors are constitutively active and their ligand-induced incorporation into the receptor complex leads to the trans-phosphorylation of the type I receptors in their cytosolic Glycine Serine (GS) domain. This phosphorylation provokes a conformational change within the cytosolic domain of the BMPRI and displaces the intracellular receptor inhibitor FK506 binding protein 1a (FKBP12). This, in turn, allows for the interaction of the cytosolic kinase domain with the BMP intracellular effectors, the different “mothers against decapentaplegic homolog” (SMAD) proteins. Upon interaction with the BMPRI-kinase domain, the receptor-regulated SMADS (R-SMADS) 1/5/8 (SMAD8, also known as SMAD9) are transphosphorylated, and thereby activated, in their C-terminal domain. The activation of the TGF-β type I receptors (ALK4/5/7) and their downstream signalling pathway, including the phosphorylation of a different subset of R-SMADS, SMADS 2/3, exhibit some peculiarities. Due to the space limitation, we will not further expound on TGF-β specific regulation, signalling, crosstalk, and particular target genes and, instead, recommend reading [39]. Inhibition of TGFβR internalization hampers R-SMADS phosphorylation. Two major proteins are responsible for the endocytosis-mediated internalization of ligand-bound membrane receptors. While SARA [40] mediates ALK4/5/7 endocytosis, endofin negatively modulates BMP signals through receptor dephosphorylation [41].
Once phosphorylated, the R-SMADS interact with a common SMAD (Co-SMAD or SMAD4) to form a heterotrimer that translocates into the nucleus (Figure 1). SMAD4 is shared by both BMP and TGF-β R-SMADs; therefore, the amount of free SMAD4 molecules may limit R-SMAD signalling. Once in the nucleus, SMAD complexes associate with further DNA binding (co-)factors to regulate the expression of specific genes. Classical BMP target genes are the differentiation and proliferation associated inhibitor-of-differentiation (ID) factors (ID1, ID2 and ID3) or genes involved in osteogenic differentiation (i.e., SP7/Osterix, RUNX2). SMAD6 is also a BMP signalling target gene. SMAD6 and SMAD7 are inhibitory SMADs (I-SMADs), which serve as negative feedback loops in the pathway. SMAD6/7 facilitate the proteosomal-mediated degradation of BMPR, and also compete with the R-SMAD for BMPR binding [42,43,44,45].
In addition to SMAD activation, BMPRs also modulate the activation of so-called non-canonical pathways (Figure 1). These are less well characterized and involve different branches of mitogen activated protein kinases (MAPK) pathways, like p38, the extracellular signal-regulated kinase (ERK), c-Jun N-terminal Kinase (JNK), phosphoinositide-3-kinase (PI3K), mammalian target of rapamycin (mTOR), glycogen synthase kinase 3 beta (GSK3-β), and small Rho GTPases in the Rho/Rac subfamily such as CDC42, as well as the LIMK/cofilin pathway, amongst others [46,47,48] (Figure 1). It was shown that, e.g., activation of BMP non-canonical pathways can result in direct non-transcriptional responses of mesenchymal progenitors, e.g., those required for cytoskeletal rearrangements and cell migration [49]. Unlike SMAD signalling, which is strong and immediate after BMPR activation, non-canonical pathways are often weakly induced by BMP ligands, and the activity can be masked by other environmental cues (i.e., inflammatory cytokines leading to inflammatory nuclear factor (NF)-kB signalling, fibroblast growth factor (FGF) signalling, vascular endothelial growth factor (VEGF) signalling, Wnt signalling, or hypoxia-HIF1α signalling). Importantly, the transcriptional output of SMAD activation can be fine-tuned by non-canonical BMPR pathways and through further crosstalk with other signalling cascades [50,51]. As such, MAPK, ERK, and JNK can crosstalk to the canonical SMAD signalling by directly phosphorylating SMADs in the linker region, thereby competing with the phosphorylation at the C-terminus domain by the BMPRs. Moreover, many other input-dependent transcription factors partner with SMAD complexes, thereby determining the binding affinity and specificity of SMAD complexes to certain target genes [52]. Therefore, the effect of BMP ligands may be dependent on multifaceted cues and factors that eventually trigger a cell-type or organ-specific response.
Figure 1. BMP signal transduction. The interaction between BMP ligands and BMP type I receptors induces the formation of membrane heterotetrameric complexes. This interaction is influenced by the relative availability of ligands and receptors, which is regulated by soluble antagonists, co-receptors, and extracellular matrix (ECM) components. The activated BMP receptors transmit signals intracellularly through their serine/threonine kinase domain, leading to the phosphorylation of the BMP canonical mediators, SMAD1/5/9, at the C-terminal domain. Moreover, BMP receptors regulate the activation of non-canonical signalling pathways, which serve as a signalling hub integrating tissue-specific environmental cues. Canonical and non-canonical signalling pathways intersect at various levels. Once activated, SMAD1/5/9 trimerize with SMAD4 and translocate into the nucleus to bind specific gene promoters in collaboration with transcription co-factors. A number of environmental cues co-influence signalling outcomes, including cellular mechanics such as mechanical shear forces or inflammatory or hypoxic microenvironments. Due to space limitations, details on the TGF-β-specific regulation of the signalling branch are not shown.
Figure 1. BMP signal transduction. The interaction between BMP ligands and BMP type I receptors induces the formation of membrane heterotetrameric complexes. This interaction is influenced by the relative availability of ligands and receptors, which is regulated by soluble antagonists, co-receptors, and extracellular matrix (ECM) components. The activated BMP receptors transmit signals intracellularly through their serine/threonine kinase domain, leading to the phosphorylation of the BMP canonical mediators, SMAD1/5/9, at the C-terminal domain. Moreover, BMP receptors regulate the activation of non-canonical signalling pathways, which serve as a signalling hub integrating tissue-specific environmental cues. Canonical and non-canonical signalling pathways intersect at various levels. Once activated, SMAD1/5/9 trimerize with SMAD4 and translocate into the nucleus to bind specific gene promoters in collaboration with transcription co-factors. A number of environmental cues co-influence signalling outcomes, including cellular mechanics such as mechanical shear forces or inflammatory or hypoxic microenvironments. Due to space limitations, details on the TGF-β-specific regulation of the signalling branch are not shown.
Cells 12 02200 g001

1.3. Cell-Type Specific BMP Activity

Physiologically, almost all BMP ligands have been found to circulate systemically at functional levels, and some BMP receptors are abundantly expressed in a plethora of cell types. However, the expression of some BMPs is restricted to specific tissues. For example, BMP-9 is mostly expressed in the liver [53], while BMP-10 expression is restricted to the right cardiac atrium in adulthood [54,55]. The localization of BMP-2 transcripts was observed in various areas, such as the developing heart, whisker follicles, tooth bud, and, as expected, in the limb bud. BMP-4 transcripts, on the other hand, were detected in limb buds, nervous tissue, and craniofacial tissues through in situ hybridization studies [56]. BMP-7 is mainly expressed in the kidney [57]. Once in circulation, BMP ligands show different affinity for BMPRs; therefore, the relative concentration of ligands and expression levels of the BMPRs can determine the activation of signalling. While ALK2 and ALK3 are broadly expressed, ALK1 expression is predominantly found in endothelial cells (ECs) and, at some level, in chondrocytes [16] and pulmonary vascular smooth muscle cells (VSMCs) [58]. ALK6 is highly expressed in ovary [59] and prostate tissue [60]. Moreover, BMP ligands are promiscuous and can bind different receptor complexes depending on their availability. BMP-9, for example, can bind to ALK1 and ALK2, although the affinity for ALK1 is much higher (EC50 = 50 pg/mL for ALK1 and 50 ng/mL for ALK2) [61]. The affinity can be severely influenced by type III receptors or co-receptors, such as betaglycan, which lack a cytosolic domain with enzymatic activity but can shield or present ligands to type I and type II receptors. The expression levels of co-receptors also differ among cell types. Endoglin is highly expressed in embryonic tissues and stem cell-like cells in adulthood, as well as in ECs and few subsets of immune cells. Furthermore, environmental cues, like inflammation, can influence the expression of BMPRs. As such, BMPR2 is inhibited at different levels by inflammatory interleukin (IL)-6 and tumour necrosis factor (TNF)-α signalling, for example [62,63,64], thereby disturbing BMP signalling under inflammatory microenvironmental conditions [63]. Hypoxia, for instance in highly condensed chondrocytes, extends the duration of SMAD1/5/9 activation. Mechanical forces also influence BMP activity (Figure 1). Shear stress can tune the endoglin co-receptor incorporation into BMPR complexes in ECs and sensitize them to low BMP-9 concentrations [65,66], and mechanical loading can promote the autocrine activity of ligands such as BMP-2 and influence BMP target gene regulation [58,67]. Moreover, culturing cells on a soft vs. a stiff matrix influences osteogenic BMP signalling, endocytosis of BMPRs, and incorporation of transcriptional co-factors into the SMAD signalling complex [68,69]. The special role of cell mechanics for the modelling of BMP rare diseases by iPSCs will be discussed in Section 3.
Although the pathways by which BMP ligands induce downstream signal activation have been roughly characterized in the previous decades, it is only now, when using more complex human cell models, that tissue-specific determinants of BMP signalling are starting to be unveiled. This highlights the need for studying BMP signalling in the right cellular and microenvironmental context, resembling, when possible, the cellular, biochemical, and biomechanical conditions which drive the onset and development of BMP related human diseases.

1.4. BMPs in iPSC Stemness and Differentiation

During the early embryonic development of vertebrates, gradients of BMP ligands and inhibitors pattern the dorsal–ventral embryonic axis [70,71]. BMP signalling specifies the ventral mesoderm [72,73] and affects neural tube formation [74], intestines (reviewed in [75]), liver (reviewed in [76]), and kidney development [77]. Importantly, BMP signalling becomes the key to maintaining stemness in embryonic stem cells, germline stem cells, hematopoietic stem cells, and intestinal stem cells in a species-dependent manner (reviewed in [78]). In mouse embryonic stem cells (mESCs), BMPs control self-renewal and promote stemness as they prevent ectoderm differentiation [79,80]. In human ES cells (hESCs), BMPs also inhibit neural differentiation but are mostly known as potent inducers of mesoderm formation. The function of BMP signalling might be different in iPSCs, where BMP signals seem to not be required for the maintenance of human iPSC stemness, while Activin-A signalling is involved in maintenance of human iPSC pluripotency [81]. Yamanaka and Conklin have shown that BMP-SMAD-ID signalling promotes episomal vector mediated reprogramming of differentiated somatic cells to pluripotency, since adding BMP-4 during the early phase of reprogramming inhibited p16/INK4A-dependent senescence, a major barrier to efficient reprogramming [81] (Figure 2). Typically, the efficacy of iPSC reprogramming is quite low (less than 1% of human primary somatic cells that have received reprogramming factors turn into iPSCs) [82]. Upon successful reprogramming, most iPSC differentiation protocols describe short-term BMP treatment if mesoderm induction and differentiation of iPSCs towards a definitive endoderm is desired. For ectodermal differentiation, BMP activity needs to be strictly repressed initially, and only at later ectodermal specification, e.g., towards skin progenitors or keratinocytes, is BMP signalling again required [82,83] (Figure 2). The most used TGF-β superfamily ligands found in human iPSC mesodermal or definite endodermal differentiation protocols are Activin-A [84,85], BMP-4 [86,87,88], BMP-2 [89,90], and BMP-7 [91,92]. Few other established iPSC differentiation protocols investigated the potency of other, less common, BMP ligands such as BMP-10. The latter is particularly interesting since it can induce human iPSC differentiation towards extraembryonic trophoblast tissue [93]. It is important to note that concentrations of different BMP ligands can vary between published protocols, even when the targeted cell differentiation is the same. For instance, in iPSC-derived EC differentiation, which often includes BMP-4-dependent mesoderm induction, one can find a comparison of different protocols in [94]. Importantly, BMP stimulation alone is not sufficient to induce iPSC differentiation towards specific cell types. Therefore, BMPs are usually combined with other growth factors (e.g., FGF, Wnt, or VEGF) (Figure 1) and/or pathway inhibitors, such as small molecule inhibitors targeting TGF-β type I receptors ALK4/5/7 or Wnt/GSK3-β signalling (SB431542 and CHIR99021, respectively) (Figure 2). We will herein limit our review on the most important mesodermal-derived cell types derived from iPSCs, which are relevant for the BMP rare diseases FOP, PAH, and HHT. These include fibroblasts, macrophages, fibro-adipogenic progenitors (FAPs), muscle-residing satellite cells, skeletal myoblasts and myocytes, osteoblasts and osteocytes, chondroblasts and chondrocytes, and cell types forming the vasculature, including ECs and mural cells such as VSMCs and pericytes.

2. iPSC Technology and BMP-Related Rare Diseases

BMP-related rare diseases affect a small number of individuals, and there are currently a limited number of clinical treatment strategies or interventions available. Mutations in genes encoding for BMP pathway proteins alter either their expression or protein function. Three conditions have garnered significant research attention in efforts to comprehend the genotype–phenotype relationship. These are Fibrodysplasia ossificans progressiva (FOP), Pulmonary arterial hypertension (PAH) and Hereditary haemorrhagic telangiectasia (HHT). However, other BMP rare diseases have been described. Diffuse intrinsic pontine glioma (DIPG) refers to a devastating form of brainstem glioma with highly proliferative and micro-vascularized solid tumours [95], often arising during childhood. These tumours are often inoperable and exhibit a very poor survival rate. Over 80% of tumours have K27M mutations in H3.3 (H3F3A) and H3.1 (HIST1H3B) genes, leading to reduced H3K27 trimethylation. DIPG exhibits heterogeneity due to additional somatic mutations, including TP53 inactivation in 40–50% of patients. Notably, ACVR1 (ALK2) somatic mutations occur in approximately 33% of DIPGs, sharing similarities with those FOP mutations described below [96,97]. Three additional mutations (L343P, R307L, and H286N) in the cytosolic domain of ALK2 have been linked with human diseases. While L343P and R307L were linked to defects in the atrioventricular septum, H286N was found in one patient with Down syndrome and congenital heart disease (CHD). CHD was also associated with several loss of function mutations in BMPR1A (ALK3) [98,99]. Germline mutations in ALK3 leading to truncated protein receptors have been traditionally associated with juvenile polyposis syndrome (JPS), which is an autosomal dominant gastrointestinal disorder that predisposes for gastrointestinal cancers [100]. Previously, JPS was linked to mutations in MADH4 (encoding SMAD4) [101]. Finally, inactivating mutations in BMPR1B (ALK6) result in uncommon inherited skeletal disorders, named acromesomelic dysplasias (AMD), which are characterized by short stature, extremely short limbs, and deformities in the hands and feet [102,103]. To the best of our knowledge, iPSC technology has been applied so far to the fields of FOP, PAH, and HHT research (Figure 2). Therefore, in the subsequent paragraphs, we will summarize the implementation of human iPSCs as model systems to investigate these three BMP rare diseases.

2.1. Fibrodysplasia Ossificans Progressiva and hiPSCs

Heterozygous germline gene mutations in the ACVR1 gene were found in all patients diagnosed with Fibrodysplasia ossificans progressiva (FOP, OMIM 135000) [104], while approximately 97% of all patients share the exact gene point mutation c.617G>A; (p.R206H) in the chromosome 2 (Table 1). The gene ACVR1 encodes for the protein ALK2, a BMP type I receptor, which classically mediates signal transduction of osteogenic BMP ligands like BMP-5, BMP-6, and BMP-7 [13] (Figure 3). Despite the fact that ALK2 is broadly expressed, the main clinical manifestation of FOP is limited to connective tissues within ligaments, fascia, tendons, and joints. Affected individuals are recognized by congenital skeletal malformations (i.e., great toe deformity or hallux valgus) and postnatal episodic heterotopic ossification (HO) (Figure 3). During the first decade of life, sporadic episodes of painful soft tissue swellings (“flare-ups”) occur often following soft tissue injury, intramuscular injections, viral infection, muscular stretching, falls, or fatigue [105]. These events usually result in the formation of extraskeletal bone plaques of endochondral origin caused by the differentiation of existing fibroadipogenic progenitor cells (FAPs) into cartilage first, and, eventually, into mature bone [106]. Other, often overlooked, features of FOP include cardiac abnormalities in new-borns, such as ventricular septal hypertrophy and cardiac conduction abnormalities in children [107,108,109,110,111,112]. The blood vessels in FOP lesions are over-represented in comparison to non-genetic forms of HO, exhibit weak junctions, and are prone to oedema [110,113], suggesting additional effects of ALK2R206H in cell types other than bone progenitor cells.
The first reported FOP iPSC line was established upon genetic reprogramming of skin fibroblasts using Sendai Virus (SeV) particles carrying OCT3/4, SOX2, KLF4, and MYC, and maintained on mitomycin-C-treated mouse embryonic fibroblast (MEF) feeder cells [114] (Table 1). Remarkably, the authors found a substantial drop in reprogramming efficiency in FOP iPSCs from donors carrying the ALK2 mutations R206H and G356D when compared with non-isogenic control cells. This finding was unexpected, as it contradicted the previously mentioned observations of enhanced reprogramming efficacy following BMP-SMAD-ID activation [82]. Here, upon incubation with the ALK2-biased small molecule kinase inhibitor LDN-193189, the efficiency was increased in a dose-dependent manner (Figure 2), suggesting that the kinase activity of ALK2 may compromise pluripotent transformation. Congruently, the expression of mesodermal and endodermal gene markers was upregulated in FOP iPSC not treated with LDN-193189, whereas the expression of the neuroectodermal markers SOX1 and NESTIN remained unaffected. In conclusion, the generation of iPSCs from FOP patients’ somatic cells may require inhibition of mutant ALK2 activity by adding small molecule inhibitors, such as ALK2 inhibitors like LDN-193189 or dorsomorphin to maintain pluripotency (Figure 2). In 2013, Matsumoto et al. described the generation of iPSCs from human dermal fibroblasts from five FOP donors and six control donors using two different reprogramming methods [115] (Table 1). The authors explored the use of retrovirus and episomal vectors carrying OCT4, SOX2, KLF4, and MYC to generate iPSCs maintained on Matrigel or feeder cells. Unlike the previous report, Matsumoto et al. did not find any differential efficiency in the generation of iPSCs from FOP donors compared to control cells, and ALK2 kinase inhibition was not required for a successful reprogramming. Interestingly, FOP iPSCs demonstrated augmented mineralization and chondrogenesis in vitro, a phenomenon that could be mitigated in the presence of LDN-193189. Cai et al., reprogrammed urine renal epithelial cells from control and FOP individuals, cultured in serum-free medium supplemented with small molecules CHIR-99021, PD-0325901, A83-01, and thiazovivin. Transient expression of OCT4, SOX2, KLF4, and the pCEP4-miR-302-367 cluster was achieved through electroporation with episomal vectors [116]. In this case, the reprogramming of FOP iPSCs was also not less efficient than in control cells. Moreover, the authors followed established protocols to differentiate control and FOP iPSCs into pericytes and ECs [117,118] (Table 1). Interestingly, FOP derived ECs were less proliferative and stable, and exhibited lower VEGFR2 expression compared with non-isogenic control cells. The expression of endothelial-specific markers was reduced, whilst fibroblast genes were upregulated in FOP cells, suggesting that FOP ECs were prone to undergo a developmental dedifferentiation process termed endothelial-to-mesenchymal transition (EndMT) [119]. Through EndMT, ECs progressively transform into multipotent MSC-like cells and may give rise to osteogenic and chondrogenic cells, at least in vitro, consistent with previous findings [112,120] (Figure 2). iECs were also investigated by Hsiao’s lab, using non-isogenic previously published iPSC lines [115] (Table 1). Unlike the previous method, in this case, EC differentiation was achieved through a protocol not including the TGF-β receptor small molecule inhibitor [121]. The expression of the EC marker VE-Cadherin (CDH5) was significantly low, whereas the mesenchymal genes FSP-1, collagen 2-alpha-1 (COL2A1), COL1A1, and ALPL were consistently elevated in FOP iECs. Despite the fact that this gene expression pattern might be suggestive of EndMT, the authors failed to obtain functional osteoblast-like mineralizing cells in vitro, either from control or from FOP iECs. Lineage tracing studies in mouse models of FOP have pointed at a population of FAP cells defined as Tie2+, CD31, CD45, PDGFRα+, SCA1+ as the independent mediator of HO in FOP [122]. FAPs may be responsible for both congenital and postnatal bone abnormalities in patients, but ubiquitous gene expression of ALK2R206H is embryonically lethal in mice [122,123,124]. Therefore, iPSC technology has been applied to FOP in order to model potential HO progenitor cells in FOP (Figure 3). In the elegant work by Nakajima et al. [125], the authors explore the chondrogenic potential of different somite derivatives obtained from isogenic control and FOP iPSCs. FOP iPSC-derived pre-somites differentiated towards MSCs were able to form chondrogenic cartilage-like structures in vitro in response to Activin-A. However, differentiation towards sclerotome resulted in cells that could successfully form chondrogenic 3D pellets in vitro, although no neo-functional response to Activin was observed in FOP cells. In FOP mice, however, tendon resident scleraxis (Scx) positive cells have been proposed as Activin-responsive sporadic bone progenitor cells [124]. This might suggest that embryonic ALK2R206H expression may impact differently the cell fate of mouse and human bone progenitor cells. Muscle regeneration in FOP was recently investigated by applying iPSC technology [126]. The authors aimed to recapitulate observations in tissues from cadaveric specimens using non-isogenic control and FOP iPSCs differentiated into myogenic progenitors combining small molecules (i.e., CHIR-99021, LDN-193189) and growth factors (i.e., bFGF). The mRNA transcriptome of primary human satellite cells (Hu-MuSCs) and iPSC-derived muscle stem/progenitor cells (iMPCs), from control and FOP donors, unveiled that iMPCs do express a more progenitor-like phenotype as compared to Hu-MuSCs (Table 1). While the FOP Hu-MuSC engraftment and regeneration capacity was compromised, this was not the case for the control and FOP iMPCs, which performed similarly. In vitro, both FOP iMPCs and Hu-MuSCs expressed high levels of the ALK2 target genes ID1 and ID3, as well as chondrogenic and ECM genes. However, neither iMPC nor Hu-MuSC implants formed HO in vivo, suggesting that other cell types are the ultimate bone mediators in FOP.
Since inflammation triggers bone formation in FOP, recent advances in iPSC differentiation protocols have enabled the study of the impact of ALK2R206H on monocyte activation and cytokine secretion. Matsuo et al. utilized two distinct macrophage differentiation protocols from iPSCs, both in 2D and 3D formats [127]. Although a direct comparison between control and FOP cells was not possible in the new 3D embryoid body-based protocol, the 2D differentiation method, using a commercially available kit, yielded a heterogeneous population of M1/M2 macrophages. These cells could be efficiently polarized into M1 or M2 states using specific stimuli (Table 1). Notable differences were observed in FOP iPSC-derived M1 macrophages, which exhibited higher expression levels of IL6, IL1a, RANTES, and Activin-A, among others. Maekawa et al. pursued a different approach by differentiating isogenic iPSC lines into the monocytic lineage and subsequently immortalizing the cells [128,129]. FOP immortalized monocytes (FOP-ML) displayed elevated levels of phosphorylated SMAD5, both at basal levels and upon stimulation with recombinant Activin-A, unlike the control cells. However, neither study investigated the potential differentiation of monocytes into osteoclasts, which play a crucial role in bone remodelling.
The implementation of iPSC technology into FOP research has led to major advances in drug development for this so far uncurable disease. Experiments performed in isogenic control and FOP-derived mesenchymal stromal cells (MSC) unveiled what seems to be the underlying mechanism driving FOP. Using a BMP transcriptional reporter assay, Hino et al. screened for most of the TGF-β superfamily ligands to discover that recombinant Activin-A induced BMP signalling and downstream chondrogenic differentiation exclusively in FOP cells [130]. Almost simultaneously, these findings were reported in humanized murine embryonic stem cells from a newly developed FOP model, where pharmacological sequestration of circulating Activin effectively prevented extraskeletal bone formation [123]. Currently, Garetosmab (a humanized anti-Activin antibody) is being clinically evaluated for its potential to prevent HO in individuals with FOP (clinical trial NCT05394116). In a different study, FOP iPSCs, incorporating a novel reporter construct based on the COL2A1 promoter and the aggrecan enhancer driving luciferase expression, were combined with drug-repurposing chemical libraries. Screening nearly 7000 compounds identified the potential therapeutic value of targeting mTOR signalling in FOP, particularly with rapamycin [131,132]. These findings have led to the initiation of the first clinical trial based on iPSC-derived cells (Japan, UMIN000028429) investigating the repositioning of rapamycin for FOP. A more recent clinical trial aims to evaluate the application of Saracatinib in FOP (clinical trial NCT04307953). Identified through the same screening unveiling rapamycin [132], Saracatinib has been repurposed as a selective ALK2 inhibitor, preventing Activin neo-function in iPSC-ECs [133]. This and other manuscripts have found how aberrant Activin signalling in FOP patients (Figure 3) is recapitulated not only in FAPs, but also in other cell types such as dermal fibroblasts, urine epithelial cells, blood-derived monocytic and endothelial cells, and periodontal ligament fibroblasts [112,116,134,135,136,137,138]. Using different primary or iPSC-derived cell types may facilitate the investigation in organ-specific Activin mechanisms, which is useful for characterizing novel druggable targets and/or evaluating the potential toxicity of drugs under investigation.

2.2. Hereditary Pulmonary Arterial Hypertension and hiPSCs

Pulmonary arterial hypertension (PAH) is a rare disease characterized by the narrowing of arterial blood vessels that carry deoxygenated blood from the right side of the heart to the lungs. PAH has been linked with gene mutations in the BMP signal transduction pathway (e.g., ACVRL1(ALK1), BMPR1B (ALK6), ENG (endoglin), and SMAD9), but also gene mutations in genes outside of the BMP pathway have been found among those patients with and without familial forms of PAH (i.e., CAV1, KCNK3, EIF2AK4). However, BMPR2 mutations are predominant, accounting for 70–80% of all cases of heritable PAH (HPAH, formerly familial PAH) [139,140,141,142,143,144] (Figure 3). Different HPAH-causing variants, mapping in different parts of the BMPR2 gene (including its promoter region), have been described, resulting in a loss of BMPR2 expression, abnormal trafficking [145], mislocalization [146], and/or a haploinsufficient phenotype [147,148]. Moreover, some mutations give rise to dominant negative BMPR2 proteins due to mechanisms escaping nonsense mediated decay [149,150,151]. BMPR2 expression is ubiquitous, but high affinity complexes predominantly with ALK1 are formed mainly in the vasculature. ALK1 cooperates with the co-receptor endoglin, strongly expressed in the endothelium [152,153], to further enhance ligand binding affinity and receptor downstream signalling [154,155]. Moreover, endoglin is required to integrate extracellular mechanical forces, such as shear stress exerted by blood flow, into the BMP signalling pathways [65,66,156,157]. In ECs, BMP-9/-10 induce SMAD1/5/9 signalling and expression of vascular cell specific target genes [158], through ALK1-ENG-BMPR2, in order to maintain EC quiescence and expression of endothelial specific genes. It is therefore not surprising that next to BMPR2 mutations, SMAD1, SMAD9, ACVRL1 (ALK1), and ENG, mutations can also drive HPAH (reviewed in [140]). The latter two (ACVRL1 and ENG mutations) are the main heritable risk factors for HHT, which will be covered in the next paragraph. For simplicity reasons, we will focus on BMPR2 mutations and HPAH (Figure 3).
BMP-9 and BMP-10 are blood circulating factors. Gene regulation by BMP-9/-10 signalling provides the endothelium with survival and quiescence signals [159,160], as well as inhibiting vascular permeability [161]. This is particularly important under challenging environmental conditions of the vasculature such as hyperglycaemia [161], hypoxia [162], or inflammation [163], which are all known to induce hyperpermeability of the vasculature. BMPRs interact with a number of vascular-specific transmembrane proteins in arterial and venous blood vessels and lymphatics, including VEGFR2 and neuropilins (NRP1/2), reviewed in [164]. Furthermore, BMP-9/-10-BMPR2-ALK1 signalling protects ECs from undergoing EndMT, which involves EC activation and extensive transcriptional reprogramming, and leads to a shift of mature ECs towards a mesenchymal phenotype [119,165,166] (Figure 3). ECs undergoing EndMT can adopt a myofibroblast-like state where cells tend to deposit aberrant fibroblast-like ECM, which changes their own and the environmental mechanical features [167,168,169,170]. We will further expand on these aspects in Section 3.1. In smooth muscle cells (SMCs), BMPR2 signalling protects their hyperproliferative growth responses towards TGF-β [171,172,173] and maintains their contractile state [58]. The major cellular consequences of HPAH are found in ECs and SMCs of the pulmonary artery vasculature, where disruption of BMP-9/-10-BMPR2-ALK1-endoglin signalling causes, as a consequence of the beforementioned cellular functions, EC apoptosis, EndMT, and SMC hyperproliferation. This results in the narrowing of the pulmonary arterial lumen of the arterial wall, through a process called neointima formation, and the increase in vessel wall thickness, through a process called medial hypertrophy (Figure 3). The resulting increase in lung blood pressure not only has negative effects for the perfusion of the lung itself but also for the right heart chamber coping with the increased arterial resistance. Unfortunately, most rodent models of HPAH do not precisely recapitulate the disease pathology; these models display less substantial pulmonary vascular remodelling in both proximal arteries and distal microvasculature, which significantly hampers drug discovery efforts [174,175,176]. Presently, there are number of drugs that have been approved by the Food and Drug Administration (FDA) for the treatment of HPAH and its idiopathic form (IPAH), whilst most address only the clinical symptoms limited to vasodilator treatment. No FDA approved PAH drug is currently able to reverse the described vascular remodelling. Recent approaches discovered that the use of the immunosuppressive drug FK506 is able to inhibit the interaction between the BMP type I receptor kinase domain and the endogenous intracellular BMP signalling inhibitor FKBP12 [177]. More recent advancements have focused on targeting HPAH by addressing the elevated levels of Activin-A using the Activin-ligand trap Sotatercept/ActRIIa-Fc [63,178,179]. This progress in drug development was partly due to the use of iPSC technology.
Figure 3. Rare BMP diseases and iPSCs. Fibrodysplasia ossificans progressiva (FOP) is a musculoskeletal disease caused by heterozygous gain-of-function mutations in ACVR1/ALK2, resulting in heightened activation of downstream ALK2 signalling in response to Activin-A. Increased ALK2 signalling leads to the formation of extraskeletal bone plaques. On the right is a photograph of the skeleton of a man with FOP, from the collections of the Anatomical Museum of the Leiden University Medical Center (LUMC). This photo has been reproduced with permission from LUMC (copyright belongs to LUMC). Hereditary pulmonary arterial hypertension (HPAH) is a cardiovascular disease characterized by progressive narrowing of the pulmonary arteries, resulting in elevated pulmonary arterial pressure (mPAP > 20 mm Hg), followed by right ventricular dilatation and hypertrophy. Loss-of-function mutations in BMPR2 exacerbate disease severity and penetrance by disrupting the overall balance of TGF-β signalling in endothelial and smooth muscle cells. The right panel illustrates a cross-section of a healthy pulmonary artery (top) compared with a diseased occluded pulmonary artery (bottom). The formation of a pronounced neointima by aberrant cell growth into the lumen is shown. The PAH patient pulmonary arterial cross-section is part of a published dataset by C.H., recreated based on details mentioned in Hiepen C. et al. [168]. Hereditary haemorrhagic telangiectasia (HHT) is associated with loss-of-function gene mutations in ACVRL1 (ALK1) and ENG (endoglin), among others, which impair BMP signalling. It is characterized by varying degrees of lesions due to the instability of the microvascular networks. The right panel displays a typical microvascular subdermal lesion seen in HHT. The picture was contributed with friendly permission by Prof. Dr. Urban Geisthoff (Philipps University of Marburg; Germany).
Figure 3. Rare BMP diseases and iPSCs. Fibrodysplasia ossificans progressiva (FOP) is a musculoskeletal disease caused by heterozygous gain-of-function mutations in ACVR1/ALK2, resulting in heightened activation of downstream ALK2 signalling in response to Activin-A. Increased ALK2 signalling leads to the formation of extraskeletal bone plaques. On the right is a photograph of the skeleton of a man with FOP, from the collections of the Anatomical Museum of the Leiden University Medical Center (LUMC). This photo has been reproduced with permission from LUMC (copyright belongs to LUMC). Hereditary pulmonary arterial hypertension (HPAH) is a cardiovascular disease characterized by progressive narrowing of the pulmonary arteries, resulting in elevated pulmonary arterial pressure (mPAP > 20 mm Hg), followed by right ventricular dilatation and hypertrophy. Loss-of-function mutations in BMPR2 exacerbate disease severity and penetrance by disrupting the overall balance of TGF-β signalling in endothelial and smooth muscle cells. The right panel illustrates a cross-section of a healthy pulmonary artery (top) compared with a diseased occluded pulmonary artery (bottom). The formation of a pronounced neointima by aberrant cell growth into the lumen is shown. The PAH patient pulmonary arterial cross-section is part of a published dataset by C.H., recreated based on details mentioned in Hiepen C. et al. [168]. Hereditary haemorrhagic telangiectasia (HHT) is associated with loss-of-function gene mutations in ACVRL1 (ALK1) and ENG (endoglin), among others, which impair BMP signalling. It is characterized by varying degrees of lesions due to the instability of the microvascular networks. The right panel displays a typical microvascular subdermal lesion seen in HHT. The picture was contributed with friendly permission by Prof. Dr. Urban Geisthoff (Philipps University of Marburg; Germany).
Cells 12 02200 g003
For HAPH disease modelling and drug development, human ECs that are typically isolated from umbilical-cord vein (HUVECs) may not represent the appropriate test system, since they are juvenile, form rather large veins, and have not experienced the same oxidative and mechanical environment as pulmonary arteries. Moreover, their access is limited, the cost of commercial HUVECs is high, and batch-to-batch reproducibility is often too low. iPSC-derived cell types can serve as an unlimited source for surrogates for different EC types, including veinous ECs, arterial ECs, (brain) microvascular ECs, and lymphatic ECs, in both functional and drug discovery studies [180,181,182]. The first HPAH-patient-derived iPSCs were generated in 2012 at the University of Cambridge by Geti, Vallier, Morrell, and co-workers [183] (Table 1). They used late-outgrowth endothelial progenitor cells (L-EPCs) from peripheral blood mononuclear cells (PBMNC) as a somatic source for iPSC reprogramming. EPCs possess several favourable characteristics as a somatic cell source for the generation of iPSCs over, e.g., dermal fibroblasts. Although skin fibroblasts are the most common cell type used for generating iPSCs, their isolation requires a surgical procedure, which is also undesirable in children, patients with skin disorders, and patients with abnormal coagulation or wound healing. In addition, fibroblasts are reprogrammed with relatively low efficacy. EPCs are also readily generated from frozen PBMNC preparations, making storage and transportation of somatic source material straightforward. As a proof-of-principle, they generated EPC-iPSCs from both healthy individuals and patients with heritable BMPR2 mutation and idiopathic PAH. Compared with the traditional approach using fibroblasts, EPCs displayed high reprogramming kinetics and efficiency, taking only 10 days to emerge in culture and form iPSC colonies. This first generation of PAH-derived iPSC was followed by the first directed differentiation approach, conducted at Vanderbilt University by West et al. in 2014 [174], in which they applied mesenchymal and EC differentiation protocols combined with a transcriptomic profiling of iPSCs and differentiated cells, respectively (Table 1). The BMPR2 mutation resulted in changes in genes that were consistently altered by the mutation, irrespective of the cell differentiation state. These included genes, e.g., associated with Wnt signalling. However, the authors also identified genes that were exclusively differentially expressed in mutated cells only upon their differentiation into specific lineages, such as secreted frizzled-related protein 2 [174]. This study was followed by work from the Rabinovitch lab in Stanford in 2017, aiming at identifying endogenous mechanisms that may revert the impact of the BMPR2 mutation [184,185]. This is particularly important because only 20% of the BMPR2 HPAH mutation carriers develop clinical symptoms, and it is not known why. For this, Gu et al. compared the transcriptomes of iPSC-derived ECs from unaffected BMPR2 mutation carriers with those from HPAH patients with clinical symptoms. The same study also used CRISPR/Cas9 for BMPR2 gene correction in patient-derived cells for the first time, generating isogenic controls [184]. Upon differentiation, the corrected HPAH iPSC-EC lines showed levels of cell adhesion molecules and survival rates in response to either serum withdrawal or after hypoxia and reoxygenation, which were similar to control cells (Table 1). In accordance with many other studies, they could show, e.g., that, in iPSC-derived ECs from HPAH patients, reduced BMPR2 expression led to impaired canonical BMPR2 signalling (pSMAD1/5-ID1) in response to BMP treatment (Figure 3). The RNA-sequencing dataset from this study is valuable and continues to be used in the identification of new BMPR2 mutation-associated mechanisms [186]. Sa et al. compared pulmonary arterial (PA) ECs and iPSC-derived EC from the same HPAH patients, since several studies suggest that PA cell-specific features (such as differential chromatin methylation patterns in the endothelial nitric oxide synthase genes) contribute to the development of PAH [185]. They also generated PAEC and iPSC-derived ECs from patients suffering from idiopathic PAH, in order to find communalities between IPAH and HPAH and to develop personalized medicine approaches. One of their major findings was that IPAH, HPAH-PAECs, and iPSC-derived ECs from the same donors exhibit similar reduced gene expression of collagen type IV (COL4A2), a major ECM protein of the vascular wall basal lamina. Subsequently, Gu et al. conducted an extensive study utilizing patient-derived iPSCs, followed by endothelial cell (EC) differentiation, to identify potential new drug candidates. This was accomplished through high-throughput drug screening combined with in silico analysis of existing transcriptomic datasets. They identified a promising lead compound named AG1296 (Tyrphostin) [187]. Tyrphostin is a tyrosine kinase inhibitor targeting platelet-derived growth factor (PDGF), c-Kit, and FGF receptor signalling, and it has potent effects on cells nanomechanical properties and cytoskeletal organization [188,189,190]. Kiskin et al. has recently generated iPSC-SMCs that recapitulated some of the important functional responses of adult-derived distal pulmonary arterial smooth muscle cells (PASMCs), as well as iPSC-ECs with enhanced expression of arterial specific markers [191] (Table 1). They obtained control and BMPR2 mutant isogenic lines using CRISPR/Cas9. This revealed, for the first time, that iPSC-derived PASMCs from HPAH patients resemble the hyperproliferative phenotype observed in primary cells. They also used those iPSC-derived PASMCs and arterial ECs to investigate the role of BMP-9 as a PAH reversing agent [162,192]. Using the same model, the role of TNFα as an important secondary driver of the PAH disease mechanism was analysed [63]. In a recent study released as preprint, HPAH patient derived iPSCs were also differentiated into vascular SMCs [193]. The study suggests that progesterone promotes the proliferation and migration of HPAH iPSCs-derived VSMCs, supporting a model that describes why penetrance for BMPR2 mutation carriers is observed with a female-sex bias [194,195,196].
Thus far, the use of hiPSCs has significantly contributed to the discovery of BMPR2 insufficiency being the major heritable risk factor for HPAH and also to the identification of underlying molecular pathomechanisms. Moreover, iPSCs from HPAH patients allowed for the identification of promising new target genes and drug candidates. Using iPSCs for stem cell therapy in PAH has also been discussed [197] (Figure 4). The first experiments in animals have suggested a therapeutic effect of injecting either iPSCs, iPSC-derived conditioned medium, or iPSC-derived exosomes in animal models of PAH [198,199]. Although further experimental validation in several preclinical models is desired, these results underline the benefits of paracrine effects in treatment of PAH.

2.3. Hereditary Haemorrhagic Telangiectasia and hiPSCs

Hereditary haemorrhagic telangiectasia (HHT) is an autosomal dominant disorder mainly affecting the microvasculature, including capillaries, venules, and arterioles. Three major genes are causally related to HHT: the ENG gene encoding the co-receptor endoglin (ENG/HHT1) [200]; the ACVRL1 gene encoding ALK1 [201,202,203] causing HHT2; and the MADH4 gene (SMAD4) (HHT3), a critical intracellular mediator in BMP and TGF-β signalling pathways [202,204] (Figure 3). GDF2 gene mutations (encoding for BMP-9) causing HHT [205] are extremely rare. For HHT2, about 350 different ACVRL1 mutations are known, including deletions, splice site mutations, and missense mutations of ACVRL1, the latter building the largest group. HHT patients exhibit characteristic severe epistaxis (nose bleeding) and haemorrhage of the subdermal microvasculature, which is called “telangiectasia” [206] (Figure 3). Other capillary beds are affected by haemorrhages and dilations, including mucosa-associated blood vessels in the nose, liver, lung, gut, and brain. Haemorrhages are due to the loss of the intervening capillary bed and the formation of dysfunctional arteriovenous malformations (also known as “shunts”) associated with a lymphocytic perivascular cell infiltrate. Those rather less harmful appearing vascular malformations can become very severe, affecting the blood flow in the brain and the heart. HHT patients often suffer from severe iron deficiency and anaemia [207]. In Acvrl1 defective HHT rodent models, it was shown that local injury triggers an abnormal re-vascularization with dilated and tortuous vessels. In mice, this might be due to excessive sprouting angiogenesis, in line with previous reports [208]. Cell–cell communication seems deficient in HHT ECs, compromising the pericyte coverage of the microvascular endothelium, which is also required for vessel stabilization and maturation [209,210]. While the role of sprouting angiogenesis to HHT development is currently under investigation, one possible explanation for increased angiogenic responses of ACVRL1 mutant ECs may be that their pericyte coverage is reduced, lacking a barrier that protects ECs from direct exposure to pro-angiogenic stimuli that induce angiogenic sprouting from pre-existing blood vessels, e.g., as observed during vascular development [211]. To date, a precise understanding of how dysfunctional BMP-SMAD1/5/9 signalling is involved in HHT is missing. Current models suggest that BMP-9/-10-ALK1-endoglin-SMAD target genes do inhibit angiogenesis [212,213] required to protect against HHT development (Figure 3). However, non-canonical BMP signalling may also contribute to the HHT pathomechanism. Moreover, there might be a link with the flow perception of ECs through endothelial primary cilia, which are flow-sensitive organelles composed of microtubules. Primary cilia act as signalling hubs that connect the mechanical forces of hemodynamic flow to canonical BMP/TGF-β family signalling [66,214,215,216]. The loss or dysfunction of endothelial primary cilia could potentially contribute to the development of arteriovenous malformations (AVMs) by affecting ALK1-SMAD signalling [217]. This only highlights the role of the local pro-angiogenic and mechanical microenvironment in the formation of arteriovenous malformations. Currently, the most effective treatments for HHT involve repurposed anti-angiogenic drugs that are already utilized in cancer therapy. These include anti-VEGF antibodies like Bevacizumab, which is a recombinant humanized monoclonal antibody that hinders blood vessel formation by inhibiting VEGF [218,219]. Additionally, several tyrosine kinase inhibitors are also being investigated as potential treatments [220].
The first approach to generate patient derived HHT iPSCs was published by Freund et al. at the Leiden University Medical Centre, in the Netherlands [221]. These efforts were followed by generation and genetic repair of two iPSC clones from a patient bearing a heterozygous ACVRL1 gene leading to HHT2. Correction was performed with CRISPR/Cas9 gene editing, creating control iPSCs from the same patient [222]. Zhou et al. generated isogenic control and HHT1 iPSCs from a patient with an ENG mutation. The iPSCs derived from this patient, carrying a novel missense mutation, displayed reduced potential for differentiation into iECs, which performed poorly in tube formation assays. Furthermore, the C30R ENG mutation resulted in the retention of the mutated endoglin protein in the endoplasmic reticulum, possibly due to protein misfolding. Functionally, BMP-9 stimulation resulted in reduced phospho-SMAD1/5-ID1 signalling in comparison to corrected cells, confirming in human iPSCs, for the first time, that mutant endoglin affects BMP-9 downstream signalling. Moreover, their HHT iECs displayed a disorganized cytoskeleton which, after CRISPR/Cas9 mediated correction, turned into a highly organized cytoskeleton [223]. This recapitulated the previous findings by others using primary blood outgrowth ECs isolated from HHT patients [224]. Very recently, Xiang-Tischhauser et al. established HHT2 iPSCs through gene editing an ACVRL1 mutation in a parental iPSC-clone, causing a frameshift mutation that led to reduced ALK1 expression. In those mutant ALK1 iPSCs, the expression levels of ENG, SMAD2/3, and TGF-βRII were not notably different between parental and mutant iPSC lines; however, they found upregulation of VEGFA and noggin expression. Their gene-edited iPSCs were also differentiated towards ECs and to form embryoid bodies (EBs) [225]. An inspiring work by Orlova et al. underlines the importance of iPSC-derived cell culture under advanced environmental conditions to better recapitulate the human disease complexity. When using 2D cell culture models, control and HHT1-isogenic iECs and pericytes were able to form indistinguishable vascular networks. However, when grown in 3D microfluidic vessel-on-chip devices with fibrin hydrogel and under flow, lumenized vessels formed, in which defective vascular organization was evident [226]. The authors found a leaky 3D vascular network in this iPSC-derived vessel-on-chip model reminiscent of what was reported from HHT patients. The interaction between iECs and primary human brain pericytes that were added to the microfluidic system was disturbed. These results highlight that iPSC culture under organ or tissue-specific conditions can be required for some BMP-related rare disease models in order to mimic cellular responses to a particular genetic defect recapitulating major aspects of the human phenotype.
Interestingly, a few HHT individuals carrying ACVRL1 mutations are predisposed to the development of PAH [227]. Currently, in those ACVR1L mutant affected HHT2 patients, PAH is seen as an extremely rare complication [228]. Similarly, GDF2 mutations cause a vascular-anomaly syndrome somehow overlapping with HHT [205], and some BMPR2 variants increase the risk of developing pulmonary AVMs reminiscent of an HHT-like condition [229]. Vice versa, it is not surprising that ACVR1L mutations can also cause PAH-like and HHT symptoms [228,230] in the same patient. In that respect, more research is required to understand why the same mutation can cause different vascular abnormalities in different vascular beds in some patients but are also very rarely combined in the same individuum.
Table 1. Overview on published BMP-related rare disease iPSC models.
Table 1. Overview on published BMP-related rare disease iPSC models.
Rare DiseaseMutation Somatic Source Cell Age and Sex of Donor (s)Re-Programming MethodDifferentiation ProtocolReference
FOPACVR1
c.617G>A (p.R206H)
c.1067G>A
(p.G356D)
Dermal fibroblasts18-M,
59-F,
22-M,
66-F
SendaiSpontaneous differentiation into mesodermal and endodermal lineages[114]
FOPACVR1
c.617G>A (p.R206H)
Dermal fibroblasts16-F,
16-F,
∅-M
∅-M
12-M,
12-M,
50-M
Retroviral,
Episomal
Osteoblast-like mineralization,
chondrogenic
EC: [121]
2D and 3D chondrogenic: [130]
Presomitic mesoderm and somites [125]
Muscle stem/progenitor cells: [126]
[115]
FOPACVR1
c.617G>A (p.R206H)
Kidney cells isolated from urine∅-M
∅-M
EpisomalEC[116]
FOPACVR1
c.617G>A (p.R206H)
Urinary cells∅-M
∅-F
SendaiEC: [133][231]
FOPACVR1
c.617G>A (p.R206H)
Periodontal ligament fibroblasts 23 yearsSendaiSpontaneous differentiation into EBs[136]
HPAHBMPR2
c.27G>A (p.W9X)
c.1039T>C (p.C347R)
Late-outgrowth endothelial progenitor cells from vein blood∅-∅
∅-∅
RetroviralDifferentiation into 3 germ layers and teratoma[183]
HPAHBMPR2
c.354T>G (p.C118W)
c.2504del (p.T835fs)
c.350G>A p.C117>Y
Dermal fibroblasts Lentiviral/
Sendai
EC[184]
HPAHBMPR2
c.1471C>T (p.Arg491Trp)
Dermal fibroblasts37-M, 33-F EC[187]
HPAHBMPR2 c.1471C>T
(p.Arg491Trp)
Dermal fibroblasts37-M, 33-FSendai EC[185]
HPAHBMPR2
W9X+/−
C2 ΔExon1
Parental wt iPSC line genome edited via CRISPR/Cas9∅-∅ SMC
EC
[191]
HPAHBMPR2
c.1598A>G (p.His533Arg)
c.248-1–418+1del171
(p.Cys84_Ser140del)
c.1471C>T (p.Arg491Trp)
CD34+ blood cells38-M, 17-F, 5-F Sendai In vitro spontaneous differentiation into 3 germ layers[232]
HHT2ACVRL1
(ALK1)
Dermal fibroblasts∅-∅RetroviralCardiomyocytes by END-2 co-culture[221]
HHT2ACVRL1
c.1120del18
Dermal fibroblasts42-MEpisomalIn vitro spontaneous differentiation and directed trilineage differentiation[222]
HHT2ACVRL1
c.772 + 3_772 + 4dup
PBMCs61-FSendaiIn vitro directed trilineage differentiation[233]
HHT2ACVRL1
heterozygous frameshift deletion of 17 base pairs in exon 8
Fibroblast-derived parental iPSC line ATCC; CRL-2097 genome edited via CRISPR/Cas9Neonate (<1 month)-MSendaiEmbryoid bodies and ECs [225]
HHT-1ENG
c.88T>C (p.C30R)
PBMNCs62-FEpisomalEC[223]
HHT-1ENG
c.1678C>T p.(G560∗)
PBMCs∅-MEpisomal EC
co-culture with primary human brain vascular pericytes
[226]
An overview of published BMP rare disease hiPSC models is given: Rare disease abbreviation (column 1), gene mutation and consequences for protein (column 2), somatic source-cells used for reprogramming (column 3), age and sex (column 4), reprogramming method (column 5), established differentiation protocols (column 6), references (column 6); ∅ = not disclosed, ∗ = translation termination, M = male, F = female, EBs = embryoid bodies.

3. Culturing iPSCs under (Patho)Physiologically Relevant Microenvironmental Conditions

Traditionally, iPSC studies use 2D monolayer cultures combined with specific differentiation conditions, consisting in basal media supplemented with defined concentrations of small molecule inhibitors and differentiation factors. A 2D culture is conducted on an iPSC compliant culture surface, such as basement membrane coating. However, to accurately mimic BMP-related rare disease mechanisms in vitro, 3D culture conditions and the combination of different cell types in a biomechanically challenging environment may be required, as demonstrated by Orlova et al. [226] (Figure 4). Achieving this necessitates expertise in (stem) cell biology, TGF-β and/or BMP signalling, cell-material interactions, mechanobiology, and the biochemical context of diseased cells in affected tissues. To replace animal disease models with human iPSC technologies, complex in vitro culture conditions using devices like microfluidic chips and bioreactors are needed to address these aspects. These systems should resemble the complex tissue architecture, circulation, and systemic effects. Current efforts aim to develop fully integrated, controlled, and sensor-equipped systems capable of simultaneously studying multiple cell types [234,235]. Microfluidic organ- and body-chips represent a promising model for iPSC-derived cell culture under complex 3D conditions. Co-culture of iPSC-derived cells in such devices allows the study of the interplay between different cell types from a common genetic background. iPSC-derived cell culture can be combined with tissue engineering and bioprinting of three-dimensional cell constructs, organoids, and grafts [236,237] (Figure 4). Based on person-centred biomedical data of the somatic cell donor (e.g., age, sex, ethnics, co-morbidities, medication, medical history), the goal for patient-specific disease modelling, as well as personalized drug screening and personalized medicine, may be to combine microfluidic organ-on-chip technology with BMP rare disease patient iPSC-derived cells in a 3D matrix (organoids), following the concept of “miniature twinning” [238,239]. Organ-on-chip, 3D organoid-on-chip, or multi-organ-chip systems currently attract a lot of attention [240,241] and have been successfully applied in recent years to create complex laboratory model systems for, e.g., systemic COVID-19 research [240], blood–brain barrier function [242], and bone regeneration [243,244]. The integration of biosensors into those high-tech systems and the concomitant application of defined cell mechanics make them an even more versatile tool with which to study disease mechanisms and accelerate drug development (Figure 4).

3.1. Microenvironmental Aspects Relevant for Mimicking BMP Rare Diseases with iPSCs

The Knudson hypothesis, also known as the “two-hit” hypothesis, states that most tumour suppressor genes require both alleles to be inactivated [245]. Considering additional genetic variations that act in addition to the primary genetic event, PAH and HHT can be considered “two or second-hit diseases” as well [246,247,248,249,250]. Both come with low penetrance, and not every BMPR2 and ALK1/ENG mutant carrier develops clinical symptoms. Multifactorial inputs seem required for PAH and HHT onset and progression, beyond genetic disruptions. For BMP rare diseases like PAH and HHT, the concept of the “genetic” second hit hypothesis could be extended towards microenvironmental cues, eventually cell/bio mechanics, hypoxia, and inflammation, that add to disease onset and progression as a “non-genetic second-hit” [66,170,216,251,252]. Any iPSC-derived in vitro model, aiming at resembling the human PAH and HHT pathomechanism as closely as possible, may consider that.
In FOP, the combination of inflammatory tissue trauma and the genetic ACVR1 mutation is necessary to drive extraskeletal ossification, strongly suggesting such non-genetic second-hit dependency as well. The expression of INHBA, encoding for the Activin-A homodimer, is induced by inflammatory signalling pathways [253], but systemic Activin levels are not increased in FOP individuals [254], suggesting a strong role for the local cellular FOP microenvironment versus systemic factors. Interestingly, recent single cell RNA sequencing data obtained from HO lesions in FOP mice indicates that Activin mRNA expression is induced locally [255]. Congruently, a recent report has shown how dermal fibroblasts from FOP donors express high levels of Activin-A compared to controls [138], and, as aforementioned, iPSC-derived inflammatory macrophages showed a significantly higher production of Activin-A [127]. This might suggest that specific local tissues or organ environments may modulate functional cell responses in FOP. Intriguingly, this has been previously discussed not only in FOP, but also in PAH and HHT [256,257,258,259]. An intricate feedforward regulatory dependency exists with regard to the local cellular ECM-based biomechanical microenvironment and how this may be influenced by mutations in the BMP pathway. Haupt et al. showed that progenitor cells from FOP-like mutant mice misinterpret the tissue microenvironment through altered sensitivity to mechanical stimuli, which lowered the threshold for commitment into chondro/osteogenic lineages [260]. Evidence suggests that ALK2R206H mutant cells contribute to an active stiffening of the surrounding ECM, e.g., by increased collagen production [121,126]. Members of the TGF-β ligand family were shown to induce osteogenic differentiation of mesenchymal progenitors more potently on stiff substrates [261]. Altered BMP-SMAD signalling influences the expression of ECM genes in various cell types, leading to changes in the mechanical microenvironment. Increased ECM deposition, e.g., affects the local stiffness and elasticity, determined by ECM composition and degree of crosslinking. The cells elaborate feedback responses to these changes through outside-in signalling, primarily mediated via integrins, which, themselves, interact with—and crosstalk to—BMPRs [262]. Integrins enable cells to transduce feedback from alterations in ECM stiffness into their signalling pathways, a phenomenon generally observed in TGF-β family signalling (reviewed in [263]). Interestingly, ECM changes, actin-cytoskeletal rearrangements, and mechanical features such as changes in cell stiffness, traction forces, and actomyosin contractility show commonalities between all three mentioned BMP rare diseases: FOP, PAH, and HHT [168,260,264,265,266]. Particularly, pathological defects on the cell cytoskeletal organization, cell morphology, migration behaviour, and adhesion molecule expression influence cell–cell and cell–substratum traction forces [267,268]. In PAH, ECM deposition and cytoskeletal defects are likely to contribute to pulmonary artery stiffening, constriction, and autocrine TGF-β signalling through increased retrieval of mature TGF-β from ECM-bound latency complexes [168]. In HHT, blood flow and changes in cellular biomechanics may affect cell junction tightness and pericyte coverage. Analogies can be drawn between HHT and the role of biomechanics in the field of cerebral cavernous malformations, where the intricate relationship between vascular malformations and dysplasias with EC mechanotransduction is better understood [269].
The important role of fluid shear stress and the mechanotransduction of vascular cells for PAH and HHT development was shown [66,270], and mechanical stimulation of different cell types can influence BMP signalling outcomes [52,65,271,272] (Figure 4). Endoglin is required to implement extracellular mechanical forces, such as shear stress exerted by blood flow, into the BMP signalling pathways [66,156,157]. The endoglin role in flow-perception during the HHT pathomechanism is currently investigated by several laboratories. To address shear stress by (micro)fluidic systems in both PAH and HHT, channel geometries and diameter, flow rates, viscosity of the medium, and the particular type of wall shear stress need to be considered. The shear stress values in arterial vessel types can be very different (5–50 dyne/cm2), and wall shear stress can follow different flow regimes [273]. During turbulent flow, regions of flow recirculation or flow separation can occur, which are more challenging to reproduce in a microfluidic device. Understanding the exact contribution of fluid shear stress and its integration into BMP signalling may help explaining why a single mutation (e.g., ALK1 or ENG) can lead to two distinct rare disease phenotypes (PAH and HHT) in different vascular beds (pulmonary arteries vs. microvasculature). The flow characteristics in the microvasculature differ significantly from those in arteries and veins impacting oxygen levels [274]. Additionally, arterial, venous, and microvascular ECs exhibit distinct responses to blood flow [275]. Therefore, in order to replicate PAH and HHT using microenvironmental mechanical shear stress cues, it is essential to consider the appropriate perfusion systems that allow for different flow rates and regimes, oxygen conditions, and the culture of specific subtypes of EC, eventually even in the presence of SMC or pericyte co-cultures (Figure 4). Several protocols have already been developed to differentiate iPSCs into arterial-like, venous-like, and microvascular-like ECs, and these have been combined with microfluidic platforms that mimic arterial, venous, and microcirculatory perfusion [180,276,277]. In summary, although iPSC technology has made progress in the study of BMP-related rare diseases, there is still a long way to go in establishing complex in vitro platforms that accurately mimic the biochemical and biomechanical cellular processes involved in the onset and progression of these pathologies.

4. Future Perspectives

The future holds iPSC-derived BMP rare disease models that can be cultured under complex microenvironmental conditions aiming at resembling closely the pathological in vivo conditions. This involves incorporating inflammatory cues, disease-specific mechanical stimulation, and a permissive microenvironment for optimal disease mimicry. In terms of BMP signalling interrogation, accurately recapitulating disease (micro)environments may not only directly impact canonical BMP-SMAD pathways, but also the crosstalk with other branches of the TGF-β superfamily and/or other non-canonical signalling cascades. As such, “non canonical” and/or “non-SMAD” signalling must be better defined and further explored [278]. While a number of perturbed signalling mechanisms related to FOP, PAH, and HHT were already discovered, further insights on diversification and cell-type specific BMP signalling modes may increase by use of iPSC technology and its combination with CRISPR/Cas9, e.g., by the design of smart biosensors and a higher resolution of the applied analytics. As such, the discovery of “lateral SMAD” signalling, i.e., the activation of BMP-SMADS by TGF-β or Activins via altered receptor complex stoichiometry, is of current interest for the mentioned pathologies [279]. This only emphasizes the need for comprehensive studies aiming to integrate input from different layers, to dissect complex regulatory networks, and to identify novel disease biomarkers or druggable targets. To overcome this, almost all omics methods are rapidly developing towards increasing depth, robustness, automatization, and affordability, while requiring smaller sample yields. These characteristics will ultimately facilitate the implementation of human iPSC-based miniaturized disease models as routine preclinical platforms in biomedical research. The FDA no longer requires animal testing for new drugs entering human clinical trials, which is also addressing ethical concerns [280]. While a number of transgenic animal models exist for studying BMP rare diseases, they may not fully replicate human disease phenotypes. However, patient-derived iPSC disease models also present challenges, i.e., obtaining rare-disease patient material, scalability, stability, automation, and material selection [281]. Even though the implementation of CRISPR/Cas9 to generate control and disease lines with an identical genetic background has substantially improved the field, other aspects of culturing and differentiating iPSCs are still a source of high variability. For example, substrate stiffness impacts stem and progenitor cell differentiation [282]. When selecting substrates for iPSC maintenance, differentiation, or organoid formation, considering their stiffness and the effect on differentiation is important [283,284]. Various hydrogel systems using natural or synthetic materials have been developed to mimic the stem cell niche in vitro, providing biochemical and biophysical signals. The choice of the right ECM for culturing iPSCs, whether in 2D or 3D organoids, can impact ECM secretion and formation by the cells. Considering the potential involvement of ECM in disease mechanisms such as FOP and PAH, caution is necessary when selecting the right ECM for BMP rare disease modelling. ECM extracts have batch-dependent variability, and their polymerization is influenced by multiple factors. The sequestration of recombinant BMPs by natural ECM proteins used as iPSC substrates can reduce the effective concentrations required for BMP stimulation of iPSCs towards mesenchymal differentiation [285]. Instead, current research focuses on designer matrices for iPSC culture [286,287], which offer better reproducibility and adaptability to, e.g., bio(ink)-based printing technologies (Figure 4). Importantly, every new technology development that claims to represent an appropriate animal-free alternative system for disease modelling needs to demonstrate that the cellular model is functionally and phenotypically representative of the native cell/tissue. For this, more comparison and referencing towards, e.g., patient-derived tissue samples is required. Finally, international harmonization and implementation of quality control standards concerning iPSC research with regard to iPSC generation, differentiation, and culturing protocols are needed. Current efforts towards more standardization in stem cell use in research by the International Society for Stem Cell Research aims at making results more comparable and interoperable, eventually facilitating a better and more efficient clinical translation of basic research findings (www.isscr.org/standards-document) accessed on 22 August 2023. This should integrate alongside a shared access to BMP rare disease patient material, which is crucial to facilitate generation of more patient-derived iPSC lines and new differentiation protocol workflows. Ultimately, modelling human diseases with iPSCs will speed up the identification of druggable targets and the identification of new pharmaceutical compounds for treating FOP, PAH, and HHT.

5. Conclusions

Since the invention of iPSC technology, FOP, PAH, and HHT research has greatly benefited from cellular human disease models in investigating underlying pathomechanisms and new drug targets. These developments complement conventional approaches, such as use of genetic animal models, which, in some cases, did not gain the anticipated insights. BMP signalling and its cellular consequences are highly complex and diverse in a cell-type specific manner, resulting in fine-tuned canonical SMAD and non-canonical pathways, which crosstalk and interplay with each other and with a variety of other signalling related to ECM, biomechanics, inflammation, hypoxia, and other growth factors. The current aim for several research groups is to mimic this complex, multifaceted cellular microenvironment as closely as technically possible to establish adequate cellular human FOP, PAH, and HHT disease models. Those models will continue to show value in the preclinical validation of promising new compounds. Besides easier access to the already established models, which we summed up here, new iPSC-derived models may be necessary to cater to the needs of increasing omics-based analysis and further exploration of more specific differentiation protocols. It is important to also expand iPSC technology towards BMP-related diseases that are currently not addressed.

Author Contributions

C.H. and G.S.-D. contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

C.H. was funded by “The Open Access Publication Fund of the Westphalian University of Applied Sciences” as well as the 5th Research Challenge on Sustainability (Internal program for science promotion of a sustainable future by the Westphalian University of Applied Sciences). G.S.-D. is sponsored by Fundació La Marató de TV3 (#202038), the Spanish Ministry of Science, through the Ramón y Cajal grants RYC2021-030866-I and PID2022-141212OA-I00, and the BHF-DZHK-DHF, 2022/23 award PROMETHEUS. Both authors would like to thank the Scientific Research Network by the Research Foundation–Flanders (WOG W0014200N) and An Zwijsen.

Acknowledgments

All figures were created with biorender.com (licence to C.H.). We thank the Anatomical Museum of the Leiden University Medical Center (LUMC) for providing a picture of an FOP patient’s skeleton (copyright belongs to LUMC). We thank Urban Geisthoff (Philipps University of Marburg, Germany) for permission to use a photograph of a subdermal lesion of an HHT patient for this review.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

2DTwo dimensional
3DThree dimensional
ACVR1Activin A receptor type 1
ALK2Activin receptor-like kinase-2
ALPLAlkaline phosphatase
BMPBone morphogenetic protein
BMPR1Bone morphogenetic protein receptor type 1
BMPR2Bone morphogenetic protein receptor type 2
Cas9CRISPR-associated protein 9
CAV1Caveolin 1
CD31Platelet endothelial cell adhesion molecule 1
CD45Tyrosin-proteinphosphatase C
C-MYCc-myelocytomatosis oncogene product
COL1A1Collagen type I1 alpha 1 chain
COL2A1Collagen type II alpha 1 chain
CRISPRClustered Regularly Interspaced Short Palindromic Repeats
ECEndothelial cell
ECMExtracellular matrix
EIF2AK4Eukaryotic translation initiation factor 2 alpha kinase 4
EndMTEndothelial-to-mesenchymal transition
ENGEndoglin
EPCEndothelial progenitor cell
ERKExtracellular signal-regulated kinase
FAPFibro/adipogenic progenitor
FDAU.S. Food and Drug Administration
FGFFibroblast growth factor
FKBP12FK506 binding protein 1a
FOPFibrodysplasia ossificans progressiva
FSP-1Ferroptosis suppressor protein 1
GSK3-βGlycogen synthase kinase 3 beta
hESCHuman embryonic stem cell
HHTHereditary haemorrhagic telangiectasia
HHT1Hereditary haemorrhagic telangiectasia type 1
HHT2Hereditary haemorrhagic telangiectasia type 2
hiPSCsHuman induced pluripotent stem cells
HOHeterotopic ossification
HPAHHereditary pulmonary arterial hypertension
MuSCsHuman primary satellite cells
IDInhibitor-of-differentiation
iECsiPSC-derived endothelial cells
IFN-g Interferon-gamma
IL-6Interleukin 6
iMACiPSC-derived macrophages
iMPCsiPSC-derived muscle stem/progenitor cells
iMSCsiPSC-derived mesenchymal stromal cells
IPAHIdiopathic pulmonary arterial hypertension
iPSCInduced pluripotent stem cell
I-SMADSInhibitory SMADs
JNKc-Jun N-terminal kinase
KCNK3Potassium two pore domain channel subfamily K member 3
KLF-4Kruppel-like factor 4
L-EPCLate-outgrowth endothelial progenitor cell
MAPKMitogen activated protein kinase
M-CSF Macrophage colony-stimulating factor
MDM2Mouse double minute 2 homolog
mESCMouse embryonic stem cell
mTORMammalian target of rapamycin
NMDNonsense mediated decay
NOTCHNeurogenic locus notch homolog protein 1
OCT-4Octamer-binding transcription factor 4
PAECPulmonary arterial endothelial cells
PAHPulmonary arterial hypertension
PASMCsPulmonary arterial smooth muscle cells
PBMNCPeripheral blood mononuclear cells
PDGFPlatelet-derived growth factor
PDGFRaPlatelet-derived growth factor receptor A
PI3KPhosphoinositide-3-kinase
R-SMADSReceptor regulated SMADs
SCA1Spinocerebellar ataxia type 1
ScxScleraxis bHLH transcription factor
SMADMothers against decapentaplegic homolog
SOX-2SRY-box transcription factor 2
TFMTraction force microscopy
TGF-βTransforming growth factor beta
TGFBRTGF-b receptor
Tie2Angiopoietin-1 receptor
VEGFVascular endothelial growth factor
VSMCVascular smooth muscle cell

References

  1. Takahashi, K.; Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [PubMed]
  2. Shi, Y.; Inoue, H.; Wu, J.C.; Yamanaka, S. Induced pluripotent stem cell technology: A decade of progress. Nat. Rev. Drug Discov. 2017, 16, 115–130. [Google Scholar] [CrossRef] [PubMed]
  3. Ferreira, C.R. The burden of rare diseases. Am. J. Med. Genet. A 2019, 179, 885–892. [Google Scholar] [CrossRef] [PubMed]
  4. Taruscio, D.; Capozzoli, F.; Frank, C. Rare diseases and orphan drugs. Ann. Dell’istituto Super. Sanita 2011, 47, 83–93. [Google Scholar]
  5. Edwards, D.S.; Clasper, J.C. Heterotopic ossification: A systematic review. J. R. Army Med. Corps 2015, 161, 315–321. [Google Scholar] [CrossRef] [PubMed]
  6. Ramirez, D.M.; Ramirez, M.R.; Reginato, A.M.; Medici, D. Molecular and cellular mechanisms of heterotopic ossification. Histol. Histopathol. 2014, 29, 1281–1285. [Google Scholar]
  7. Sun, D.; Gao, W.; Hu, H.; Zhou, S. Why 90% of clinical drug development fails and how to improve it? Acta Pharm. Sin. B 2022, 12, 3049–3062. [Google Scholar] [CrossRef]
  8. Happe, C.; Kurakula, K.; Sun, X.Q.; da Silva Goncalves Bos, D.; Rol, N.; Guignabert, C.; Tu, L.; Schalij, I.; Wiesmeijer, K.C.; Tura-Ceide, O.; et al. The BMP Receptor 2 in Pulmonary Arterial Hypertension: When and Where the Animal Model Matches the Patient. Cells 2020, 9, 1422. [Google Scholar] [CrossRef]
  9. Wang, H.; Brown, P.C.; Chow, E.C.Y.; Ewart, L.; Ferguson, S.S.; Fitzpatrick, S.; Freedman, B.S.; Guo, G.L.; Hedrich, W.; Heyward, S.; et al. 3D cell culture models: Drug pharmacokinetics, safety assessment, and regulatory consideration. Clin. Transl. Sci. 2021, 14, 1659–1680. [Google Scholar] [CrossRef]
  10. Urist, M.R. Bone: Formation by autoinduction. Science 1965, 150, 893–899. [Google Scholar] [CrossRef]
  11. Urist, M.R.; Strates, B.S. Bone morphogenetic protein. J. Dent. Res. 1971, 50, 1392–1406. [Google Scholar] [CrossRef] [PubMed]
  12. Wozney, J.M.; Rosen, V.; Celeste, A.J.; Mitsock, L.M.; Whitters, M.J.; Kriz, R.W.; Hewick, R.M.; Wang, E.A. Novel regulators of bone formation: Molecular clones and activities. Science 1988, 242, 1528–1534. [Google Scholar] [CrossRef] [PubMed]
  13. 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]
  14. 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]
  15. 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]
  16. Garcia de Vinuesa, A.; Sanchez-Duffhues, G.; Blaney-Davidson, E.; van Caam, A.; Lodder, K.; Ramos, Y.; Kloppenburg, M.; Meulenbelt, I.; van der Kraan, P.; Goumans, M.J.; et al. Cripto favors chondrocyte hypertrophy via TGF-beta SMAD1/5 signaling during development of osteoarthritis. J. Pathol. 2021, 255, 330–342. [Google Scholar] [CrossRef]
  17. Perez-Gomez, E.; Villa-Morales, M.; Santos, J.; Fernandez-Piqueras, J.; Gamallo, C.; Dotor, J.; Bernabeu, C.; Quintanilla, M. A role for endoglin as a suppressor of malignancy during mouse skin carcinogenesis. Cancer Res. 2007, 67, 10268–10277. [Google Scholar] [CrossRef]
  18. Watanabe, K.; Bianco, C.; Strizzi, L.; Hamada, S.; Mancino, M.; Bailly, V.; Mo, W.; Wen, D.; Miatkowski, K.; Gonzales, M.; et al. Growth factor induction of Cripto-1 shedding by glycosylphosphatidylinositol-phospholipase D and enhancement of endothelial cell migration. J. Biol. Chem. 2007, 282, 31643–31655. [Google Scholar] [CrossRef]
  19. Siebold, C.; Yamashita, T.; Monnier, P.P.; Mueller, B.K.; Pasterkamp, R.J. RGMs: Structural Insights, Molecular Regulation, and Downstream Signaling. Trends Cell Biol. 2017, 27, 365–378. [Google Scholar] [CrossRef]
  20. Onichtchouk, D.; Chen, Y.G.; Dosch, R.; Gawantka, V.; Delius, H.; Massague, J.; Niehrs, C. Silencing of TGF-beta signalling by the pseudoreceptor BAMBI. Nature 1999, 401, 480–485. [Google Scholar] [CrossRef]
  21. Hagihara, M.; Endo, M.; Hata, K.; Higuchi, C.; Takaoka, K.; Yoshikawa, H.; Yamashita, T. Neogenin, a receptor for bone morphogenetic proteins. J. Biol. Chem. 2011, 286, 5157–5165. [Google Scholar] [CrossRef] [PubMed]
  22. Bragdon, B.; Thinakaran, S.; Moseychuk, O.; King, D.; Young, K.; Litchfield, D.W.; Petersen, N.O.; Nohe, A. Casein kinase 2 beta-subunit is a regulator of bone morphogenetic protein 2 signaling. Biophys. J. 2010, 99, 897–904. [Google Scholar] [CrossRef] [PubMed]
  23. Dannewitz Prosseda, S.; Tian, X.; Kuramoto, K.; Boehm, M.; Sudheendra, D.; Miyagawa, K.; Zhang, F.; Solow-Cordero, D.; Saldivar, J.C.; Austin, E.D.; et al. FHIT, a Novel Modifier Gene in Pulmonary Arterial Hypertension. Am. J. Respir. Crit. Care Med. 2019, 199, 83–98. [Google Scholar] [CrossRef]
  24. Andruska, A.M.; Ali, M.K.; Tian, X.; Spiekerkoetter, E. Selective Src-Family B Kinase Inhibition Promotes Pulmonary Artery Endothelial Cell Dysfunction. bioRxiv 2021. [Google Scholar] [CrossRef]
  25. Ali, M.K.; Tian, X.; Zhao, L.; Schimmel, K.; Rhodes, C.J.; Wilkins, M.R.; Nicolls, M.R.; Spiekerkoetter, E.F. PTPN1 Deficiency Modulates BMPR2 Signaling and Induces Endothelial Dysfunction in Pulmonary Arterial Hypertension. Cells 2023, 12, 316. [Google Scholar] [CrossRef]
  26. Wilkinson, L.; Kolle, G.; Wen, D.; Piper, M.; Scott, J.; Little, M. CRIM1 regulates the rate of processing and delivery of bone morphogenetic proteins to the cell surface. J. Biol. Chem. 2003, 278, 34181–34188. [Google Scholar] [CrossRef] [PubMed]
  27. Martino, M.M.; Hubbell, J.A. The 12th–14th type III repeats of fibronectin function as a highly promiscuous growth factor-binding domain. FASEB J. 2010, 24, 4711–4721. [Google Scholar]
  28. Hauff, K.; Zambarda, C.; Dietrich, M.; Halbig, M.; Grab, A.L.; Medda, R.; Cavalcanti-Adam, E.A. Matrix-Immobilized BMP-2 on Microcontact Printed Fibronectin as an in vitro Tool to Study BMP-Mediated Signaling and Cell Migration. Front. Bioeng. Biotechnol. 2015, 3, 62. [Google Scholar] [CrossRef]
  29. Xu, E.R.; Blythe, E.E.; Fischer, G.; Hyvonen, M. Structural analyses of von Willebrand factor C domains of collagen 2A and CCN3 reveal an alternative mode of binding to bone morphogenetic protein-2. J. Biol. Chem. 2017, 292, 12516–12527. [Google Scholar] [CrossRef]
  30. Sieron, A.L.; Louneva, N.; Fertala, A. Site-specific interaction of bone morphogenetic protein 2 with procollagen II. Cytokine 2002, 18, 214–221. [Google Scholar] [CrossRef]
  31. Sengle, G.; Charbonneau, N.L.; Ono, R.N.; Sasaki, T.; Alvarez, J.; Keene, D.R.; Bachinger, H.P.; Sakai, L.Y. Targeting of bone morphogenetic protein growth factor complexes to fibrillin. J. Biol. Chem. 2008, 283, 13874–13888. [Google Scholar] [CrossRef]
  32. Wohl, A.P.; Troilo, H.; Collins, R.F.; Baldock, C.; Sengle, G. Extracellular Regulation of Bone Morphogenetic Protein Activity by the Microfibril Component Fibrillin-1. J. Biol. Chem. 2016, 291, 12732–12746. [Google Scholar] [CrossRef] [PubMed]
  33. Spanou, C.E.S.; Wohl, A.P.; Doherr, S.; Correns, A.; Sonntag, N.; Lutke, S.; Morgelin, M.; Imhof, T.; Gebauer, J.M.; Baumann, U.; et al. Targeting of bone morphogenetic protein complexes to heparin/heparan sulfate glycosaminoglycans in bioactive conformation. FASEB J. 2023, 37, e22717. [Google Scholar] [CrossRef]
  34. Kanzaki, S.; Takahashi, T.; Kanno, T.; Ariyoshi, W.; Shinmyouzu, K.; Tujisawa, T.; Nishihara, T. Heparin inhibits BMP-2 osteogenic bioactivity by binding to both BMP-2 and BMP receptor. J. Cell. Physiol. 2008, 216, 844–850. [Google Scholar] [CrossRef] [PubMed]
  35. Kanzaki, S.; Ariyoshi, W.; Takahashi, T.; Okinaga, T.; Kaneuji, T.; Mitsugi, S.; Nakashima, K.; Tsujisawa, T.; Nishihara, T. Dual effects of heparin on BMP-2-induced osteogenic activity in MC3T3-E1 cells. Pharmacol. Rep. 2011, 63, 1222–1230. [Google Scholar] [CrossRef] [PubMed]
  36. Rider, C.C. Heparin/heparan sulphate binding in the TGF-beta cytokine superfamily. Biochem. Soc. Trans. 2006, 34 Pt 3, 458–460. [Google Scholar] [CrossRef] [PubMed]
  37. Choi, Y.J.; Lee, J.Y.; Park, J.H.; Park, J.B.; Suh, J.S.; Choi, Y.S.; Lee, S.J.; Chung, C.P.; Park, Y.J. The identification of a heparin binding domain peptide from bone morphogenetic protein-4 and its role on osteogenesis. Biomaterials 2010, 31, 7226–7238. [Google Scholar] [CrossRef] [PubMed]
  38. Correns, A.; Zimmermann, L.A.; Baldock, C.; Sengle, G. BMP antagonists in tissue development and disease. Matrix Biol. Plus 2021, 11, 100071. [Google Scholar] [CrossRef]
  39. Massague, J. TGFbeta signalling in context. Nat. Rev. Mol. Cell Biol. 2012, 13, 616–630. [Google Scholar] [CrossRef]
  40. Runyan, C.E.; Schnaper, H.W.; Poncelet, A.C. The role of internalization in transforming growth factor beta1-induced Smad2 association with Smad anchor for receptor activation (SARA) and Smad2-dependent signaling in human mesangial cells. J. Biol. Chem. 2005, 280, 8300–8308. [Google Scholar] [CrossRef]
  41. Shi, W.; Chang, C.; Nie, S.; Xie, S.; Wan, M.; Cao, X. Endofin acts as a Smad anchor for receptor activation in BMP signaling. J. Cell Sci. 2007, 120 Pt 7, 1216–1224. [Google Scholar] [CrossRef] [PubMed]
  42. Hanyu, A.; Ishidou, Y.; Ebisawa, T.; Shimanuki, T.; Imamura, T.; Miyazono, K. The N domain of Smad7 is essential for specific inhibition of transforming growth factor-beta signaling. J. Cell Biol. 2001, 155, 1017–1027. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, Y.S.; Park, J.S.; Kim, J.H.; Jung, S.M.; Lee, J.Y.; Kim, S.J.; Park, S.H. Smad6-specific recruitment of Smurf E3 ligases mediates TGF-beta1-induced degradation of MyD88 in TLR4 signalling. Nat. Commun. 2011, 2, 460. [Google Scholar] [CrossRef] [PubMed]
  44. Goto, K.; Kamiya, Y.; Imamura, T.; Miyazono, K.; Miyazawa, K. Selective inhibitory effects of Smad6 on bone morphogenetic protein type I receptors. J. Biol. Chem. 2007, 282, 20603–20611. [Google Scholar] [CrossRef] [PubMed]
  45. Murakami, G.; Watabe, T.; Takaoka, K.; Miyazono, K.; Imamura, T. Cooperative inhibition of bone morphogenetic protein signaling by Smurf1 and inhibitory Smads. Mol. Biol. Cell 2003, 14, 2809–2817. [Google Scholar] [CrossRef]
  46. Sieber, C.; Kopf, J.; Hiepen, C.; Knaus, P. Recent advances in BMP receptor signaling. Cytokine Growth Factor Rev. 2009, 20, 343–355. [Google Scholar] [CrossRef]
  47. Derynck, R.; Zhang, Y.E. Smad-dependent and Smad-independent pathways in TGF-beta family signalling. Nature 2003, 425, 577–584. [Google Scholar] [CrossRef]
  48. Wakayama, Y.; Fukuhara, S.; Ando, K.; Matsuda, M.; Mochizuki, N. Cdc42 mediates Bmp-induced sprouting angiogenesis through Fmnl3-driven assembly of endothelial filopodia in zebrafish. Dev. Cell 2015, 32, 109–122. [Google Scholar] [CrossRef]
  49. Hiepen, C.; Benn, A.; Denkis, A.; Lukonin, I.; Weise, C.; Boergermann, J.H.; Knaus, P. BMP2-induced chemotaxis requires PI3K p55gamma/p110alpha-dependent phosphatidylinositol (3,4,5)-triphosphate production and LL5beta recruitment at the cytocortex. BMC Biol. 2014, 12, 43. [Google Scholar] [CrossRef]
  50. Sapkota, G.; Alarcon, C.; Spagnoli, F.M.; Brivanlou, A.H.; Massague, J. Balancing BMP signaling through integrated inputs into the Smad1 linker. Mol. Cell 2007, 25, 441–454. [Google Scholar] [CrossRef]
  51. Fuentealba, L.C.; Eivers, E.; Ikeda, A.; Hurtado, C.; Kuroda, H.; Pera, E.M.; De Robertis, E.M. Integrating patterning signals: Wnt/GSK3 regulates the duration of the BMP/Smad1 signal. Cell 2007, 131, 980–993. [Google Scholar] [CrossRef] [PubMed]
  52. Reichenbach, M.; Mendez, P.L.; da Silva Madaleno, C.; Ugorets, V.; Rikeit, P.; Boerno, S.; Jatzlau, J.; Knaus, P. Differential Impact of Fluid Shear Stress and YAP/TAZ on BMP/TGF-beta Induced Osteogenic Target Genes. Adv. Biol. 2021, 5, e2000051. [Google Scholar] [CrossRef] [PubMed]
  53. 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] [PubMed]
  54. Neuhaus, H.; Rosen, V.; Thies, R.S. Heart specific expression of mouse BMP-10 a novel member of the TGF-beta superfamily. Mech. Dev. 1999, 80, 181–184. [Google Scholar] [CrossRef]
  55. Chen, H.; Shi, S.; Acosta, L.; Li, W.; Lu, J.; Bao, S.; Chen, Z.; Yang, Z.; Schneider, M.D.; Chien, K.R.; et al. BMP10 is essential for maintaining cardiac growth during murine cardiogenesis. Development 2004, 131, 2219–2231. [Google Scholar] [CrossRef]
  56. Jones, C.M.; Lyons, K.M.; Hogan, B.L. Involvement of Bone Morphogenetic Protein-4 (BMP-4) and Vgr-1 in morphogenesis and neurogenesis in the mouse. Development 1991, 111, 531–542. [Google Scholar] [CrossRef]
  57. Ozkaynak, E.; Schnegelsberg, P.N.; Oppermann, H. Murine osteogenic protein (OP-1): High levels of mRNA in kidney. Biochem. Biophys. Res. Commun. 1991, 179, 116–123. [Google Scholar] [CrossRef]
  58. Wang, L.; Rice, M.; Swist, S.; Kubin, T.; Wu, F.; Wang, S.; Kraut, S.; Weissmann, N.; Bottger, T.; Wheeler, M.; et al. BMP9 and BMP10 Act Directly on Vascular Smooth Muscle Cells for Generation and Maintenance of the Contractile State. Circulation 2021, 143, 1394–1410. [Google Scholar] [CrossRef]
  59. Astrom, A.K.; Jin, D.; Imamura, T.; Roijer, E.; Rosenzweig, B.; Miyazono, K.; ten Dijke, P.; Stenman, G. Chromosomal localization of three human genes encoding bone morphogenetic protein receptors. Mamm. Genome 1999, 10, 299–302. [Google Scholar]
  60. Ide, H.; Katoh, M.; Sasaki, H.; Yoshida, T.; Aoki, K.; Nawa, Y.; Osada, Y.; Sugimura, T.; Terada, M. Cloning of human bone morphogenetic protein type IB receptor (BMPR-IB) and its expression in prostate cancer in comparison with other BMPRs. Oncogene 1997, 14, 1377–1382. [Google Scholar] [CrossRef]
  61. 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. 2012, 69, 313–324. [Google Scholar] [CrossRef] [PubMed]
  62. Brock, M.; Trenkmann, M.; Gay, R.E.; Michel, B.A.; Gay, S.; Fischler, M.; Ulrich, S.; Speich, R.; Huber, L.C. Interleukin-6 modulates the expression of the bone morphogenic protein receptor type II through a novel STAT3-microRNA cluster 17/92 pathway. Circ. Res. 2009, 104, 1184–1191. [Google Scholar] [CrossRef]
  63. Hurst, L.A.; Dunmore, B.J.; Long, L.; Crosby, A.; Al-Lamki, R.; Deighton, J.; Southwood, M.; Yang, X.; Nikolic, M.Z.; Herrera, B.; et al. TNFalpha drives pulmonary arterial hypertension by suppressing the BMP type-II receptor and altering NOTCH signalling. Nat. Commun. 2017, 8, 14079. [Google Scholar] [CrossRef] [PubMed]
  64. Kim, C.W.; Song, H.; Kumar, S.; Nam, D.; Kwon, H.S.; Chang, K.H.; Son, D.J.; Kang, D.W.; Brodie, S.A.; Weiss, D.; et al. Anti-inflammatory and antiatherogenic role of BMP receptor II in endothelial cells. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 1350–1359. [Google Scholar] [CrossRef] [PubMed]
  65. Mendez, P.L.; Obendorf, L.; Jatzlau, J.; Burdzinski, W.; Reichenbach, M.; Nageswaran, V.; Haghikia, A.; Stangl, V.; Hiepen, C.; Knaus, P. Atheroprone fluid shear stress-regulated ALK1-Endoglin-SMAD signaling originates from early endosomes. BMC Biol. 2022, 20, 210. [Google Scholar] [CrossRef]
  66. Baeyens, N.; Larrivee, B.; Ola, R.; Hayward-Piatkowskyi, B.; Dubrac, A.; Huang, B.; Ross, T.D.; Coon, B.G.; Min, E.; Tsarfati, M.; et al. Defective fluid shear stress mechanotransduction mediates hereditary hemorrhagic telangiectasia. J. Cell Biol. 2016, 214, 807–816. [Google Scholar] [CrossRef]
  67. Rui, Y.F.; Lui, P.P.; Ni, M.; Chan, L.S.; Lee, Y.W.; Chan, K.M. Mechanical loading increased BMP-2 expression which promoted osteogenic differentiation of tendon-derived stem cells. J. Orthop. Res. 2011, 29, 390–396. [Google Scholar] [CrossRef]
  68. Gilde, F.; Fourel, L.; Guillot, R.; Pignot-Paintrand, I.; Okada, T.; Fitzpatrick, V.; Boudou, T.; Albiges-Rizo, C.; Picart, C. Stiffness-dependent cellular internalization of matrix-bound BMP-2 and its relation to Smad and non-Smad signaling. Acta Biomater. 2016, 46, 55–67. [Google Scholar] [CrossRef]
  69. Wei, Q.; Holle, A.; Li, J.; Posa, F.; Biagioni, F.; Croci, O.; Benk, A.S.; Young, J.; Noureddine, F.; Deng, J.; et al. BMP-2 Signaling and Mechanotransduction Synergize to Drive Osteogenic Differentiation via YAP/TAZ. Adv. Sci. 2020, 7, 1902931. [Google Scholar] [CrossRef]
  70. De Robertis, E.M.; Kuroda, H. Dorsal-ventral patterning and neural induction in Xenopus embryos. Annu. Rev. Cell Dev. Biol. 2004, 20, 285–308. [Google Scholar] [CrossRef]
  71. Zinski, J.; Tajer, B.; Mullins, M.C. TGF-beta Family Signaling in Early Vertebrate Development. Cold Spring Harb. Perspect. Biol. 2018, 10, a033274. [Google Scholar] [CrossRef] [PubMed]
  72. Nishimatsu, S.; Thomsen, G.H. Ventral mesoderm induction and patterning by bone morphogenetic protein heterodimers in Xenopus embryos. Mech. Dev. 1998, 74, 75–88. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, P.; Li, J.; Tan, Z.; Wang, C.; Liu, T.; Chen, L.; Yong, J.; Jiang, W.; Sun, X.; Du, L.; et al. Short-term BMP-4 treatment initiates mesoderm induction in human embryonic stem cells. Blood 2008, 111, 1933–1941. [Google Scholar] [CrossRef] [PubMed]
  74. Sporle, R.; Schughart, K. Neural tube morphogenesis. Curr. Opin. Genet. Dev. 1997, 7, 507–512. [Google Scholar] [CrossRef] [PubMed]
  75. Shroyer, N.F.; Wong, M.H. BMP signaling in the intestine: Cross-talk is key. Gastroenterology 2007, 133, 1035–1038. [Google Scholar] [CrossRef]
  76. Herrera, B.; Sanchez, A.; Fabregat, I. BMPS and liver: More questions than answers. Curr. Pharm. Des. 2012, 18, 4114–4125. [Google Scholar] [CrossRef]
  77. Godin, R.E.; Robertson, E.J.; Dudley, A.T. Role of BMP family members during kidney development. Int. J. Dev. Biol. 1999, 43, 405–411. [Google Scholar]
  78. Zhang, J.; Li, L. BMP signaling and stem cell regulation. Dev. Biol. 2005, 284, 1–11. [Google Scholar] [CrossRef]
  79. Qi, X.; Li, T.G.; Hao, J.; Hu, J.; Wang, J.; Simmons, H.; Miura, S.; Mishina, Y.; Zhao, G.Q. BMP4 supports self-renewal of embryonic stem cells by inhibiting mitogen-activated protein kinase pathways. Proc. Natl. Acad. Sci. USA 2004, 101, 6027–6032. [Google Scholar] [CrossRef]
  80. Ying, Q.L.; Nichols, J.; Chambers, I.; Smith, A. BMP induction of Id proteins suppresses differentiation and sustains embryonic stem cell self-renewal in collaboration with STAT3. Cell 2003, 115, 281–292. [Google Scholar] [CrossRef]
  81. Tomizawa, M.; Shinozaki, F.; Sugiyama, T.; Yamamoto, S.; Sueishi, M.; Yoshida, T. Activin A maintains pluripotency markers and proliferative potential of human induced pluripotent stem cells. Exp. Ther. Med. 2011, 2, 405–408. [Google Scholar] [CrossRef] [PubMed]
  82. Hayashi, Y.; Hsiao, E.C.; Sami, S.; Lancero, M.; Schlieve, C.R.; Nguyen, T.; Yano, K.; Nagahashi, A.; Ikeya, M.; Matsumoto, Y.; et al. BMP-SMAD-ID promotes reprogramming to pluripotency by inhibiting p16/INK4A-dependent senescence. Proc. Natl. Acad. Sci. USA 2016, 113, 13057–13062. [Google Scholar] [CrossRef] [PubMed]
  83. Ali, G.; Abdelalim, E.M. Directed differentiation of human pluripotent stem cells into epidermal keratinocyte-like cells. STAR Protoc. 2022, 3, 101613. [Google Scholar] [CrossRef] [PubMed]
  84. Ghorbani-Dalini, S.; Azarpira, N.; Sangtarash, M.H.; Soleimanpour-Lichaei, H.R.; Yaghobi, R.; Lorzadeh, S.; Sabet, A.; Sarshar, M.; Al-Abdullah, I.H. Optimization of activin-A: A breakthrough in differentiation of human induced pluripotent stem cell into definitive endoderm. 3 Biotech 2020, 10, 215. [Google Scholar] [CrossRef] [PubMed]
  85. Bogacheva, M.S.; Harjumaki, R.; Flander, E.; Taalas, A.; Bystriakova, M.A.; Yliperttula, M.; Xiang, X.; Leung, A.W.; Lou, Y.R. Differentiation of Human Pluripotent Stem Cells into Definitive Endoderm Cells in Various Flexible Three-Dimensional Cell Culture Systems: Possibilities and Limitations. Front. Cell Dev. Biol. 2021, 9, 726499. [Google Scholar] [CrossRef]
  86. Patsch, C.; Challet-Meylan, L.; Thoma, E.C.; Urich, E.; Heckel, T.; O’Sullivan, J.F.; Grainger, S.J.; Kapp, F.G.; Sun, L.; Christensen, K.; et al. Generation of vascular endothelial and smooth muscle cells from human pluripotent stem cells. Nat. Cell Biol. 2015, 17, 994–1003. [Google Scholar] [CrossRef]
  87. Kim, M.S.; Horst, A.; Blinka, S.; Stamm, K.; Mahnke, D.; Schuman, J.; Gundry, R.; Tomita-Mitchell, A.; Lough, J. Activin-A and Bmp4 levels modulate cell type specification during CHIR-induced cardiomyogenesis. PLoS ONE 2015, 10, e0118670. [Google Scholar] [CrossRef]
  88. Park, S.W.; Jun Koh, Y.; Jeon, J.; Cho, Y.H.; Jang, M.J.; Kang, Y.; Kim, M.J.; Choi, C.; Sook Cho, Y.; Chung, H.M.; et al. Efficient differentiation of human pluripotent stem cells into functional CD34+ progenitor cells by combined modulation of the MEK/ERK and BMP4 signaling pathways. Blood 2010, 116, 5762–5772. [Google Scholar] [CrossRef]
  89. Gai, H.; Leung, E.L.; Costantino, P.D.; Aguila, J.R.; Nguyen, D.M.; Fink, L.M.; Ward, D.C.; Ma, Y. Generation and characterization of functional cardiomyocytes using induced pluripotent stem cells derived from human fibroblasts. Cell Biol. Int. 2009, 33, 1184–1193. [Google Scholar] [CrossRef]
  90. Guzzo, R.M.; Drissi, H. Differentiation of Human Induced Pluripotent Stem Cells to Chondrocytes. Methods Mol. Biol. 2015, 1340, 79–95. [Google Scholar]
  91. Song, B.; Smink, A.M.; Jones, C.V.; Callaghan, J.M.; Firth, S.D.; Bernard, C.A.; Laslett, A.L.; Kerr, P.G.; Ricardo, S.D. The directed differentiation of human iPS cells into kidney podocytes. PLoS ONE 2012, 7, e46453. [Google Scholar] [CrossRef] [PubMed]
  92. Tanigawa, S.; Naganuma, H.; Kaku, Y.; Era, T.; Sakuma, T.; Yamamoto, T.; Taguchi, A.; Nishinakamura, R. Activin Is Superior to BMP7 for Efficient Maintenance of Human iPSC-Derived Nephron Progenitors. Stem Cell Rep. 2019, 13, 322–337. [Google Scholar] [CrossRef] [PubMed]
  93. Lichtner, B.; Knaus, P.; Lehrach, H.; Adjaye, J. BMP10 as a potent inducer of trophoblast differentiation in human embryonic and induced pluripotent stem cells. Biomaterials 2013, 34, 9789–9802. [Google Scholar] [CrossRef] [PubMed]
  94. Williams, I.M.; Wu, J.C. Generation of Endothelial Cells from Human Pluripotent Stem Cells. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 1317–1329. [Google Scholar] [CrossRef]
  95. Jansen, M.H.; van Vuurden, D.G.; Vandertop, W.P.; Kaspers, G.J. Diffuse intrinsic pontine gliomas: A systematic update on clinical trials and biology. Cancer Treat. Rev. 2012, 38, 27–35. [Google Scholar] [CrossRef]
  96. Buczkowicz, P.; Hoeman, C.; Rakopoulos, P.; Pajovic, S.; Letourneau, L.; Dzamba, M.; Morrison, A.; Lewis, P.; Bouffet, E.; Bartels, U.; et al. Genomic analysis of diffuse intrinsic pontine gliomas identifies three molecular subgroups and recurrent activating ACVR1 mutations. Nat. Genet. 2014, 46, 451–456. [Google Scholar] [CrossRef]
  97. Taylor, K.R.; Mackay, A.; Truffaux, N.; Butterfield, Y.; Morozova, O.; Philippe, C.; Castel, D.; Grasso, C.S.; Vinci, M.; Carvalho, D.; et al. Recurrent activating ACVR1 mutations in diffuse intrinsic pontine glioma. Nat. Genet. 2014, 46, 457–461. [Google Scholar] [CrossRef]
  98. D’Alessandro, L.C.; Al Turki, S.; Manickaraj, A.K.; Manase, D.; Mulder, B.J.; Bergin, L.; Rosenberg, H.C.; Mondal, T.; Gordon, E.; Lougheed, J.; et al. Exome sequencing identifies rare variants in multiple genes in atrioventricular septal defect. Genet. Med. 2016, 18, 189–198. [Google Scholar] [CrossRef]
  99. Gaussin, V.; Morley, G.E.; Cox, L.; Zwijsen, A.; Vance, K.M.; Emile, L.; Tian, Y.; Liu, J.; Hong, C.; Myers, D.; et al. Alk3/Bmpr1a receptor is required for development of the atrioventricular canal into valves and annulus fibrosus. Circ. Res. 2005, 97, 219–226. [Google Scholar] [CrossRef]
  100. Howe, J.R.; Bair, J.L.; Sayed, M.G.; Anderson, M.E.; Mitros, F.A.; Petersen, G.M.; Velculescu, V.E.; Traverso, G.; Vogelstein, B. Germline mutations of the gene encoding bone morphogenetic protein receptor 1A in juvenile polyposis. Nat. Genet. 2001, 28, 184–187. [Google Scholar] [CrossRef]
  101. Howe, J.R.; Roth, S.; Ringold, J.C.; Summers, R.W.; Jarvinen, H.J.; Sistonen, P.; Tomlinson, I.P.; Houlston, R.S.; Bevan, S.; Mitros, F.A.; et al. Mutations in the SMAD4/DPC4 gene in juvenile polyposis. Science 1998, 280, 1086–1088. [Google Scholar] [CrossRef] [PubMed]
  102. Demirhan, O.; Turkmen, S.; Schwabe, G.C.; Soyupak, S.; Akgul, E.; Tastemir, D.; Karahan, D.; Mundlos, S.; Lehmann, K. A homozygous BMPR1B mutation causes a new subtype of acromesomelic chondrodysplasia with genital anomalies. J. Med. Genet. 2005, 42, 314–317. [Google Scholar] [CrossRef] [PubMed]
  103. Ullah, A.; Umair, M.; Muhammad, D.; Bilal, M.; Lee, K.; Leal, S.M.; Ahmad, W. A novel homozygous variant in BMPR1B underlies acromesomelic dysplasia Hunter-Thompson type. Ann. Hum. Genet. 2018, 82, 129–134. [Google Scholar] [CrossRef] [PubMed]
  104. Shore, E.M.; Xu, M.; Feldman, G.J.; Fenstermacher, D.A.; Cho, T.J.; Choi, I.H.; Connor, J.M.; Delai, P.; Glaser, D.L.; LeMerrer, M.; et al. A recurrent mutation in the BMP type I receptor ACVR1 causes inherited and sporadic fibrodysplasia ossificans progressiva. Nat. Genet. 2006, 38, 525–527. [Google Scholar] [CrossRef]
  105. Pignolo, R.J.; Shore, E.M.; Kaplan, F.S. Fibrodysplasia ossificans progressiva: Clinical and genetic aspects. Orphanet J. Rare Dis. 2011, 6, 80. [Google Scholar] [CrossRef]
  106. Shore, E.M.; Kaplan, F.S. Inherited human diseases of heterotopic bone formation. Nat. Rev. Rheumatol. 2010, 6, 518–527. [Google Scholar] [CrossRef]
  107. Kou, S.; De Cunto, C.; Baujat, G.; Wentworth, K.L.; Grogan, D.R.; Brown, M.A.; Di Rocco, M.; Keen, R.; Al Mukaddam, M.; le Quan Sang, K.H.; et al. Patients with ACVR1(R206H) mutations have an increased prevalence of cardiac conduction abnormalities on electrocardiogram in a natural history study of Fibrodysplasia Ossificans Progressiva. Orphanet J. Rare Dis. 2020, 15, 193. [Google Scholar] [CrossRef]
  108. Marseglia, L.; D’Angelo, G.; Manti, S.; Manganaro, A.; Calabro, M.P.; Salpietro, C.; Gitto, E. Fibrodysplasia ossificans progressiva in a newborn with cardiac involvement. Pediatr. Int. 2015, 57, 719–721. [Google Scholar] [CrossRef]
  109. Ware, A.D.; Brewer, N.; Meyers, C.; Morris, C.; McCarthy, E.; Shore, E.M.; James, A.W. Differential Vascularity in Genetic and Nonhereditary Heterotopic Ossification. Int. J. Surg. Pathol. 2019, 27, 859–867. [Google Scholar] [CrossRef]
  110. el-Labban, N.G.; Hopper, C.; Barber, P. Ultrastructural finding of vascular degeneration in fibrodysplasia ossificans progressiva (FOP). J. Oral Pathol. Med. 1995, 24, 125–129. [Google Scholar] [CrossRef]
  111. Hegyi, L.; Gannon, F.H.; Glaser, D.L.; Shore, E.M.; Kaplan, F.S.; Shanahan, C.M. Stromal cells of fibrodysplasia ossificans progressiva lesions express smooth muscle lineage markers and the osteogenic transcription factor Runx2/Cbfa-1: Clues to a vascular origin of heterotopic ossification? J. Pathol. 2003, 201, 141–148. [Google Scholar] [CrossRef] [PubMed]
  112. Sanchez-Duffhues, G.; Williams, E.; Benderitter, P.; Orlova, V.; van Wijhe, M.; Garcia de Vinuesa, A.; Kerr, G.; Caradec, J.; Lodder, K.; de Boer, H.C.; et al. Development of Macrocycle Kinase Inhibitors for ALK2 Using Fibrodysplasia Ossificans Progressiva-Derived Endothelial Cells. JBMR Plus 2019, 3, e10230. [Google Scholar] [CrossRef] [PubMed]
  113. Wentworth, K.L.; Bigay, K.; Chan, T.V.; Ho, J.P.; Morales, B.M.; Connor, J.; Brooks, E.; Shahriar Salamat, M.; Sanchez, H.C.; Wool, G.; et al. Clinical-pathological correlations in three patients with fibrodysplasia ossificans progressiva. Bone 2018, 109, 104–110. [Google Scholar] [CrossRef]
  114. Hamasaki, M.; Hashizume, Y.; Yamada, Y.; Katayama, T.; Hohjoh, H.; Fusaki, N.; Nakashima, Y.; Furuya, H.; Haga, N.; Takami, Y.; et al. Pathogenic mutation of ALK2 inhibits induced pluripotent stem cell reprogramming and maintenance: Mechanisms of reprogramming and strategy for drug identification. Stem Cells 2012, 30, 2437–2449. [Google Scholar] [CrossRef] [PubMed]
  115. Matsumoto, Y.; Hayashi, Y.; Schlieve, C.R.; Ikeya, M.; Kim, H.; Nguyen, T.D.; Sami, S.; Baba, S.; Barruet, E.; Nasu, A.; et al. Induced pluripotent stem cells from patients with human fibrodysplasia ossificans progressiva show increased mineralization and cartilage formation. Orphanet J. Rare Dis. 2013, 8, 190. [Google Scholar] [CrossRef] [PubMed]
  116. Cai, J.; Orlova, V.V.; Cai, X.; Eekhoff, E.M.W.; Zhang, K.; Pei, D.; Pan, G.; Mummery, C.L.; Ten Dijke, P. Induced Pluripotent Stem Cells to Model Human Fibrodysplasia Ossificans Progressiva. Stem Cell Rep. 2015, 5, 963–970. [Google Scholar] [CrossRef]
  117. Orlova, V.V.; van den Hil, F.E.; Petrus-Reurer, S.; Drabsch, Y.; Ten Dijke, P.; Mummery, C.L. Generation, expansion and functional analysis of endothelial cells and pericytes derived from human pluripotent stem cells. Nat. Protoc. 2014, 9, 1514–1531. [Google Scholar] [CrossRef]
  118. Orlova, V.V.; Drabsch, Y.; Freund, C.; Petrus-Reurer, S.; van den Hil, F.E.; Muenthaisong, S.; Dijke, P.T.; Mummery, C.L. Functionality of endothelial cells and pericytes from human pluripotent stem cells demonstrated in cultured vascular plexus and zebrafish xenografts. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 177–186. [Google Scholar] [CrossRef]
  119. Sanchez-Duffhues, G.; Garcia de Vinuesa, A.; Ten Dijke, P. Endothelial-to-mesenchymal transition in cardiovascular diseases: Developmental signaling pathways gone awry. Dev. Dyn. 2018, 247, 492–508. [Google Scholar] [CrossRef]
  120. Medici, D.; Shore, E.M.; Lounev, V.Y.; Kaplan, F.S.; Kalluri, R.; Olsen, B.R. Conversion of vascular endothelial cells into multipotent stem-like cells. Nat. Med. 2010, 16, 1400–1406. [Google Scholar] [CrossRef]
  121. Barruet, E.; Morales, B.M.; Lwin, W.; White, M.P.; Theodoris, C.V.; Kim, H.; Urrutia, A.; Wong, S.A.; Srivastava, D.; Hsiao, E.C. The ACVR1 R206H mutation found in fibrodysplasia ossificans progressiva increases human induced pluripotent stem cell-derived endothelial cell formation and collagen production through BMP-mediated SMAD1/5/8 signaling. Stem Cell Res. Ther. 2016, 7, 115. [Google Scholar] [CrossRef] [PubMed]
  122. Lees-Shepard, J.B.; Yamamoto, M.; Biswas, A.A.; Stoessel, S.J.; Nicholas, S.E.; Cogswell, C.A.; Devarakonda, P.M.; Schneider, M.J., Jr.; Cummins, S.M.; Legendre, N.P.; et al. Activin-dependent signaling in fibro/adipogenic progenitors causes fibrodysplasia ossificans progressiva. Nat. Commun. 2018, 9, 471. [Google Scholar] [CrossRef] [PubMed]
  123. Hatsell, S.J.; Idone, V.; Wolken, D.M.; Huang, L.; Kim, H.J.; Wang, L.; Wen, X.; Nannuru, K.C.; Jimenez, J.; Xie, L.; et al. ACVR1R206H receptor mutation causes fibrodysplasia ossificans progressiva by imparting responsiveness to activin A. Sci. Transl. Med. 2015, 7, 303ra137. [Google Scholar] [CrossRef]
  124. Dey, D.; Bagarova, J.; Hatsell, S.J.; Armstrong, K.A.; Huang, L.; Ermann, J.; Vonner, A.J.; Shen, Y.; Mohedas, A.H.; Lee, A.; et al. Two tissue-resident progenitor lineages drive distinct phenotypes of heterotopic ossification. Sci. Transl. Med. 2016, 8, 366ra163. [Google Scholar] [CrossRef] [PubMed]
  125. Nakajima, T.; Shibata, M.; Nishio, M.; Nagata, S.; Alev, C.; Sakurai, H.; Toguchida, J.; Ikeya, M. Modeling human somite development and fibrodysplasia ossificans progressiva with induced pluripotent stem cells. Development 2018, 145, dev165431. [Google Scholar] [CrossRef] [PubMed]
  126. Barruet, E.; Garcia, S.M.; Wu, J.; Morales, B.M.; Tamaki, S.; Moody, T.; Pomerantz, J.H.; Hsiao, E.C. Modeling the ACVR1(R206H) mutation in human skeletal muscle stem cells. Elife 2021, 10, e66107. [Google Scholar] [CrossRef] [PubMed]
  127. Matsuo, K.; Lepinski, A.; Chavez, R.D.; Barruet, E.; Pereira, A.; Moody, T.A.; Ton, A.N.; Sharma, A.; Hellman, J.; Tomoda, K.; et al. ACVR1(R206H) extends inflammatory responses in human induced pluripotent stem cell-derived macrophages. Bone 2021, 153, 116129. [Google Scholar] [CrossRef]
  128. Maekawa, H.; Jin, Y.; Nishio, M.; Kawai, S.; Nagata, S.; Kamakura, T.; Yoshitomi, H.; Niwa, A.; Saito, M.K.; Matsuda, S.; et al. Recapitulation of pro-inflammatory signature of monocytes with ACVR1A mutation using FOP patient-derived iPSCs. Orphanet J. Rare Dis. 2022, 17, 364. [Google Scholar] [CrossRef]
  129. Yanagimachi, M.D.; Niwa, A.; Tanaka, T.; Honda-Ozaki, F.; Nishimoto, S.; Murata, Y.; Yasumi, T.; Ito, J.; Tomida, S.; Oshima, K.; et al. Robust and highly-efficient differentiation of functional monocytic cells from human pluripotent stem cells under serum- and feeder cell-free conditions. PLoS ONE 2013, 8, e59243. [Google Scholar]
  130. Hino, K.; Ikeya, M.; Horigome, K.; Matsumoto, Y.; Ebise, H.; Nishio, M.; Sekiguchi, K.; Shibata, M.; Nagata, S.; Matsuda, S.; et al. Neofunction of ACVR1 in fibrodysplasia ossificans progressiva. Proc. Natl. Acad. Sci. USA 2015, 112, 15438–15443. [Google Scholar] [CrossRef]
  131. Hino, K.; Horigome, K.; Nishio, M.; Komura, S.; Nagata, S.; Zhao, C.; Jin, Y.; Kawakami, K.; Yamada, Y.; Ohta, A.; et al. Activin-A enhances mTOR signaling to promote aberrant chondrogenesis in fibrodysplasia ossificans progressiva. J. Clin. Investig. 2017, 127, 3339–3352. [Google Scholar] [CrossRef] [PubMed]
  132. Hino, K.; Zhao, C.; Horigome, K.; Nishio, M.; Okanishi, Y.; Nagata, S.; Komura, S.; Yamada, Y.; Toguchida, J.; Ohta, A.; et al. An mTOR Signaling Modulator Suppressed Heterotopic Ossification of Fibrodysplasia Ossificans Progressiva. Stem Cell Rep. 2018, 11, 1106–1119. [Google Scholar] [CrossRef] [PubMed]
  133. Hildebrandt, S.; Kampfrath, B.; Fischer, K.; Hildebrand, L.; Haupt, J.; Stachelscheid, H.; Knaus, P. ActivinA Induced SMAD1/5 Signaling in an iPSC Derived EC Model of Fibrodysplasia Ossificans Progressiva (FOP) Can Be Rescued by the Drug Candidate Saracatinib. Stem Cell Rev. Rep. 2021, 17, 1039–1052. [Google Scholar] [CrossRef] [PubMed]
  134. Micha, D.; Voermans, E.; Eekhoff, M.E.W.; van Essen, H.W.; Zandieh-Doulabi, B.; Netelenbos, C.; Rustemeyer, T.; Sistermans, E.A.; Pals, G.; Bravenboer, N. Inhibition of TGFbeta signaling decreases osteogenic differentiation of fibrodysplasia ossificans progressiva fibroblasts in a novel in vitro model of the disease. Bone 2016, 84, 169–180. [Google Scholar] [CrossRef] [PubMed]
  135. Del Zotto, G.; Antonini, F.; Azzari, I.; Ortolani, C.; Tripodi, G.; Giacopelli, F.; Cappato, S.; Moretta, L.; Ravazzolo, R.; Bocciardi, R. Peripheral Blood Mononuclear Cell Immunophenotyping in Fibrodysplasia Ossificans Progressiva Patients: Evidence for Monocyte DNAM1 Up-regulation. Cytom. B Clin. Cytom. 2018, 94, 613–622. [Google Scholar] [CrossRef] [PubMed]
  136. Sanchez-Duffhues, G.; Mikkers, H.; de Jong, D.; Szuhai, K.; de Vries, T.J.; Freund, C.; Bravenboer, N.; van Es, R.J.J.; Netelenbos, J.C.; Goumans, M.J.; et al. Generation of Fibrodysplasia ossificans progressiva and control integration free iPSC lines from periodontal ligament fibroblasts. Stem Cell Res. 2019, 41, 101639. [Google Scholar] [CrossRef]
  137. Schoenmaker, T.; Mokry, M.; Micha, D.; Netelenbos, C.; Bravenboer, N.; Gilijamse, M.; Eekhoff, E.M.W.; de Vries, T.J. Activin-A Induces Early Differential Gene Expression Exclusively in Periodontal Ligament Fibroblasts from Fibrodysplasia Ossificans Progressiva Patients. Biomedicines 2021, 9, 629. [Google Scholar] [CrossRef]
  138. de Ruiter, R.D.; Wisse, L.E.; Schoenmaker, T.; Yaqub, M.; Sanchez-Duffhues, G.; Eekhoff, E.M.W.; Micha, D. TGF-Beta Induces Activin A Production in Dermal Fibroblasts Derived from Patients with Fibrodysplasia Ossificans Progressiva. Int. J. Mol. Sci. 2023, 24, 2299. [Google Scholar] [CrossRef]
  139. Evans, J.D.; Girerd, B.; Montani, D.; Wang, X.J.; Galie, N.; Austin, E.D.; Elliott, G.; Asano, K.; Grunig, E.; Yan, Y.; et al. BMPR2 mutations and survival in pulmonary arterial hypertension: An individual participant data meta-analysis. Lancet Respir. Med. 2016, 4, 129–137. [Google Scholar] [CrossRef]
  140. Morrell, N.W.; Aldred, M.A.; Chung, W.K.; Elliott, C.G.; Nichols, W.C.; Soubrier, F.; Trembath, R.C.; Loyd, J.E. Genetics and genomics of pulmonary arterial hypertension. Eur. Respir. J. 2019, 53, 1801899. [Google Scholar] [CrossRef]
  141. Machado, R.D.; Southgate, L.; Eichstaedt, C.A.; Aldred, M.A.; Austin, E.D.; Best, D.H.; Chung, W.K.; Benjamin, N.; Elliott, C.G.; Eyries, M.; et al. Pulmonary Arterial Hypertension: A Current Perspective on Established and Emerging Molecular Genetic Defects. Hum. Mutat. 2015, 36, 1113–1127. [Google Scholar] [CrossRef] [PubMed]
  142. Soubrier, F.; Chung, W.K.; Machado, R.; Grunig, E.; Aldred, M.; Geraci, M.; Loyd, J.E.; Elliott, C.G.; Trembath, R.C.; Newman, J.H.; et al. Genetics and genomics of pulmonary arterial hypertension. J. Am. Coll. Cardiol. 2013, 62, D13–D21. [Google Scholar] [CrossRef] [PubMed]
  143. Newman, J.H.; Trembath, R.C.; Morse, J.A.; Grunig, E.; Loyd, J.E.; Adnot, S.; Coccolo, F.; Ventura, C.; Phillips, J.A., 3rd; Knowles, J.A.; et al. Genetic basis of pulmonary arterial hypertension: Current understanding and future directions. J. Am. Coll. Cardiol. 2004, 43, 33S–39S. [Google Scholar] [CrossRef] [PubMed]
  144. Chida, A.; Shintani, M.; Nakayama, T.; Furutani, Y.; Hayama, E.; Inai, K.; Saji, T.; Nonoyama, S.; Nakanishi, T. Missense mutations of the BMPR1B (ALK6) gene in childhood idiopathic pulmonary arterial hypertension. Circ. J. 2012, 76, 1501–1508. [Google Scholar] [CrossRef]
  145. Frump, A.L.; Lowery, J.W.; Hamid, R.; Austin, E.D.; de Caestecker, M. Abnormal trafficking of endogenously expressed BMPR2 mutant allelic products in patients with heritable pulmonary arterial hypertension. PLoS ONE 2013, 8, e80319. [Google Scholar]
  146. Yang, Y.M.; Lane, K.B.; Sehgal, P.B. Subcellular mechanisms in pulmonary arterial hypertension: Combinatorial modalities that inhibit anterograde trafficking and cause bone morphogenetic protein receptor type 2 mislocalization. Pulm. Circ. 2013, 3, 533–550. [Google Scholar] [CrossRef]
  147. Machado, R.D.; Aldred, M.A.; James, V.; Harrison, R.E.; Patel, B.; Schwalbe, E.C.; Gruenig, E.; Janssen, B.; Koehler, R.; Seeger, W.; et al. Mutations of the TGF-beta type II receptor BMPR2 in pulmonary arterial hypertension. Hum. Mutat. 2006, 27, 121–132. [Google Scholar] [CrossRef]
  148. Machado, R.D.; Pauciulo, M.W.; Thomson, J.R.; Lane, K.B.; Morgan, N.V.; Wheeler, L.; Phillips, J.A., 3rd; Newman, J.; Williams, D.; Galie, N.; et al. BMPR2 haploinsufficiency as the inherited molecular mechanism for primary pulmonary hypertension. Am. J. Hum. Genet. 2001, 68, 92–102. [Google Scholar] [CrossRef]
  149. Newman, J.H.; Wheeler, L.; Lane, K.B.; Loyd, E.; Gaddipati, R.; Phillips, J.A., 3rd; Loyd, J.E. Mutation in the gene for bone morphogenetic protein receptor II as a cause of primary pulmonary hypertension in a large kindred. N. Engl. J. Med. 2001, 345, 319–324. [Google Scholar] [CrossRef]
  150. Cogan, J.D.; Pauciulo, M.W.; Batchman, A.P.; Prince, M.A.; Robbins, I.M.; Hedges, L.K.; Stanton, K.C.; Wheeler, L.A.; Phillips, J.A., 3rd; Loyd, J.E.; et al. High frequency of BMPR2 exonic deletions/duplications in familial pulmonary arterial hypertension. Am. J. Respir. Crit. Care Med. 2006, 174, 590–598. [Google Scholar] [CrossRef]
  151. Cogan, J.; Austin, E.; Hedges, L.; Womack, B.; West, J.; Loyd, J.; Hamid, R. Role of BMPR2 alternative splicing in heritable pulmonary arterial hypertension penetrance. Circulation 2012, 126, 1907–1916. [Google Scholar] [CrossRef]
  152. Seki, T.; Hong, K.H.; Oh, S.P. Nonoverlapping expression patterns of ALK1 and ALK5 reveal distinct roles of each receptor in vascular development. Lab. Investig. 2006, 86, 116–129. [Google Scholar] [CrossRef] [PubMed]
  153. Seki, T.; Yun, J.; Oh, S.P. Arterial endothelium-specific activin receptor-like kinase 1 expression suggests its role in arterialization and vascular remodeling. Circ. Res. 2003, 93, 682–689. [Google Scholar] [CrossRef] [PubMed]
  154. Alt, A.; Miguel-Romero, L.; Donderis, J.; Aristorena, M.; Blanco, F.J.; Round, A.; Rubio, V.; Bernabeu, C.; Marina, A. Structural and functional insights into endoglin ligand recognition and binding. PLoS ONE 2012, 7, e29948. [Google Scholar]
  155. Tazat, K.; Pomeraniec-Abudy, L.; Hector-Greene, M.; Szilagyi, S.S.; Sharma, S.; Cai, E.M.; Corona, A.L.; Ehrlich, M.; Blobe, G.C.; Henis, Y.I. ALK1 regulates the internalization of endoglin and the type III TGF-beta receptor. Mol. Biol. Cell 2021, 32, 605–621. [Google Scholar] [CrossRef] [PubMed]
  156. Sugden, W.W.; Meissner, R.; Aegerter-Wilmsen, T.; Tsaryk, R.; Leonard, E.V.; Bussmann, J.; Hamm, M.J.; Herzog, W.; Jin, Y.; Jakobsson, L.; et al. Endoglin controls blood vessel diameter through endothelial cell shape changes in response to haemodynamic cues. Nat. Cell Biol. 2017, 19, 653–665. [Google Scholar] [CrossRef] [PubMed]
  157. Seghers, L.; de Vries, M.R.; Pardali, E.; Hoefer, I.E.; Hierck, B.P.; ten Dijke, P.; Goumans, M.J.; Quax, P.H. Shear induced collateral artery growth modulated by endoglin but not by ALK1. J. Cell. Mol. Med. 2012, 16, 2440–2450. [Google Scholar] [CrossRef]
  158. 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]
  159. David, L.; Mallet, C.; Keramidas, M.; Lamande, N.; Gasc, J.M.; Dupuis-Girod, S.; Plauchu, H.; Feige, J.J.; Bailly, S. Bone morphogenetic protein-9 is a circulating vascular quiescence factor. Circ. Res. 2008, 102, 914–922. [Google Scholar] [CrossRef]
  160. Desroches-Castan, A.; Tillet, E.; Bouvard, C.; Bailly, S. BMP9 and BMP10: Two close vascular quiescence partners that stand out. Dev. Dyn. 2022, 251, 178–197. [Google Scholar] [CrossRef]
  161. Akla, N.; Viallard, C.; Popovic, N.; Lora Gil, C.; Sapieha, P.; Larrivee, B. BMP9 (Bone Morphogenetic Protein-9)/Alk1 (Activin-Like Kinase Receptor Type I) Signaling Prevents Hyperglycemia-Induced Vascular Permeability. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 1821–1836. [Google Scholar] [CrossRef] [PubMed]
  162. Long, L.; Ormiston, M.L.; Yang, X.; Southwood, M.; Graf, S.; Machado, R.D.; Mueller, M.; Kinzel, B.; Yung, L.M.; Wilkinson, J.M.; et al. Selective enhancement of endothelial BMPR-II with BMP9 reverses pulmonary arterial hypertension. Nat. Med. 2015, 21, 777–785. [Google Scholar] [CrossRef] [PubMed]
  163. Li, W.; Long, L.; Yang, X.; Tong, Z.; Southwood, M.; King, R.; Caruso, P.; Upton, P.D.; Yang, P.; Bocobo, G.A.; et al. Circulating BMP9 Protects the Pulmonary Endothelium during Inflammation-induced Lung Injury in Mice. Am. J. Respir. Crit. Care Med. 2021, 203, 1419–1430. [Google Scholar] [CrossRef] [PubMed]
  164. Ponomarev, L.C.; Ksiazkiewicz, J.; Staring, M.W.; Luttun, A.; Zwijsen, A. The BMP Pathway in Blood Vessel and Lymphatic Vessel Biology. Int. J. Mol. Sci. 2021, 22, 6364. [Google Scholar] [CrossRef]
  165. Gorelova, A.; Berman, M.; Al Ghouleh, I. Endothelial-to-Mesenchymal Transition in Pulmonary Arterial Hypertension. Antioxid. Redox Signal. 2021, 34, 891–914. [Google Scholar] [CrossRef]
  166. Yun, E.; Kook, Y.; Yoo, K.H.; Kim, K.I.; Lee, M.S.; Kim, J.; Lee, A. Endothelial to Mesenchymal Transition in Pulmonary Vascular Diseases. Biomedicines 2020, 8, 639. [Google Scholar] [CrossRef]
  167. Ihida-Stansbury, K.; McKean, D.M.; Lane, K.B.; Loyd, J.E.; Wheeler, L.A.; Morrell, N.W.; Jones, P.L. Tenascin-C is induced by mutated BMP type II receptors in familial forms of pulmonary arterial hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2006, 291, L694–L702. [Google Scholar] [CrossRef]
  168. Hiepen, C.; Jatzlau, J.; Hildebrandt, S.; Kampfrath, B.; Goktas, M.; Murgai, A.; Cuellar Camacho, J.L.; Haag, R.; Ruppert, C.; Sengle, G.; et al. BMPR2 acts as a gatekeeper to protect endothelial cells from increased TGFbeta responses and altered cell mechanics. PLoS Biol. 2019, 17, e3000557. [Google Scholar] [CrossRef]
  169. Thenappan, T.; Chan, S.Y.; Weir, E.K. Role of extracellular matrix in the pathogenesis of pulmonary arterial hypertension. Am. J. Physiol. Heart Circ. Physiol. 2018, 315, H1322–H1331. [Google Scholar] [CrossRef]
  170. Hiepen, C.; Jatzlau, J.; Knaus, P. Biomechanical stress provides a second hit in the establishment of BMP/TGFbeta-related vascular disorders. Cell Stress 2020, 4, 44–47. [Google Scholar] [CrossRef]
  171. Wang, L.; Moonen, J.R.; Cao, A.; Isobe, S.; Li, C.G.; Tojais, N.F.; Taylor, S.; Marciano, D.P.; Chen, P.I.; Gu, M.; et al. Dysregulated Smooth Muscle Cell BMPR2-ARRB2 Axis Causes Pulmonary Hypertension. Circ. Res. 2023, 132, 545–564. [Google Scholar] [CrossRef] [PubMed]
  172. Tada, Y.; Majka, S.; Carr, M.; Harral, J.; Crona, D.; Kuriyama, T.; West, J. Molecular effects of loss of BMPR2 signaling in smooth muscle in a transgenic mouse model of PAH. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007, 292, L1556–L1563. [Google Scholar] [CrossRef] [PubMed]
  173. Andruska, A.; Spiekerkoetter, E. Consequences of BMPR2 Deficiency in the Pulmonary Vasculature and Beyond: Contributions to Pulmonary Arterial Hypertension. Int. J. Mol. Sci. 2018, 19, 2499. [Google Scholar] [CrossRef] [PubMed]
  174. West, J.D.; Austin, E.D.; Gaskill, C.; Marriott, S.; Baskir, R.; Bilousova, G.; Jean, J.C.; Hemnes, A.R.; Menon, S.; Bloodworth, N.C.; et al. Identification of a common Wnt-associated genetic signature across multiple cell types in pulmonary arterial hypertension. Am. J. Physiol. Cell Physiol. 2014, 307, C415–C430. [Google Scholar] [CrossRef] [PubMed]
  175. Wu, X.H.; Ma, J.L.; Ding, D.; Ma, Y.J.; Wei, Y.P.; Jing, Z.C. Experimental animal models of pulmonary hypertension: Development and challenges. Anim. Model Exp. Med. 2022, 5, 207–216. [Google Scholar] [CrossRef]
  176. Jasinska-Stroschein, M. A review of genetically-driven rodent models of pulmonary hypertension. Vasc. Pharmacol. 2022, 144, 106970. [Google Scholar] [CrossRef]
  177. Spiekerkoetter, E.; Tian, X.; Cai, J.; Hopper, R.K.; Sudheendra, D.; Li, C.G.; El-Bizri, N.; Sawada, H.; Haghighat, R.; Chan, R.; et al. FK506 activates BMPR2, rescues endothelial dysfunction, and reverses pulmonary hypertension. J. Clin. Investig. 2013, 123, 3600–3613. [Google Scholar] [CrossRef]
  178. Yung, L.M.; Yang, P.; Joshi, S.; Augur, Z.M.; Kim, S.S.J.; Bocobo, G.A.; Dinter, T.; Troncone, L.; Chen, P.S.; McNeil, M.E.; et al. ACTRIIA-Fc rebalances activin/GDF versus BMP signaling in pulmonary hypertension. Sci. Transl. Med. 2020, 12, eaaz5660. [Google Scholar] [CrossRef]
  179. Yndestad, A.; Larsen, K.O.; Oie, E.; Ueland, T.; Smith, C.; Halvorsen, B.; Sjaastad, I.; Skjonsberg, O.H.; Pedersen, T.M.; Anfinsen, O.G.; et al. Elevated levels of activin A in clinical and experimental pulmonary hypertension. J. Appl. Physiol. 2009, 106, 1356–1364. [Google Scholar] [CrossRef]
  180. Rosa, S.; Praca, C.; Pitrez, P.R.; Gouveia, P.J.; Aranguren, X.L.; Ricotti, L.; Ferreira, L.S. Functional characterization of iPSC-derived arterial- and venous-like endothelial cells. Sci. Rep. 2019, 9, 3826. [Google Scholar] [CrossRef]
  181. Nguyen, J.; Lin, Y.Y.; Gerecht, S. The next generation of endothelial differentiation: Tissue-specific ECs. Cell Stem Cell 2021, 28, 1188–1204. [Google Scholar] [CrossRef] [PubMed]
  182. Hamid, R.; Yan, L. Induced Pluripotent Stem Cells in Pulmonary Arterial Hypertension. Am. J. Respir. Crit. Care Med. 2017, 195, 852–853. [Google Scholar] [CrossRef] [PubMed]
  183. Geti, I.; Ormiston, M.L.; Rouhani, F.; Toshner, M.; Movassagh, M.; Nichols, J.; Mansfield, W.; Southwood, M.; Bradley, A.; Rana, A.A.; et al. A practical and efficient cellular substrate for the generation of induced pluripotent stem cells from adults: Blood-derived endothelial progenitor cells. Stem Cells Transl. Med. 2012, 1, 855–865. [Google Scholar] [CrossRef] [PubMed]
  184. Gu, M.; Shao, N.Y.; Sa, S.; Li, D.; Termglinchan, V.; Ameen, M.; Karakikes, I.; Sosa, G.; Grubert, F.; Lee, J.; et al. Patient-Specific iPSC-Derived Endothelial Cells Uncover Pathways that Protect against Pulmonary Hypertension in BMPR2 Mutation Carriers. Cell Stem Cell 2017, 20, 490–504.e5. [Google Scholar] [CrossRef] [PubMed]
  185. Sa, S.; Gu, M.; Chappell, J.; Shao, N.Y.; Ameen, M.; Elliott, K.A.; Li, D.; Grubert, F.; Li, C.G.; Taylor, S.; et al. Induced Pluripotent Stem Cell Model of Pulmonary Arterial Hypertension Reveals Novel Gene Expression and Patient Specificity. Am. J. Respir. Crit. Care Med. 2017, 195, 930–941. [Google Scholar] [CrossRef] [PubMed]
  186. Sun, W.; Tang, Y.; Tai, Y.Y.; Handen, A.; Zhao, J.; Speyer, G.; Al Aaraj, Y.; Watson, A.; Romanelli, M.E.; Sembrat, J.; et al. SCUBE1 Controls BMPR2-Relevant Pulmonary Endothelial Function: Implications for Diagnostic Marker Development in Pulmonary Arterial Hypertension. JACC Basic Transl. Sci. 2020, 5, 1073–1092. [Google Scholar] [CrossRef]
  187. Gu, M.; Donato, M.; Guo, M.; Wary, N.; Miao, Y.; Mao, S.; Saito, T.; Otsuki, S.; Wang, L.; Harper, R.L.; et al. iPSC-endothelial cell phenotypic drug screening and in silico analyses identify tyrphostin-AG1296 for pulmonary arterial hypertension. Sci. Transl. Med. 2021, 13, eaba6480. [Google Scholar] [CrossRef]
  188. Li, Y.; Li, Y.; Liu, Q.; Wang, A. Tyrphostin AG1296, a platelet-derived growth factor receptor inhibitor, induces apoptosis, and reduces viability and migration of PLX4032-resistant melanoma cells. OncoTargets Ther. 2015, 8, 1043–1051. [Google Scholar] [CrossRef]
  189. Lasota, M.; Bentke-Imiolek, A.; Skrzypek, K.; Bobrowska, J.; Jagusiak, A.; Bryniarska-Kubiak, N.; Zagajewski, J.; Kot, M.; Szydlak, R.; Lekka, M.; et al. Small-molecule inhibitor-tyrphostin AG1296 regulates proliferation, survival and migration of rhabdomyosarcoma cells. J. Physiol. Pharmacol. 2021, 72, 881–893. [Google Scholar]
  190. Ueda, S.; Ikeda, H.; Mizuki, M.; Ishiko, J.; Matsumura, I.; Tanaka, H.; Shibayama, H.; Sugahara, H.; Takai, E.; Zhang, X.; et al. Constitutive activation of c-kit by the juxtamembrane but not the catalytic domain mutations is inhibited selectively by tyrosine kinase inhibitors STI571 and AG1296. Int. J. Hematol. 2002, 76, 427–435. [Google Scholar] [CrossRef]
  191. Kiskin, F.N.; Chang, C.H.; Huang, C.J.Z.; Kwieder, B.; Cheung, C.; Dunmore, B.J.; Serrano, F.; Sinha, S.; Morrell, N.W.; Rana, A.A. Contributions of BMPR2 Mutations and Extrinsic Factors to Cellular Phenotypes of Pulmonary Arterial Hypertension Revealed by Induced Pluripotent Stem Cell Modeling. Am. J. Respir. Crit. Care Med. 2018, 198, 271–275. [Google Scholar] [CrossRef] [PubMed]
  192. Morrell, N.W. Finding the needle in the haystack: BMP9 and 10 emerge from the genome in pulmonary arterial hypertension. Eur. Respir. J. 2019, 53, 1900078. [Google Scholar] [CrossRef] [PubMed]
  193. Hu, W.-P.; Zhang, S.-J.; Ding, Y.-J.; Fang, J.; Zhou, L.; Xie, S.-M.; Ge, X.; Fu, L.-J.; Li, Q.-Y.; Qu, J.-M.; et al. Progesterone is an Inducement of Heritable Pulmonary Arterial Hypertension with BMPR2 Mutation. bioRxiv 2023. [Google Scholar] [CrossRef]
  194. Chen, X.; Talati, M.; Fessel, J.P.; Hemnes, A.R.; Gladson, S.; French, J.; Shay, S.; Trammell, A.; Phillips, J.A.; Hamid, R.; et al. Estrogen Metabolite 16alpha-Hydroxyestrone Exacerbates Bone Morphogenetic Protein Receptor Type II-Associated Pulmonary Arterial Hypertension Through MicroRNA-29-Mediated Modulation of Cellular Metabolism. Circulation 2016, 133, 82–97. [Google Scholar] [CrossRef]
  195. Singla, S.; Machado, R.F. The imitation game in pulmonary arterial hypertension. Sex, bone morphogenetic protein receptor, and the estrogen paradox. Am. J. Respir. Crit. Care Med. 2015, 191, 612–613. [Google Scholar] [CrossRef] [PubMed]
  196. Liu, D.; Wu, W.H.; Mao, Y.M.; Yuan, P.; Zhang, R.; Ju, F.L.; Jing, Z.C. BMPR2 mutations influence phenotype more obviously in male patients with pulmonary arterial hypertension. Circ. Cardiovasc. Genet. 2012, 5, 511–518. [Google Scholar] [CrossRef]
  197. Devendran, A.; Kar, S.; Bailey, R.; Trivieri, M.G. The Role of Bone Morphogenetic Protein Receptor Type 2 (BMPR2) and the Prospects of Utilizing Induced Pluripotent Stem Cells (iPSCs) in Pulmonary Arterial Hypertension Disease Modeling. Cells 2022, 11, 3823. [Google Scholar] [CrossRef] [PubMed]
  198. Huang, W.C.; Ke, M.W.; Cheng, C.C.; Chiou, S.H.; Wann, S.R.; Shu, C.W.; Chiou, K.R.; Tseng, C.J.; Pan, H.W.; Mar, G.Y.; et al. Therapeutic Benefits of Induced Pluripotent Stem Cells in Monocrotaline-Induced Pulmonary Arterial Hypertension. PLoS ONE 2016, 11, e0142476. [Google Scholar] [CrossRef]
  199. Oh, S.; Jung, J.H.; Ahn, K.J.; Jang, A.Y.; Byun, K.; Yang, P.C.; Chung, W.J. Stem Cell and Exosome Therapy in Pulmonary Hypertension. Korean Circ. J. 2022, 52, 110–122. [Google Scholar] [CrossRef]
  200. McAllister, K.A.; Grogg, K.M.; Johnson, D.W.; Gallione, C.J.; Baldwin, M.A.; Jackson, C.E.; Helmbold, E.A.; Markel, D.S.; McKinnon, W.C.; Murrell, J.; et al. Endoglin, a TGF-beta binding protein of endothelial cells, is the gene for hereditary haemorrhagic telangiectasia type 1. Nat. Genet. 1994, 8, 345–351. [Google Scholar] [CrossRef]
  201. Gallione, C.J.; Pasyk, K.A.; Boon, L.M.; Lennon, F.; Johnson, D.W.; Helmbold, E.A.; Markel, D.S.; Vikkula, M.; Mulliken, J.B.; Warman, M.L.; et al. A gene for familial venous malformations maps to chromosome 9p in a second large kindred. J. Med. Genet. 1995, 32, 197–199. [Google Scholar] [CrossRef]
  202. Johnson, D.W.; Berg, J.N.; Baldwin, M.A.; Gallione, C.J.; Marondel, I.; Yoon, S.J.; Stenzel, T.T.; Speer, M.; Pericak-Vance, M.A.; Diamond, A.; et al. Mutations in the activin receptor-like kinase 1 gene in hereditary haemorrhagic telangiectasia type 2. Nat. Genet. 1996, 13, 189–195. [Google Scholar] [CrossRef]
  203. Abdalla, S.A.; Cymerman, U.; Johnson, R.M.; Deber, C.M.; Letarte, M. Disease-associated mutations in conserved residues of ALK-1 kinase domain. Eur. J. Hum. Genet. 2003, 11, 279–287. [Google Scholar] [CrossRef]
  204. McAllister, K.A.; Baldwin, M.A.; Thukkani, A.K.; Gallione, C.J.; Berg, J.N.; Porteous, M.E.; Guttmacher, A.E.; Marchuk, D.A. Six novel mutations in the endoglin gene in hereditary hemorrhagic telangiectasia type 1 suggest a dominant-negative effect of receptor function. Hum. Mol. Genet. 1995, 4, 1983–1985. [Google Scholar] [CrossRef]
  205. Wooderchak-Donahue, W.L.; McDonald, J.; O’Fallon, B.; Upton, P.D.; Li, W.; Roman, B.L.; Young, S.; Plant, P.; Fulop, G.T.; Langa, C.; et al. BMP9 mutations cause a vascular-anomaly syndrome with phenotypic overlap with hereditary hemorrhagic telangiectasia. Am. J. Hum. Genet. 2013, 93, 530–537. [Google Scholar] [CrossRef] [PubMed]
  206. Dupuis-Girod, S.; Bailly, S.; Plauchu, H. Hereditary hemorrhagic telangiectasia: From molecular biology to patient care. J. Thromb. Haemost. 2010, 8, 1447–1456. [Google Scholar] [CrossRef] [PubMed]
  207. Shovlin, C.L. Hereditary haemorrhagic telangiectasia: Pathophysiology, diagnosis and treatment. Blood Rev. 2010, 24, 203–219. [Google Scholar] [CrossRef] [PubMed]
  208. Thalgott, J.H.; Dos-Santos-Luis, D.; Hosman, A.E.; Martin, S.; Lamande, N.; Bracquart, D.; Srun, S.; Galaris, G.; de Boer, H.C.; Tual-Chalot, S.; et al. Decreased Expression of Vascular Endothelial Growth Factor Receptor 1 Contributes to the Pathogenesis of Hereditary Hemorrhagic Telangiectasia Type 2. Circulation 2018, 138, 2698–2712. [Google Scholar] [CrossRef] [PubMed]
  209. Carvalho, R.L.; Jonker, L.; Goumans, M.J.; Larsson, J.; Bouwman, P.; Karlsson, S.; Dijke, P.T.; Arthur, H.M.; Mummery, C.L. Defective paracrine signalling by TGFbeta in yolk sac vasculature of endoglin mutant mice: A paradigm for hereditary haemorrhagic telangiectasia. Development 2004, 131, 6237–6247. [Google Scholar] [CrossRef] [PubMed]
  210. Lebrin, F.; Srun, S.; Raymond, K.; Martin, S.; van den Brink, S.; Freitas, C.; Breant, C.; Mathivet, T.; Larrivee, B.; Thomas, J.L.; et al. Thalidomide stimulates vessel maturation and reduces epistaxis in individuals with hereditary hemorrhagic telangiectasia. Nat. Med. 2010, 16, 420–428. [Google Scholar] [CrossRef]
  211. Gerhardt, H. VEGF and endothelial guidance in angiogenic sprouting. Organogenesis 2008, 4, 241–246. [Google Scholar] [CrossRef] [PubMed]
  212. Larrivee, B.; Prahst, C.; Gordon, E.; del Toro, R.; Mathivet, T.; Duarte, A.; Simons, M.; Eichmann, A. ALK1 signaling inhibits angiogenesis by cooperating with the Notch pathway. Dev. Cell 2012, 22, 489–500. [Google Scholar] [CrossRef] [PubMed]
  213. Choi, E.J.; Kim, Y.H.; Choe, S.W.; Tak, Y.G.; Garrido-Martin, E.M.; Chang, M.; Lee, Y.J.; Oh, S.P. Enhanced responses to angiogenic cues underlie the pathogenesis of hereditary hemorrhagic telangiectasia 2. PLoS ONE 2013, 8, e63138. [Google Scholar] [CrossRef] [PubMed]
  214. Vion, A.C.; Alt, S.; Klaus-Bergmann, A.; Szymborska, A.; Zheng, T.; Perovic, T.; Hammoutene, A.; Oliveira, M.B.; Bartels-Klein, E.; Hollfinger, I.; et al. Primary cilia sensitize endothelial cells to BMP and prevent excessive vascular regression. J. Cell Biol. 2018, 217, 1651–1665. [Google Scholar] [CrossRef]
  215. Corti, P.; Young, S.; Chen, C.Y.; Patrick, M.J.; Rochon, E.R.; Pekkan, K.; Roman, B.L. Interaction between alk1 and blood flow in the development of arteriovenous malformations. Development 2011, 138, 1573–1582. [Google Scholar] [CrossRef]
  216. Hiepen, C.; Mendez, P.L.; Knaus, P. It Takes Two to Tango: Endothelial TGFbeta/BMP Signaling Crosstalk with Mechanobiology. Cells 2020, 9, 1965. [Google Scholar] [CrossRef]
  217. Eisa-Beygi, S.; Burrows, P.E.; Link, B.A. Endothelial cilia dysfunction in pathogenesis of hereditary hemorrhagic telangiectasia. Front. Cell Dev. Biol. 2022, 10, 1037453. [Google Scholar] [CrossRef]
  218. Riss, D.; Burian, M.; Wolf, A.; Kranebitter, V.; Kaider, A.; Arnoldner, C. Intranasal submucosal bevacizumab for epistaxis in hereditary hemorrhagic telangiectasia: A double-blind, randomized, placebo-controlled trial. Head Neck 2015, 37, 783–787. [Google Scholar] [CrossRef]
  219. Robert, F.; Desroches-Castan, A.; Bailly, S.; Dupuis-Girod, S.; Feige, J.J. Future treatments for hereditary hemorrhagic telangiectasia. Orphanet. J. Rare Dis. 2020, 15, 4. [Google Scholar] [CrossRef]
  220. Kim, Y.H.; Kim, M.J.; Choe, S.W.; Sprecher, D.; Lee, Y.J.; Oh, S.P. Selective effects of oral antiangiogenic tyrosine kinase inhibitors on an animal model of hereditary hemorrhagic telangiectasia. J. Thromb. Haemost. 2017, 15, 1095–1102. [Google Scholar] [CrossRef]
  221. Freund, C.; Davis, R.P.; Gkatzis, K.; Ward-van Oostwaard, D.; Mummery, C.L. The first reported generation of human induced pluripotent stem cells (iPS cells) and iPS cell-derived cardiomyocytes in the Netherlands. Neth. Heart J. 2010, 18, 51–54. [Google Scholar] [PubMed]
  222. Bouma, M.J.; Orlova, V.; van den Hil, F.E.; Mager, H.J.; Baas, F.; de Knijff, P.; Mummery, C.L.; Mikkers, H.; Freund, C. Generation and genetic repair of 2 iPSC clones from a patient bearing a heterozygous c.1120del18 mutation in the ACVRL1 gene leading to Hereditary Hemorrhagic Telangiectasia (HHT) type 2. Stem Cell Res. 2020, 46, 101786. [Google Scholar] [CrossRef] [PubMed]
  223. Zhou, F.; Zhao, X.; Liu, X.; Liu, Y.; Ma, F.; Liu, B.; Yang, J. Autologous correction in patient induced pluripotent stem cell-endothelial cells to identify a novel pathogenic mutation of hereditary hemorrhagic telangiectasia. Pulm. Circ. 2020, 10, 2045894019885357. [Google Scholar] [CrossRef]
  224. Fernandez, L.A.; Sanz-Rodriguez, F.; Zarrabeitia, R.; Perez-Molino, A.; Hebbel, R.P.; Nguyen, J.; Bernabeu, C.; Botella, L.M. Blood outgrowth endothelial cells from Hereditary Haemorrhagic Telangiectasia patients reveal abnormalities compatible with vascular lesions. Cardiovasc. Res. 2005, 68, 235–248. [Google Scholar] [CrossRef] [PubMed]
  225. Xiang-Tischhauser, L.; Bette, M.; Rusche, J.R.; Roth, K.; Kasahara, N.; Stuck, B.A.; Bakowsky, U.; Wartenberg, M.; Sauer, H.; Geisthoff, U.W.; et al. Generation of a Syngeneic Heterozygous ACVRL1((wt/mut)) Knockout iPS Cell Line for the in Vitro Study of HHT2-Associated Angiogenesis. Cells 2023, 12, 1600. [Google Scholar] [CrossRef]
  226. Orlova, V.V.; Nahon, D.M.; Cochrane, A.; Cao, X.; Freund, C.; van den Hil, F.; Westermann, C.J.J.; Snijder, R.J.; Ploos van Amstel, J.K.; Ten Dijke, P.; et al. Vascular defects associated with hereditary hemorrhagic telangiectasia revealed in patient-derived isogenic iPSCs in 3D vessels on chip. Stem Cell Rep. 2022, 17, 1536–1545. [Google Scholar] [CrossRef]
  227. Girerd, B.; Montani, D.; Coulet, F.; Sztrymf, B.; Yaici, A.; Jais, X.; Tregouet, D.; Reis, A.; Drouin-Garraud, V.; Fraisse, A.; et al. Clinical outcomes of pulmonary arterial hypertension in patients carrying an ACVRL1 (ALK1) mutation. Am. J. Respir. Crit. Care Med. 2010, 181, 851–861. [Google Scholar] [CrossRef]
  228. Walsh, L.J.; Collins, C.; Ibrahim, H.; Kerins, D.M.; Brady, A.P.; TM, O.C. Pulmonary arterial hypertension in hereditary hemorrhagic telangiectasia associated with ACVRL1 mutation: A case report. J. Med. Case Rep. 2022, 16, 99. [Google Scholar] [CrossRef]
  229. Kularatne, M.; Eyries, M.; Savale, L.; Humbert, M.; Montani, D. Isolated Pulmonary Arteriovenous Malformations Associated With BMPR2 Pathogenic Variants. Chest 2023, 164, e23–e26. [Google Scholar] [CrossRef]
  230. Harrison, R.E.; Flanagan, J.A.; Sankelo, M.; Abdalla, S.A.; Rowell, J.; Machado, R.D.; Elliott, C.G.; Robbins, I.M.; Olschewski, H.; McLaughlin, V.; et al. Molecular and functional analysis identifies ALK-1 as the predominant cause of pulmonary hypertension related to hereditary haemorrhagic telangiectasia. J. Med. Genet. 2003, 40, 865–871. [Google Scholar] [CrossRef]
  231. Hildebrand, L.; Rossbach, B.; Kuhnen, P.; Gossen, M.; Kurtz, A.; Reinke, P.; Seemann, P.; Stachelscheid, H. Generation of integration free induced pluripotent stem cells from fibrodysplasia ossificans progressiva (FOP) patients from urine samples. Stem Cell Res. 2016, 16, 54–58. [Google Scholar] [CrossRef] [PubMed]
  232. Usman, A.; Haase, A.; Merkert, S.; Gohring, G.; Hansmann, G.; Gall, H.; Schermuly, R.; Martin, U.; Olmer, R. Generation of pulmonary arterial hypertension patient-specific induced pluripotent stem cell lines from three unrelated patients with a heterozygous missense mutation in exon 12, a heterozygous in-frame deletion in exon 3 and a missense mutation in exon 11 of the BMPR2 gene. Stem Cell Res. 2021, 55, 102488. [Google Scholar] [PubMed]
  233. Ura, H.; Togi, S.; Iwata, Y.; Ozaki, M.; Niida, Y. Establishment of a human induced pluripotent stem cell line, KMUGMCi001-A, from a patient bearing a heterozygous c.772 + 3_772 + 4dup mutation in the ACVRL1 gene leading Telangiectasia, hereditary hemorrhagic, type 2 (HHT2). Stem Cell Res. 2022, 61, 102743. [Google Scholar] [CrossRef] [PubMed]
  234. Morales, I.A.; Boghdady, C.M.; Campbell, B.E.; Moraes, C. Integrating mechanical sensor readouts into organ-on-a-chip platforms. Front. Bioeng. Biotechnol. 2022, 10, 1060895. [Google Scholar] [CrossRef]
  235. Zhang, Y.S.; Aleman, J.; Shin, S.R.; Kilic, T.; Kim, D.; Mousavi Shaegh, S.A.; Massa, S.; Riahi, R.; Chae, S.; Hu, N.; et al. Multisensor-integrated organs-on-chips platform for automated and continual in situ monitoring of organoid behaviors. Proc. Natl. Acad. Sci. USA 2017, 114, E2293–E2302. [Google Scholar] [CrossRef]
  236. Fanizza, F.; Campanile, M.; Forloni, G.; Giordano, C.; Albani, D. Induced pluripotent stem cell-based organ-on-a-chip as personalized drug screening tools: A focus on neurodegenerative disorders. J. Tissue Eng. 2022, 13, 20417314221095339. [Google Scholar] [CrossRef]
  237. Hockney, S.; Parker, J.; Turner, J.E.; Todd, X.; Todryk, S.; Gieling, R.G.; Hilgen, G.; Simoes, D.C.M.; Pal, D. Next generation organoid engineering to replace animals in cancer drug testing. Biochem. Pharmacol. 2023, 213, 115586. [Google Scholar] [CrossRef]
  238. Fuhr, A.; Kurtz, A.; Hiepen, C.; Müller, S. Organoids as Miniature Twins—Challenges for Comparability and Need for Data Standardization and Access. Organoids 2022, 1, 28–36. [Google Scholar] [CrossRef]
  239. Yin, X.; Mead, B.E.; Safaee, H.; Langer, R.; Karp, J.M.; Levy, O. Engineering Stem Cell Organoids. Cell Stem Cell 2016, 18, 25–38. [Google Scholar] [CrossRef]
  240. Wang, P.; Wang, Y.; Qin, J. Multi-organ microphysiological system: A new paradigm for COVID-19 research. Organs Chip 2023, 5, 100029. [Google Scholar] [CrossRef]
  241. Zommiti, M.; Connil, N.; Tahrioui, A.; Groboillot, A.; Barbey, C.; Konto-Ghiorghi, Y.; Lesouhaitier, O.; Chevalier, S.; Feuilloley, M.G.J. Organs-on-Chips Platforms Are Everywhere: A Zoom on Biomedical Investigation. Bioengineering 2022, 9, 646. [Google Scholar] [CrossRef] [PubMed]
  242. Koenig, L.; Ramme, A.P.; Faust, D.; Mayer, M.; Flotke, T.; Gerhartl, A.; Brachner, A.; Neuhaus, W.; Appelt-Menzel, A.; Metzger, M.; et al. A Human Stem Cell-Derived Brain-Liver Chip for Assessing Blood-Brain-Barrier Permeation of Pharmaceutical Drugs. Cells 2022, 11, 3295. [Google Scholar] [CrossRef] [PubMed]
  243. Whelan, I.T.; Burdis, R.; Shahreza, S.; Moeendarbary, E.; Hoey, D.A.; Kelly, D.J. A microphysiological model of bone development and regeneration. Biofabrication 2023, 15, 034103. [Google Scholar] [CrossRef]
  244. Knowles, H.J.; Chanalaris, A.; Koutsikouni, A.; Cribbs, A.P.; Grover, L.M.; Hulley, P.A. Mature primary human osteocytes in mini organotypic cultures secrete FGF23 and PTH1-34-regulated sclerostin. Front. Endocrinol. 2023, 14, 1167734. [Google Scholar] [CrossRef] [PubMed]
  245. Knudson, A.G., Jr. Mutation and cancer: Statistical study of retinoblastoma. Proc. Natl. Acad. Sci. USA 1971, 68, 820–823. [Google Scholar] [CrossRef]
  246. Yang, P.; Yu, P.B. In Search of the Second Hit in Pulmonary Arterial Hypertension. Circ. Res. 2019, 124, 6–8. [Google Scholar] [CrossRef]
  247. Eichstaedt, C.A.; Song, J.; Benjamin, N.; Harutyunova, S.; Fischer, C.; Grunig, E.; Hinderhofer, K. EIF2AK4 mutation as “second hit” in hereditary pulmonary arterial hypertension. Respir. Res. 2016, 17, 141. [Google Scholar] [CrossRef] [PubMed]
  248. Viales, R.R.; Eichstaedt, C.A.; Ehlken, N.; Fischer, C.; Lichtblau, M.; Grunig, E.; Hinderhofer, K. Mutation in BMPR2 Promoter: A ‘Second Hit’ for Manifestation of Pulmonary Arterial Hypertension? PLoS ONE 2015, 10, e0133042. [Google Scholar] [CrossRef]
  249. Bernabeu, C.; Bayrak-Toydemir, P.; McDonald, J.; Letarte, M. Potential Second-Hits in Hereditary Hemorrhagic Telangiectasia. J. Clin. Med. 2020, 9, 3571. [Google Scholar] [CrossRef]
  250. Jaliawala, H.A.; Parmar, M.; Summers, K.; Bernardo, R.J. A second hit? Pulmonary arterial hypertension, BMPR2 mutation, and exposure to prescription amphetamines. Pulm. Circ. 2022, 12, e12053. [Google Scholar] [CrossRef]
  251. Vengethasamy, L.; Hautefort, A.; Tielemans, B.; Belge, C.; Perros, F.; Verleden, S.; Fadel, E.; Van Raemdonck, D.; Delcroix, M.; Quarck, R. BMPRII influences the response of pulmonary microvascular endothelial cells to inflammatory mediators. Pflugers Arch. 2016, 468, 1969–1983. [Google Scholar] [CrossRef]
  252. Hagen, M.; Fagan, K.; Steudel, W.; Carr, M.; Lane, K.; Rodman, D.M.; West, J. Interaction of interleukin-6 and the BMP pathway in pulmonary smooth muscle. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007, 292, L1473–L1479. [Google Scholar] [CrossRef] [PubMed]
  253. Arai, K.Y.; Ono, M.; Kudo, C.; Fujioka, A.; Okamura, R.; Nomura, Y.; Nishiyama, T. IL-1beta stimulates activin betaA mRNA expression in human skin fibroblasts through the MAPK pathways, the nuclear factor-kappaB pathway, and prostaglandin E2. Endocrinology 2011, 152, 3779–3790. [Google Scholar] [CrossRef] [PubMed]
  254. Hildebrand, L.; Gaber, T.; Kuhnen, P.; Morhart, R.; Unterborsch, H.; Schomburg, L.; Seemann, P. Trace element and cytokine concentrations in patients with Fibrodysplasia Ossificans Progressiva (FOP): A case control study. J. Trace Elem. Med. Biol. 2017, 39, 186–192. [Google Scholar] [CrossRef] [PubMed]
  255. Hwang, C.; Pagani, C.A.; Das, N.; Marini, S.; Huber, A.K.; Xie, L.; Jimenez, J.; Brydges, S.; Lim, W.K.; Nannuru, K.C.; et al. Activin A does not drive post-traumatic heterotopic ossification. Bone 2020, 138, 115473. [Google Scholar] [CrossRef] [PubMed]
  256. Davies, R.J.; Holmes, A.M.; Deighton, J.; Long, L.; Yang, X.; Barker, L.; Walker, C.; Budd, D.C.; Upton, P.D.; Morrell, N.W. BMP type II receptor deficiency confers resistance to growth inhibition by TGF-beta in pulmonary artery smooth muscle cells: Role of proinflammatory cytokines. Am. J. Physiol. Lung Cell. Mol. Physiol. 2012, 302, L604–L615. [Google Scholar] [CrossRef] [PubMed]
  257. Soon, E.; Crosby, A.; Southwood, M.; Yang, P.; Tajsic, T.; Toshner, M.; Appleby, S.; Shanahan, C.M.; Bloch, K.D.; Pepke-Zaba, J.; et al. Bone morphogenetic protein receptor type II deficiency and increased inflammatory cytokine production. A gateway to pulmonary arterial hypertension. Am. J. Respir. Crit. Care Med. 2015, 192, 859–872. [Google Scholar] [CrossRef]
  258. Furuya, Y.; Satoh, T.; Kuwana, M. Interleukin-6 as a potential therapeutic target for pulmonary arterial hypertension. Int. J. Rheumatol. 2010, 2010, 720305. [Google Scholar] [CrossRef]
  259. Zhang, R.; Han, Z.; Degos, V.; Shen, F.; Choi, E.J.; Sun, Z.; Kang, S.; Wong, M.; Zhu, W.; Zhan, L.; et al. Persistent infiltration and pro-inflammatory differentiation of monocytes cause unresolved inflammation in brain arteriovenous malformation. Angiogenesis 2016, 19, 451–461. [Google Scholar] [CrossRef]
  260. Haupt, J.; Stanley, A.; McLeod, C.M.; Cosgrove, B.D.; Culbert, A.L.; Wang, L.; Mourkioti, F.; Mauck, R.L.; Shore, E.M. ACVR1(R206H) FOP mutation alters mechanosensing and tissue stiffness during heterotopic ossification. Mol. Biol. Cell 2019, 30, 17–29. [Google Scholar] [CrossRef]
  261. Park, J.S.; Chu, J.S.; Tsou, A.D.; Diop, R.; Tang, Z.; Wang, A.; Li, S. The effect of matrix stiffness on the differentiation of mesenchymal stem cells in response to TGF-beta. Biomaterials 2011, 32, 3921–3930. [Google Scholar] [CrossRef] [PubMed]
  262. Harburger, D.S.; Calderwood, D.A. Integrin signalling at a glance. J. Cell Sci. 2009, 122 Pt 2, 159–163. [Google Scholar] [CrossRef] [PubMed]
  263. Ashe, H.L. Modulation of BMP signalling by integrins. Biochem. Soc. Trans. 2016, 44, 1465–1473. [Google Scholar] [CrossRef] [PubMed]
  264. Johnson, J.A.; Hemnes, A.R.; Perrien, D.S.; Schuster, M.; Robinson, L.J.; Gladson, S.; Loibner, H.; Bai, S.; Blackwell, T.R.; Tada, Y.; et al. Cytoskeletal defects in Bmpr2-associated pulmonary arterial hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2012, 302, L474–L484. [Google Scholar] [CrossRef]
  265. Fernandez, L.A.; Sanz-Rodriguez, F.; Blanco, F.J.; Bernabeu, C.; Botella, L.M. Hereditary hemorrhagic telangiectasia, a vascular dysplasia affecting the TGF-beta signaling pathway. Clin. Med. Res. 2006, 4, 66–78. [Google Scholar] [CrossRef]
  266. Melchionna, R.; Trono, P.; Tocci, A.; Nistico, P. Actin Cytoskeleton and Regulation of TGFbeta Signaling: Exploring Their Links. Biomolecules 2021, 11, 336. [Google Scholar] [CrossRef]
  267. Ingber, D.E. Mechanobiology and diseases of mechanotransduction. Ann. Med. 2003, 35, 564–577. [Google Scholar] [CrossRef]
  268. Zuela-Sopilniak, N.; Lammerding, J. Can’t handle the stress? Mechanobiology and disease. Trends Mol. Med. 2022, 28, 710–725. [Google Scholar] [CrossRef]
  269. Abdelilah-Seyfried, S.; Iruela-Arispe, M.L.; Penninger, J.M.; Tournier-Lasserve, E.; Vikkula, M.; Cleaver, O. Recalibrating vascular malformations and mechanotransduction by pharmacological intervention. J. Clin. Investig. 2022, 132, e160227. [Google Scholar] [CrossRef]
  270. Tan, W.; Madhavan, K.; Hunter, K.S.; Park, D.; Stenmark, K.R. Vascular stiffening in pulmonary hypertension: Cause or consequence? (2013 Grover Conference series). Pulm. Circ. 2014, 4, 560–580. [Google Scholar] [CrossRef]
  271. Deng, H.; Min, E.; Baeyens, N.; Coon, B.G.; Hu, R.; Zhuang, Z.W.; Chen, M.; Huang, B.; Afolabi, T.; Zarkada, G.; et al. Activation of Smad2/3 signaling by low fluid shear stress mediates artery inward remodeling. Proc. Natl. Acad. Sci. USA 2021, 118, e2105339118. [Google Scholar] [CrossRef]
  272. Kopf, J.; Petersen, A.; Duda, G.N.; Knaus, P. BMP2 and mechanical loading cooperatively regulate immediate early signalling events in the BMP pathway. BMC Biol. 2012, 10, 37. [Google Scholar] [CrossRef]
  273. Papaioannou, T.G.; Stefanadis, C. Vascular wall shear stress: Basic principles and methods. Hellenic. J. Cardiol. 2005, 46, 9–15. [Google Scholar] [PubMed]
  274. Munoz, C.J.; Lucas, A.; Williams, A.T.; Cabrales, P. A Review on Microvascular Hemodynamics: The Control of Blood Flow Distribution and Tissue Oxygenation. Crit. Care Clin. 2020, 36, 293–305. [Google Scholar] [CrossRef] [PubMed]
  275. Campinho, P.; Vilfan, A.; Vermot, J. Blood Flow Forces in Shaping the Vascular System: A Focus on Endothelial Cell Behavior. Front. Physiol. 2020, 11, 552. [Google Scholar] [CrossRef] [PubMed]
  276. Katt, M.E.; Xu, Z.S.; Gerecht, S.; Searson, P.C. Human Brain Microvascular Endothelial Cells Derived from the BC1 iPS Cell Line Exhibit a Blood-Brain Barrier Phenotype. PLoS ONE 2016, 11, e0152105. [Google Scholar] [CrossRef] [PubMed]
  277. Linville, R.M.; DeStefano, J.G.; Sklar, M.B.; Xu, Z.; Farrell, A.M.; Bogorad, M.I.; Chu, C.; Walczak, P.; Cheng, L.; Mahairaki, V.; et al. Human iPSC-derived blood-brain barrier microvessels: Validation of barrier function and endothelial cell behavior. Biomaterials 2019, 190-191, 24–37. [Google Scholar] [CrossRef]
  278. Zhang, Y.E. Non-Smad Signaling Pathways of the TGF-beta Family. Cold Spring Harb. Perspect. Biol. 2017, 9, a022129. [Google Scholar] [CrossRef]
  279. Byfield, S.D.; Roberts, A.B. Lateral signaling enhances TGF-beta response complexity. Trends Cell Biol. 2004, 14, 107–111. [Google Scholar] [CrossRef]
  280. Wadman, M. FDA no longer has to require animal testing for new drugs. Science 2023, 379, 127–128. [Google Scholar] [CrossRef]
  281. Middelkamp, H.H.T.; Verboven, A.H.A.; De Sa Vivas, A.G.; Schoenmaker, C.; Klein Gunnewiek, T.M.; Passier, R.; Albers, C.A.; t Hoen, P.A.C.; Nadif Kasri, N.; van der Meer, A.D. Cell type-specific changes in transcriptomic profiles of endothelial cells, iPSC-derived neurons and astrocytes cultured on microfluidic chips. Sci. Rep. 2021, 11, 2281. [Google Scholar] [CrossRef] [PubMed]
  282. Discher, D.E.; Janmey, P.; Wang, Y.L. Tissue cells feel and respond to the stiffness of their substrate. Science 2005, 310, 1139–1143. [Google Scholar] [CrossRef] [PubMed]
  283. Cao, H.; Zhou, Q.; Liu, C.; Zhang, Y.; Xie, M.; Qiao, W.; Dong, N. Substrate stiffness regulates differentiation of induced pluripotent stem cells into heart valve endothelial cells. Acta Biomater. 2022, 143, 115–126. [Google Scholar] [CrossRef] [PubMed]
  284. Heo, J.H.; Kang, D.; Seo, S.J.; Jin, Y. Engineering the Extracellular Matrix for Organoid Culture. Int. J. Stem Cells 2022, 15, 60–69. [Google Scholar] [CrossRef]
  285. Sedlmeier, G.; Sleeman, J.P. Extracellular regulation of BMP signaling: Welcome to the matrix. Biochem. Soc. Trans. 2017, 45, 173–181. [Google Scholar] [CrossRef]
  286. Xiang, Y.; Miller, K.; Guan, J.; Kiratitanaporn, W.; Tang, M.; Chen, S. 3D bioprinting of complex tissues in vitro: State-of-the-art and future perspectives. Arch. Toxicol. 2022, 96, 691–710. [Google Scholar] [CrossRef]
  287. Gjorevski, N.; Sachs, N.; Manfrin, A.; Giger, S.; Bragina, M.E.; Ordonez-Moran, P.; Clevers, H.; Lutolf, M.P. Designer matrices for intestinal stem cell and organoid culture. Nature 2016, 539, 560–564. [Google Scholar] [CrossRef]
Figure 2. iPSC generation and cell type differentiation in rare BMP diseases. So far, iPSC lines have been successfully generated for three rare BMP diseases: Fibrodysplasia ossificans progressiva (FOP), Hereditary pulmonary arterial hypertension (HPAH), and Hereditary haemorrhagic telangiectasia (HHT). Various sources of somatic tissue have been explored for this purpose. Reprogramming efficiency may be affected by the BMP gene mutation, and small molecules, like LDN-193189, that aim to restore normal BMP signalling can be utilized. BMPs are potent inducers of mesoderm. Therefore, a combination of BMP agonists and Wnt antagonists (such as CHIR-99021) is employed to induce mesodermal differentiation. In order to generate vascular cells, ALK4/5/7 inhibition is often necessary in a subsequent step.
Figure 2. iPSC generation and cell type differentiation in rare BMP diseases. So far, iPSC lines have been successfully generated for three rare BMP diseases: Fibrodysplasia ossificans progressiva (FOP), Hereditary pulmonary arterial hypertension (HPAH), and Hereditary haemorrhagic telangiectasia (HHT). Various sources of somatic tissue have been explored for this purpose. Reprogramming efficiency may be affected by the BMP gene mutation, and small molecules, like LDN-193189, that aim to restore normal BMP signalling can be utilized. BMPs are potent inducers of mesoderm. Therefore, a combination of BMP agonists and Wnt antagonists (such as CHIR-99021) is employed to induce mesodermal differentiation. In order to generate vascular cells, ALK4/5/7 inhibition is often necessary in a subsequent step.
Cells 12 02200 g002
Figure 4. Current and Future Development of iPSC-based technology in BMP rare diseases. Since the inception of the first iPSC-based cell models, significant progress has been made in developing more intricate and robust systems that accurately replicate diseased tissues and organs. Of particular interest in monogenic diseases is the utilization of gene editing tools such as CRISPR/Cas9 to create isogenic cell lines. While the first culturing approaches opted for 2 dimensional methods, more complex three dimensional culture conditions, such as embryoid bodies (EBs) and organoids, are emerging. This can be combined with co-culture of different cell types differentiated from the same parental iPSC line. Culturing those cells under conditions that expose them to microenvironmental cues, such as hypoxia, biomechanics, and inflammation, allows for effective and complex disease modelling. These systems have proven to be valuable platforms for drug screening and in-depth mechanistic investigations. Further advances, such as construction of complex cellular grafts or therapeutic extracellular vesicles (EVs) derived from iPSCs, hold the potential to revolutionize tissue engineering and iPSC-based cell therapies, presenting therapeutic alternatives for patients affected by rare BMP diseases.
Figure 4. Current and Future Development of iPSC-based technology in BMP rare diseases. Since the inception of the first iPSC-based cell models, significant progress has been made in developing more intricate and robust systems that accurately replicate diseased tissues and organs. Of particular interest in monogenic diseases is the utilization of gene editing tools such as CRISPR/Cas9 to create isogenic cell lines. While the first culturing approaches opted for 2 dimensional methods, more complex three dimensional culture conditions, such as embryoid bodies (EBs) and organoids, are emerging. This can be combined with co-culture of different cell types differentiated from the same parental iPSC line. Culturing those cells under conditions that expose them to microenvironmental cues, such as hypoxia, biomechanics, and inflammation, allows for effective and complex disease modelling. These systems have proven to be valuable platforms for drug screening and in-depth mechanistic investigations. Further advances, such as construction of complex cellular grafts or therapeutic extracellular vesicles (EVs) derived from iPSCs, hold the potential to revolutionize tissue engineering and iPSC-based cell therapies, presenting therapeutic alternatives for patients affected by rare BMP diseases.
Cells 12 02200 g004
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

Sánchez-Duffhues, G.; Hiepen, C. Human iPSCs as Model Systems for BMP-Related Rare Diseases. Cells 2023, 12, 2200. https://doi.org/10.3390/cells12172200

AMA Style

Sánchez-Duffhues G, Hiepen C. Human iPSCs as Model Systems for BMP-Related Rare Diseases. Cells. 2023; 12(17):2200. https://doi.org/10.3390/cells12172200

Chicago/Turabian Style

Sánchez-Duffhues, Gonzalo, and Christian Hiepen. 2023. "Human iPSCs as Model Systems for BMP-Related Rare Diseases" Cells 12, no. 17: 2200. https://doi.org/10.3390/cells12172200

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